Next Article in Journal
Band Gap Engineering in Quadruple-Layered Sillén–Aurivillius Perovskite Oxychlorides Bi7Fe2Ti2O17X (X = Cl, Br, I) for Enhanced Photocatalytic Performance
Previous Article in Journal
Photocatalytic Activity and Stability of Organically Modified Layered Perovskite-like Titanates HLnTiO4 (Ln = La, Nd) in the Reaction of Hydrogen Evolution from Aqueous Methanol
Previous Article in Special Issue
Discovery and Heterologous Expression of Unspecific Peroxygenases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Laccase Catalysis in Bond Formation and Breakage: A Review

1
CAS Key Laboratory of Biobased Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, No. 189 Song Ling Road, Qingdao 266101, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
Shandong Energy Institute, Qingdao 266101, China
4
Qingdao New Energy Shandong Laboratory, Qingdao 266101, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(4), 750; https://doi.org/10.3390/catal13040750
Submission received: 9 March 2023 / Revised: 8 April 2023 / Accepted: 11 April 2023 / Published: 14 April 2023
(This article belongs to the Special Issue Advances in Biocatalysis and Enzyme Engineering)

Abstract

:
Laccase belongs to the superfamily of multicopper oxidases and has been widely investigated in recent decades. Due to its mild and efficient oxidation of substrates, laccase has been successfully applied in organic catalytic synthesis, the degradation of harmful substances, and other green catalytic fields. Nevertheless, there are few reports on the green catalysis with laccase. This review focuses on reporting and collating some of the latest interesting laccase-catalyzed bond formation and breakage research. This is discussed with a focus on the effects of the medium system on the laccase-catalyzed reaction, as well as the formation and the breakage of C–N, C–C, and C–O bonds catalyzed by laccase. It provides abundant references and novel insights for furthering the industrial applications of laccase.

1. Introduction

Laccase (EC 1.10.3.2) is a copper-containing polyphenol oxidase that belongs to the blue copper oxidase (MCO) family [1,2]. It was first discovered in the Japanese lacquer tree Rhus Vernicifera [3,4]. Subsequently, laccases were found in different plant species [5], microbes [6,7], and animals [8]. There have been more investigations on microbial laccases than on animal and plant laccases, for which there have been relatively few. Microbial laccases are divided into fungal laccases and bacterial laccases. Bacterial laccase mainly plays a role in melanin production, spore wall defense, morphological change, and copper ion detoxification [9,10]. Fungal laccase is mainly related to pigment generation, plant disease, and lignin degradation [11,12]. Plant laccase is closely related to lignin biosynthesis [13]. At the same time, the primary function of animal laccase protein is to control the ossification of the epidermis [14].
Laccase is a glycoprotein with a molecular mass ranging from 50 to 140 kDa. Their amino acid sequence can span from 220 to 800 amino acids and may contain three cupredoxin-like domains. These domains bind copper centers involved in intermolecular electron transfer reactions and constitute the catalytic core of laccases (Figure 1) [15]. The active copper center of laccase generally contains four copper ions: a type I copper ion (T1-Cu), a type II copper ion (T2-Cu), and two type III copper ions (T3-Cu). T1-Cu is a mononuclear center that can gain electrons from the substrate and then transfer them to the trinuclear cluster (TNC), and the oxidation of the substrate occurs there. T2-Cu is a single-electron acceptor, whereas T3-Cu forms coupled ion pairs and is a double-electron acceptor. T2-Cu and T3-Cu together form a trinuclear cluster (TNC). Oxygen accepts four electrons and four protons to form water, which joins the bulk solvent [16,17]. Laccases can perform the single-electron oxidation of the substrate without using hydrogen peroxide while reducing molecular oxygen to water; therefore, they have a surprisingly broad substrate spectrum and can oxidize simple diphenols, polyphenols, diamines, and aromatic amines. The optimum temperature and pH of laccases depend on the enzyme source and substrate properties; those ranges are, respectively, from 20 °C to 75 °C and 3 to 8 (Table 1).
As an oxidase, the ability of laccase to oxidize the substrate is directly related to its redox potential (E0) [33,34]. E0 is an important characteristic of the catalytic oxidation capacity of laccase, which is the energy required for laccase to capture an electron from a reducing substrate. The E0 is critical to the reactivity of laccase and the overall reaction characteristics. Laccase can directly oxidize substrates with low E0, whereas some mediators are needed to assist laccase in oxidizing substrates with high E0. Adding a mediator not only effectively improves the reaction efficiency of laccase but also expands the scope of the substrate. For instance, with the mediator’s help, laccase can oxidize nonphenolic structures with high E0 and is used in pulp bleaching [35]. Recently, laccase and the laccase–mediator system (LMS) have received extensive attention in green catalysis, such as synthesizing complex organic compounds, the selective modification of natural products, and the degradation of harmful substances [36,37,38,39]. For example, the C–N bond breakage of amines catalyzed by laccase is essential for synthesizing amino acids and nucleosides [40,41]. Existing research indicates that using Pleurotus ostreatus laccase and its natural mediator (syringaldehyde) to catalyze C–C bond breakage results in the removal of up to 100% and 85% of BPA at concentrations of 0.44 and 0.88 mmol/L in wastewater within 1 h [42].
Furthermore, laccase-catalyzed C–N, C–C, and C–O bond formation can be widely employed in the green synthesis of some functional polymers, environmental remediation, and other areas. In contrast, there are few reviews on laccase catalysis at present. This review starts with the application of the laccase–mediator system; it also focuses on the analysis of the laccase-catalyzed formation of small-molecule compounds through C–N, C–O, and C–C bonds to obtain functional phenol polymers and important chemical pharmaceutical skeletons, as well as the laccase-catalyzed breakage of the C–N, C–O, and C–C bonds of macromolecular refractory organic compounds to form organic molecules organics. Furthermore, this review summarizes and forecasts the prospects of the laccase-catalyzed formation and breakage of C–N, C–O, and C–C bonds, aiming to provide rich reference value and novel ideas for expanding and developing the multifunctional applications of laccase in the field of biotechnology.

2. The Effect of the Mediator System on Laccase Catalysis

Currently, some problems still need to be solved urgently to directly apply laccase to industrial production. For instance: (1) Numerous substrates cannot directly bind to laccase specifically. (2) With laccase, it is difficult to oxidize nonphenolic compounds with high E0 (E0 > 1.3 V) due to its low E0 (E0 < 0.8 V). Thus, the development of laccase in industries such as lignin degradation and bio-bleaching is limited [43,44]. In order to reduce the oxidation potential of substrates and improve the oxidation efficiency, some mediators can be used as an intermediate substrate for laccase to form new intermediate states to transport electrons [41,45]. These mediators are compounds with low molecular mass and low E0, such as 2,2′-Azino-bis-(3-ethylbenzothiazoline-sulphonate) (ABTS) and 2,2,6,6-Tetramethyl-1-piperidinylox (TEMPO), which can easily gain and lose electrons. They can form highly active and stable intermediates under the action of laccase and act on the substrate to is oxidized. In the LMS, laccase first oxidizes the mediator into a free radical. The oxidized mediator rapidly applies to compounds above the E0 of laccase and to those polymers that cannot directly access the laccase active center [46,47,48] (Figure 2).
Laccase mediators are usually divided into artificial and natural mediators (Figure 3) [49]. Due to their high efficiency and inexpensive availability, artificial mediators are widely used in lignin degradation, polycyclic aromatic hydrocarbon (PAH) oxidation, and dye decolorization. Common artificial mediators include ABTS, TEMPO, and 1-hydroxy-benzotriazole (HBT) (Figure 3B) [50,51,52]. Three mechanisms have been proposed for the function of mediators in the LMS: (1) hydrogen atom transfer (HAT), (2) electron transfer (ET), and (3) the ionic mechanism (IM). ABTS was the first mediator found to promote the laccase-catalyzed oxidation of nonphenolic lignin. The action mechanism of ABTS belongs to the electron transfer mechanism (ET), which undergoes two stages (Figure 4): forming an ABTS+· cationic radical and slowly oxidizing to ABTS2+. ABTS2+ with higher reduction potential (but not ABTS+·) performs a more critical function in the laccase–ABTS system, which mediates the oxidation of nonphenolic lignin substrates [53,54,55]. The HAT mechanism, which is generally the oxidation mechanism mediated by the N-OH type mediator, uses a form of nitryl (>N-O·) to perform oxidation, such as the HBT system. Meanwhile, the purpose of the ion mechanism (IM) is mainly to form an ammonium oxide ion (>N=O·) through the nitryl group (>N-O·) to carry out the oxidation, such as the TEMPO system [39,56].
Artificial mediators have potential applications in the areas of lignin degradation and polycyclic aromatic hydrocarbon (PAH) oxidation for dye decolorization, but some disadvantages limit their use. For instance: (1) poor stability, (2) potential toxicity [57], and (3) difficulty in regeneration when the molar ratio of mediators to substrates is as high as 40:1 [53,58]. Compared to artificial mediators, natural mediators have more economic value because they are readily obtained, environmentally friendly, and reproducible (Figure 3A) [35,53]. Some fungal metabolites and lignin derivatives could be used as natural mediators of laccase, including but not limited to vanillin, acetyl vanillin, acetosyringone, syringaldehyde, 2,4,6-trimethyl phenol, and p-coumaric acid [59]. Taking the laccase p-coumaric acid system as an example, it can remove 95% anthracene (80% with HBT) and benzoin anthracene within 24 h [57].
Besides indirectly assisting laccase-catalyzing substrates, mediators show synergism with each other, and the degradation efficiency increases with the increase in mediator concentration [60]. For example, the complex mediator system composed of laccase, ABTS, and HBT can oxidize phenanthrene with only one intermediary phase with a degradation rate that can be increased by 30–40% compared with a single-mediator system (such as the ABTS system or HBT system) [61]. Therefore, with intensive research on and development of the LMS, the biocatalytic substrates of laccase can be further expanded.

3. Laccase-Catalyzed C–N, C–C, and C–O Bond Breakage

Laccase-catalyzed C–N, C–C, and C–O bond breakage has been used in preparing multifunctional polymer materials, the purification of water pollution, organic synthesis, and other fields. It is similar to the catabolic process in vivo, in which laccase mediates the oxidative degradation of large organic compounds to form small molecules [62,63,64,65]. The laccase-TEMPO system can catalyze C–N bond breakage to deprotect and provide amino donors for organic synthesis. In this chapter, the breakage of the C–N, C–C, and C–O bonds will be introduced from three perspectives.

3.1. C–N Bond Breakage

Due to the abundance of amines in organic molecules, the employment of amino-protecting groups is commonplace in many synthetic schemes, such as carbamates, amides, and sulfonamides [66]. The chemical method is generally used to remove the amino-protecting groups and has the disadvantages of high price, harsh reaction conditions, and severe pollution [66,67,68,69]. The N-para-methoxyphenyl (N-PMP) and N-benzyl group could be removed in an aqueous solution with the oxidative fracture of the C–N bond catalysis by laccase [41].

3.1.1. N-PMP Removal

The para-methoxyphenyl (PMP) group is being increasingly used as a nitrogen-protecting group for amines [70]. In most cases, ceric ammonium nitrate (CAN) has been used for deprotection, which is expensive and highly toxic. It requires increasingly cost-effective, environmentally friendly, and scalable deprotection procedures to make these processes commercially viable. Recently, a novel enzymatic method for the oxidative deprotection of p-methoxyphenyl (PMP)-protected amines has been reported [71]. Using laccase AB (laccase from Agaricus bisporus) under mildly acidic conditions results in the highest efficiency (with the highest yield reaching 89%) when the cosolvent is DMSO and pH = 3 is used. Further screening of the substrate showed that laccase had activity on substrates 1–7 (Scheme 1, Table 2). Due to the reaction conditions being green and mild with a high catalytic efficiency, laccase has excellent potential as a conventional PMP removal biocatalyst in organic synthesis.

3.1.2. N-Benzyl Removal

N-Benzyl groups are typically cleaved using reductive methods, often employing palladium-catalyzed hydrogenolysis [72], which is accompanied by high cost, poor reaction conditions, and severe pollution. Reports have indicated that laccase from Trametes versicolor and TEMPO catalyzes the debenzylation of n-benzyl-1-phenylethylamine to α-phenylethylamine at 30 °C, pH 5, and O2 (Scheme 2, Table 3) [40]. This methodology could be successfully applied over different aliphatic, cyclic, and aromatic amines affording deprotected derivatives of 1a–7a smoothly and immaculately, with no undesired by-products or oxidative transformations (Conversion > 97%). In contrast to the high cost, poor reaction conditions, and severe pollution of palladium-catalyzed hydrogenolysis, laccase-catalyzed N-benzyl removal is accompanied by mild reaction conditions, pollution-free, and easy extraction of the final derivatives.
Although this reaction occurred efficiently for N-benzylated secondary amines (Scheme 2, Table 3), it did not modify O-benzylated alcohols or N-protected tertiary amines (Scheme 3, Table 4). It is speculated that laccase has a particular selectivity for N-benzyl, which effectively removes N-benzyl from the secondary amine and has no activity for O-benzylated amine. Compared to the secondary amine, the steric hindrance of the tertiary amine of the N-benzyl group is more significant, which results in the low efficiency in the benzyl removal of tertiary amines by laccase. In addition, the laccases that catalyze secondary N-benzyl removal are the fungal laccases with the highest E0. Tertiary amines and o-benzylated amines with higher dissociation energies are not catalyzed because the E0 of laccases is insufficient. Modifying laccase to magnify the binding pocket of laccase and increase the E0 of laccase is necessary to catalyze N-benzylated tertiary amines and O-benzylated alcohols with laccase [73,74,75].

3.2. C–C Bond Breakage

As an important organic chemical raw material, bisphenol A (BPA) is extensively involved in manufacturing polycarbonate and epoxy resin plastics [76,77]. It causes imbalances and disturbances in the endocrine system of organisms and even induces a risk of cancer when it accumulates in wild organisms and humans [78]. BPA can be effectively removed from wastewater using adsorption, Fenton oxidation, electrochemistry, photodegradation, and biofilm filtration [79,80,81]. However, these conventional methods struggle to eliminate trace BPA and easily cause secondary damage to the ecological environment. The degradation rates of BPA can be up to 100% catalyzed by laccase-hydroxy benzotriazole (HBT) within 4 h, which can cause the effective removal of BPA from industrial wastewater [82].
The intermediate products of BPA catalyzed by laccase include macromolecular polymerization and small-molecule oxidative decomposition, which are mainly affected by the source of laccase and the type of media [37,83]. Laccase can catalyze BPA to form phenoxy radical intermediates and covalently couple outside the enzymatic reaction site to form BPA oligomers and polymers [84,85]. Although in the LMS, the mediator as an electron shuttle prompts laccase to catalyze the cleavage of the C–C bond of oxidized BPA, forming a variety of small-molecule oxidative decomposition products [42,86,87] (Scheme 4). Furthermore, the LMS can also catalyze the breakage of C–C bonds of estrogen, antibiotics, polycyclic aromatic hydrocarbons, and so on to effectively degrade organic pollutants [88]. For instance, the LMS with HBT can effectively degrade polyethylene and nylon-66 by catalyzing the breakage of C–C bonds. In previous experiments, after 3 days of treatment with the LMS, the weight average molecular mass (Mw) of polyethylene decreased from 242,000 to 28,300, and that of nylon-66 from 79,300 to 14,700. LMS also decreased the polydispersity (weight average molecular mass/number average molecular mass, Mw/Mn) of polyethylene and nylon-66. Furthermore, these reductions in elongation, tensile strength, and molecular mass were accompanied with the morphological disintegration of the polyethylene and nylon-66 membranes [89,90,91].
Although laccase is effective in wastewater treatment, it is challenging to separate, purify, and recover free laccase due to its high price, which limits its practical application in wastewater treatment [92,93]. The recovery and reuse of laccase benefits from immobilized enzyme technology, which can significantly improve laccase’s stability and catalytic efficiency [94,95,96]. In one study, amino-functionalized magnetic nanoparticles were attached to laccase nanoflowers, and it was found that the degradation rate of BPA was up to 100% within 5 min, with a prolonged storage period and excellent recycling performance [97]. Therefore, this technology is expected to improve the removal efficiency of bisphenol A by laccase and guarantee the mass implementation of this enzyme in sewage treatment [98].

3.3. C–O Bond Breakage

Triclosan has been widely used in detergents, skin creams, and other dairy products as a broad-spectrum antibacterial agent. However, when the concentration of triclosan in water is higher than a specific range, it will lead to the death of fish and other organisms [99]. Triclosan can be effectively removed from wastewater using photochemical degradation and oxidation. Nevertheless, secondary damage to the ecological environment is caused by these conventional methods. Reports have shown that laccase can effectively remove and convert triclosan in water by catalyzing its C–O bond break to produce small-molecule chemicals. In these reports, the removal efficiency of triclosan catalyzed by the laccase–HBT system was about 80% (the HBT concentration was 1.0 mmol/L) [100,101].
In the laccase catalytic system, the triclosan removal efficiency can be improved by adding HBT. Without HBT, the primary mechanism of laccase-catalyzed oxidation of triclosan is the formation of oligomers, such as a dimer, trimer, and tetramer, through a free-radical-mediated coupling reaction. In the laccase–HBT system, triclosan conversion primarily involves generating 2,4-dichlorophenol and 3-chlorophenol through ether bond cleavage [102,103] (Scheme 5).
The mechanism of laccase-catalyzed C–N, C–C, and C–O bond breakage is similar to that of anabolic metabolism in vivo. First, the mediators are oxidized into an ionic type with high redox potential and oxygen is reduced to water by laccase in the presence of the mediators [104]. Then, the ionic mediators attack the substrate’s C–N, C–C, and C–O bonds, generating an unstable oxidation intermediate containing double bonds. Finally, the oxidation state intermediates are oxidized to the product in the presence of H+, reducing oxygen to water (Figure 5). The laccase-catalyzed breakage of C–N, C–C, and C–O bonds can be used in preparing multifunctional polymeric materials, water pollution purification, and organic synthesis. Among these, laccase-catalyzed C–N bond breakage can be applied to remove the amino-protecting group in chemical synthesis. This type of bond breakage as well as that of C–O can be harnessed in sewage treatment, environmental remediation, and other fields [105,106]. In conclusion, the laccase-catalyzed breakage of C–N, C–C, and C–O bonds has essential applications in preparing multifunctional polymeric materials, water pollution purification, and organic synthesis [102].

4. Laccase-Catalyzed C–N, C–C, C–O Bond Formation

The formation of C–C, C–N, and C–O bonds catalyzed by laccase has crucial application potential for preparing multifunctional polymer materials and the organic synthesis of amino acid derivatives [107]. It is similar to the anabolic process in vivo, which catalyzes the homologous/heterologous oxidative coupling of low-molecular-weight organic compounds from polymeric products [108]. For example, laccase from Trametes versicolor can catalyze the C–C homologous coupling of free radicals to prepare biaryl compounds, with the highest yield of the product being 85% [109]. In this chapter, the formation of the C–N, C–C, and C–O bonds will be introduced from three perspectives.

4.1. C–N Bond Formation

A single low-molecular-weight organic compound can be catalyzed by laccase to form multiple homologous isomers through a mild, green, and efficient process that can be widely used in the green synthesis of some essential antibacterial drugs [110]. It has been reported that 2-amino-4,6-dimethyl-3-oxo-phenoxazine-1,9-dicarboxylic acid (actinocin) was prepared with the oxidative coupling of 4-methyl-3-hydroxy-2-aminobenzoic acid catalyzed by laccase (from Trametes versicolor) with a yield of 53% (Scheme 6). Actinocin is a kind of actinomycin which can effectively block the transcription of cancer cell DNA, so it has a good anticancer ability [111]. Moreover, 2-amino-3H-phenoxazine-3-ketone has been synthesized by using laccase catalysis under similar conditions [112].
Furthermore, the amination of hydroquinone catalyzed by laccase has excellent advantages in synthesizing novel antibiotics and amino acid derivatives. For instance, eight new penicillin can be synthesized through the coupling of ampicillin or amoxicillin-2,5-dihydroxybenzoic acid derivatives catalyzed by laccase (Scheme 7) [112]. L-phenylalanine and L-tryptophan derivatives are synthesized by laccase catalysis (Scheme 8 and Scheme 9) [113,114,115].
Multifunctional dimers materials are synthesized through the C–N formation of low-molecular-weight organic compounds catalyzed by laccase. The process is mild, green, and efficient and can be widely utilized in the green synthesis of some drugs and amino acids [111,114]. However, its practical applications in the synthesis of some critical antibacterial drugs and acid derivatives are limited because the efficiency of laccase-catalyzed C–N formation is low, and laccase is expensive and not easy to separate, purify, or recycle. The stability of laccase can be improved through immobilization, achieving recyclability and reducing costs, which will help promote the popularization and application of this enzyme in green synthesis [116].

4.2. C–C Bond Formation

Biaryl compounds are essential components of many natural organic compounds, and are becoming increasingly widely used in the chemical industry, medicine, food, and other fields. In recent years, biaryl compounds have generally been synthesized using reductive coupling and oxidative coupling, with the problems of low atomic utilization rates, high cost, and severe environmental pollution. Consequently, the mild catalytic synthesis of biaryl compounds is one of the most studied areas in organic chemistry [117]. In 2005, a small library of biaryl compounds was prepared using the laccase-catalyzed C–C homologous coupling of free radicals (Table 5), with the highest yield of the product being 85% [109].
In order to highlight the potential of laccase-mediated biaryl synthesis, Beifuss and coworkers provided an elegant approach for 3-tert-butyl-1H-pyrazole-5(4H)-one that contains biaryl compounds using the bio-oxidation of catechol, which obtained a very high biaryl compound yield of 98% when isolated (Table 6) [118].
Although laccase-catalyzed C–C bond formation can be used to obtain biaryl compounds mildly and efficiently, using laccases to prepare biaryl compounds is still a challenging synthetic application which appears difficult to generalize as it is entirely rationalized. As exemplified by Constantin and coworkers, laccase-mediated bio-oxidations are prone to forming unexpected and unpredicted products because subtle differences in the structure of the oxidized substrates can largely influence the outcome of a reaction [119]. The application of laccase in biaryl compound synthesis could be improved by enhancing the substrate specificity of laccase catalysis by modifying laccase [120]. For instance, bacterial laccase with single or multiple mutations more efficiently oxidizes benzo[α]pyrene than the wild-type enzyme (the mutant exhibits about seven fold higher activity than wild-type CueO) [121].

4.3. C–O Bond Formation

Moreover, laccase-catalyzed coupling reactions can add hydroxyl groups from water or fatty alcohol nucleophiles to p-dihydroxy aromatic substrates to form new C–O binding products [109]. Due to their mild conditions (air, room temperature, atmospheric pressure, and lack of toxic substances), these laccase-catalyzed reactions provide a cost-effective, environmentally friendly method for the simple and rapid cross-coupling of hydroxylated substances to the C–O binding of alcohols (Scheme 10) [122,123].
Moreover, laccase from different sources has been used for the selective oxidation of diphenyl phytotoxin trans-resveratrol (3,5,4-trihydroxystilbene) [108,124]: specifically, the laccases from Myceliophthora Thermophila and Chrysophyceae catalyze the production of homologous dimers from trans-resveratrol (3,5,4-trihydroxystilbene) (Scheme 11) [125] and homodimers that synthesize trans-resveratrol for the removal of ROS, respectively. In addition, the formation of actinomycin and hexazinone can be involved in antimicrobials, which are synthesized by laccase-catalyzed aminophenol C–C and C–O bond formation (Scheme 12) [111].
The laccase-catalyzed mechanism of forming C–N, C–C, and C–O bonds is similar to anabolism in vivo. First, unstable phenoxy active radical intermediates are formed by the loss of electrons from the hydroxyl phenol functional group and aromatic ring structure while reducing oxygen to water. Subsequently, oligomers are spontaneously formed by free radical intermediates; then, polymer self- or cross-polymerization products are formed with long-term repeated coupling. Laccase-catalyzed C–N, C–C, and C–O bond formation in single low-molecular-weight organic compounds can be used to obtain various functional polymer materials, for example, some critical antibacterial drugs and amino acid derivatives. In summary, the laccase-catalyzed formation of C–N, C–C, and C–O bonds has crucial application potential in preparing multifunctional polymer materials and in the organic synthesis of amino acid derivatives [67,114].

5. Conclusions and Perspectives

With the development of molecular biology and bioengineering, more laccase protein structures and mechanisms of action have been clarified. The laccase-catalyzed formation and breakage of C–C, C–N, and C–O bonds have also been increasingly involved in multifunctional polymer materials, water pollution purification, and organic synthesis. Moreover, promising theoretical breakthroughs in formation and breakage these bonds in organic compounds have been made since the development of the LMS. For example, the conversion rate for LMS-catalyzed secondary amine C–N breakage can reach up to 99%. Thus, as a green catalyst, laccase has great potential in industrial applications.
Although laccase has been widely used, the further promotion and application of this enzyme are limited by the problems of its primary source being direct extraction, its high price, and its poor stability. Therefore, research on laccase can be strengthened in the following aspects: (1) improving the efficiency of autonomous expression and reducing costs, (2) immobilizing laccase to improve stability and realize recovery, and (3) modifying the selectivity to improve the oxidation activity and increase pocket size. These studies will create a solid foundation for the further application of laccase.

Author Contributions

H.L.: conceptualization, writing—original draft, writing—review and editing. Z.Y.: conceptualization, writing—original draft, writing—review and editing. Q.W.: validation, formal analysis. Y.L.: validation, formal analysis. L.J.: validation, formal analysis. M.X.: writing—review and editing, supervision. C.X.: writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Taishan Scholar Project of Shandong Province [ts201712076].

Data Availability Statement

Data sharing is not applicable to this article. No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ozer, A.; Ay Sal, F.; Belduz, A.O.; Kirci, H.; Canakci, S. Use of feruloyl esterase as laccase-mediator system in paper bleaching. Appl. Biotech. Biochem. 2020, 190, 721–731. [Google Scholar] [CrossRef] [PubMed]
  2. Nakamura, K.; Go, N. Function and molecular evolution of multicopper blue proteins. Cell. Mol. Life Sci. 2005, 62, 2050–2066. [Google Scholar] [CrossRef] [PubMed]
  3. Singh, G.; Arya, S.K. Utility of laccase in pulp and paper industry: A progressive step towards the green technology. Int. J. Biol. Macromol. 2019, 134, 1070–1084. [Google Scholar] [CrossRef] [PubMed]
  4. Claus, H. Laccases: Structure, reactions, distribution. Micron 2004, 35, 93–96. [Google Scholar] [CrossRef] [PubMed]
  5. Kagalkar, A.N.; Khandare, R.V.; Govindwar, S.P. Textile dye degradation potential of plant laccase significantly enhances upon augmentation with redox mediators. RSC Adv. 2015, 5, 80505–80517. [Google Scholar] [CrossRef]
  6. Chauhan, P.S.; Goradia, B.; Saxena, A. Bacterial laccase: Recent update on production, properties and industrial applications. 3 Biotech 2017, 7, 323. [Google Scholar] [CrossRef]
  7. Balaes, T. Ligninolytic enzyme system in ecological adaptation of lignicolous macrofungi. Appl. Ecol. Env. Res. 2017, 15, 207–224. [Google Scholar] [CrossRef]
  8. Janusz, G.; Pawlik, A.; Swiderska-Burek, U.; Polak, J.; Sulej, J.; Jarosz-Wilkolazka, A.; Paszczynski, A. Laccase properties, physiological functions, and evolution. Int. J. Mol. Sci. 2020, 21, 966. [Google Scholar] [CrossRef] [Green Version]
  9. Guan, Z.B.; Luo, Q.; Wang, H.R.; Chen, Y.; Liao, X.R. Bacterial laccases: Promising biological green tools for industrial applications. Cell. Mol. Life Sci. 2018, 75, 3569–3592. [Google Scholar] [CrossRef]
  10. Singh, D.; Gupta, N. Microbial laccase: A robust enzyme and its industrial applications. Biologia 2020, 75, 1183–1193. [Google Scholar] [CrossRef]
  11. Piscitelli, A.; Del Vecchio, C.; Faraco, V.; Giardina, P.; Macellaro, G.; Miele, A.; Pezzella, C.; Sannia, G. Fungal laccases: Versatile tools for lignocellulose transformation. Comptes Rendus Biol. 2011, 334, 789–794. [Google Scholar] [CrossRef] [PubMed]
  12. Sousa, A.C.; Martins, L.O.; Robalo, M.P. Laccases: Versatile biocatalysts for the synthesis of heterocyclic cores. Molecules 2021, 26, 3719. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, J.; Feng, J.; Jia, W.; Chang, S.; Li, S.; Li, Y. Lignin engineering through laccase modification: A promising field for energy plant improvement. Biotechnol. Biofuels 2015, 8, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Asano, T.; Taoka, M.; Yamauchi, Y.; Craig Everroad, R.; Seto, Y.; Isobe, T.; Kamo, M.; Chosa, N. Re-examination of a α-chymotrypsin-solubilized laccase in the pupal cuticle of the silkworm, Bombyx mori: Insights into the regulation system for laccase activation during the ecdysis process. Insect. Biochem. Mol. Biol. 2014, 55, 61–69. [Google Scholar] [CrossRef]
  15. Agrawal, K.; Chaturvedi, V.; Verma, P. Fungal laccase discovered but yet undiscovered. Bioresour. Bioprocess. 2018, 5, 4. [Google Scholar] [CrossRef] [Green Version]
  16. Jones, S.M.; Solomon, E.I. Electron transfer and reaction mechanism of laccases. Cell. Mol. Life Sci. 2015, 72, 869–883. [Google Scholar] [CrossRef] [Green Version]
  17. Malmstrom, B.G.; Agro, A.F.; Antonini, E. The mechanism of laccase-catalyzed oxidations: Kinetic evidence for the involvement of several electron-accepting sites in the enzyme. Eur. J. Biochem. 1969, 9, 383–391. [Google Scholar] [CrossRef]
  18. Dong, J.L.; Zhang, Y.Z. Purification and characterization of two laccase isoenzymes from a ligninolytic fungus Trametes gallica. Prep. Biochem. Biotech. 2004, 34, 179–194. [Google Scholar] [CrossRef]
  19. Chakroun, H.; Mechichi, T.; Martinez, M.J.; Dhouib, A.; Sayadi, S. Purification and characterization of a novel laccase from the ascomycete Trichoderma atroviride: Application on bioremediation of phenolic compounds. Process Biochem. 2010, 45, 507–513. [Google Scholar] [CrossRef]
  20. Giovanelli, G.; Ravasini, G. Apple juice stabilization by combined enzyme-membrane filtration process. LWT Food Sci. Technol. 1993, 26, 1–7. [Google Scholar] [CrossRef]
  21. Marques de Souza, C.G.; Peralta, R.M. Purification and characterization of the main laccase produced by the white-rot fungus Pleurotus pulmonarius on wheat bran solid state medium. J. Basic Microbiol. 2003, 43, 278–286. [Google Scholar] [CrossRef] [PubMed]
  22. Martins, L.O.; Soares, C.M.; Pereira, M.M.; Teixeira, M.; Costa, T.; Jones, G.H.; Henriques, A.O. Molecular and biochemical characterization of a highly stable bacterial laccase that occurs as a structural component of the Bacillus subtilis endospore coat. J. Biol. Chem. 2002, 277, 18849–18859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Endo, K.; Hayashi, Y.; Hibi, T.; Hosono, K.; Beppu, T.; Ueda, K. Enzymological characterization of EpoA, a laccase-like phenol oxidase produced by Streptomyces griseus. J. Biochem. 2003, 133, 671–677. [Google Scholar] [CrossRef] [PubMed]
  24. Molina-Guijarro, J.M.; Pérez Torres, J.; Muñoz Dorado, J.; Guillén Carretero, F.; Moya Lobo, R.; Hernández Cutuli, M.; Arias Fernández, M.E. Detoxification of azo dyes by a novel pH-versatile, salt-resistant laccase from Streptomyces ipomoea. Int. Microbiol. 2009, 12, 13–21. [Google Scholar]
  25. Niladevi, K.N.; Jacob, N.; Prema, P. Evidence for a halotolerant-alkaline laccase in Streptomyces psammoticus: Purification and characterization. Process Biochem. 2008, 43, 654–660. [Google Scholar] [CrossRef]
  26. Grass, G.; Rensing, C. CueO is a multi-copper oxidase that confers copper tolerance in Escherichia coli. Biochem. Biophys. Res. Commun. 2001, 286, 902–908. [Google Scholar] [CrossRef]
  27. Ranocha, P.; McDougall, G.; Hawkins, S.; Sterjiades, R.; Borderies, G.; Stewart, D.; Cabanes-Macheteau, M.; Boudet, A.M.; Goffner, D. Biochemical characterization, molecular cloning and expression of laccases–a divergent gene family–in poplar. Eur. J. Biochem. 1999, 259, 485–495. [Google Scholar] [CrossRef]
  28. Sato, Y.; Bao, W.; Sederoff, R.; Whetten, R. Molecular cloning and expression of eight laccase cDNAs in loblolly pine (Pinus taeda)*. J. Plant. Res. 2001, 114, 147–155. [Google Scholar] [CrossRef]
  29. Hüttermann, A.; Mai, C.; Kharazipour, A. Modification of lignin for the production of new compounded materials. Appl. Microbiol. Biotechnol. 2001, 55, 387–394. [Google Scholar] [CrossRef]
  30. Dittmer, N.T.; Suderman, R.J.; Jiang, H.; Zhu, Y.C.; Gorman, M.J.; Kramer, K.J.; Kanost, M.R. Characterization of cDNAs encoding putative laccase-like multicopper oxidases and developmental expression in the tobacco hornworm, Manduca sexta, and the malaria mosquito, Anopheles gambiae. Insect Biochem. Mol. Biol. 2004, 34, 29–41. [Google Scholar] [CrossRef]
  31. Niu, B.L.; Shen, W.F.; Liu, Y.; Weng, H.B.; He, L.H.; Mu, J.J.; Wu, Z.L.; Jiang, P.; Tao, Y.Z.; Meng, Z.Q. Cloning and RNAi-mediated functional characterization of MaLac2 of the pine sawyer, Monochamus alternatus. Insect Mol. Biol. 2008, 17, 303–312. [Google Scholar] [CrossRef]
  32. Lang, M.; Kanost, M.R.; Gorman, M.P.O. Multicopper oxidase-3 is a laccase associated with the peritrophic matrix of Anopheles gambiae. PLoS ONE 2012, 7, e33985. [Google Scholar] [CrossRef] [Green Version]
  33. Fabbrini, M.; Galli, C.; Gentili, P. Radical or electron-transfer mechanism of oxidation with some laccase/mediator systems. J. Mol. Catal. B-enzmy. 2002, 18, 169–171. [Google Scholar] [CrossRef]
  34. Kunamneni, A.; Plou, F.J.; Ballesteros, A.; Alcalde, M. Laccases and their applications: A patent review. Recent Pat. Biotechnol. 2008, 2, 10–24. [Google Scholar] [CrossRef] [Green Version]
  35. Sigoillot, C.; Record, E.; Belle, V.; Robert, J.L.; Levasseur, A.; Punt, P.J.; van den Hondel, C.A.M.J.J.; Fournel, A.; Sigoillot, J.C.; Asther, M. Natural and recombinant fungal laccases for paper pulp bleaching. Appl. Microbiol. Biotechnol. 2004, 64, 346–352. [Google Scholar] [CrossRef] [PubMed]
  36. Mayolo-Deloisa, K.; Gonzalez-Gonzalez, M.; Rito-Palomares, M. Laccases in food industry: Bioprocessing, potential industrial and biotechnological applications. Front. Bioeng. Biotechnol. 2020, 8, 222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Lassouane, F.; Ait-Amar, H.; Amrani, S.; Rodriguez-Couto, S. A promising laccase immobilization approach for Bisphenol A removal from aqueous solutions. Bioresour. Technol. 2019, 271, 360–367. [Google Scholar] [CrossRef] [PubMed]
  38. Parra Guardado, A.L.; Belleville, M.-P.; Rostro Alanis, M.d.J.; Parra Saldivar, R.; Sanchez-Marcano, J. Effect of redox mediators in pharmaceuticals degradation by laccase: A comparative study. Process Biochem. 2019, 78, 123–131. [Google Scholar] [CrossRef]
  39. Morozova, O.V.; Shumakovich, G.P.; Shleev, S.V.; Yaropolov, Y.I. Laccase–mediator systems and their applications: A Review. Appl. Biochem. Microbiol. 2007, 43, 523–535. [Google Scholar] [CrossRef]
  40. Martínez-Montero, L.; Díaz-Rodríguez, A.; Gotor, V.; Gotor-Fernández, V.; Lavandera, I. Broadening the chemical scope of laccases: Selective deprotection of N-benzyl groups. Green Chem. 2015, 17, 2794–2798. [Google Scholar] [CrossRef]
  41. Verkade, J.M.M.; van Hemert, L.J.C.; Quaedflieg, P.J.L.M.; Schoemaker, H.E.; Schürmann, M.; van Delft, F.L.; Rutjes, F.P.J.T. Laccase-mediated deprotection of para-methoxyphenyl (PMP)-protected amines. Adv. Synth. Catal. 2007, 349, 1332–1336. [Google Scholar] [CrossRef]
  42. Freitas, E.D.; Bubna, G.A.; Brugnari, T.; Kato, C.G.; Nolli, M.; Rauen, T.G.; Regina, D.; Peralta, R.A.; Bracht, A.; Souza, C.D. Removal of bisphenol A by laccases from Pleurotus ostreatus and Pleurotus pulmonarius and evaluation of ecotoxicity of degradation products. Chem. Eng. J. 2017, 330, 1361–1369. [Google Scholar] [CrossRef]
  43. Cañas, A.I.; Camarero, S. Laccases and their natural mediators: Biotechnological tools for sustainable eco-friendly processes. Biotechnol. Adv. 2010, 28, 694–705. [Google Scholar] [CrossRef] [PubMed]
  44. Bourbonnais, R.; Paice, M.G. Oxidation of non-phenolic substrates: An expanded role for laccase in lignin biodegradation. FEBS Lett. 1990, 267, 99–102. [Google Scholar] [CrossRef] [Green Version]
  45. Könst, P.; Kara, S.; Kochius, S.; Holtmann, D.; Arends, I.W.; Ludwig, R.; Hollmann, F. Expanding the scope of laccase-mediator systems. ChemCatChem 2013, 5, 3027–3032. [Google Scholar] [CrossRef]
  46. Bourbonnais, R.; Paice, M.; Freiermuth, B.; Bodie, E.; Borneman, S. Reactivities of various mediators and laccases with kraft pulp and lignin model compounds. Appl. Environ. Microbiol. 1997, 63, 4627–4632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Eggert, C.; Temp, U.; Dean, J.F.; Eriksson, K.-E.L. A fungal metabolite mediates degradation of non-phenolic lignin structures and synthetic lignin by laccase. FEBS Lett. 1996, 391, 144–148. [Google Scholar] [CrossRef] [Green Version]
  48. Kawai, S.; Umezawa, T.; Higuchi, T. < original> Oxidation of methoxylated benzyl alcohols by laccase of Coriolus versicolor in the presence of syringaldehyde. Wood Res. Bull. Wood Res. Inst. Kyoto Univ. 1989, 76, 10–16. [Google Scholar]
  49. Park, S.; Jung, D.; Do, H.; Yun, J.; Lee, D.; Hwang, S.; Lee, S.H. Laccase-mediator system using a natural mediator as a whitening agent for the decolorization of melanin. Polymers 2021, 13, 3671. [Google Scholar] [CrossRef]
  50. Valls, C.; Colom, J.F.; Baffert, C.; Gimbert, I.; Roncero, M.B.; Sigoillot, J.-C. Comparing the efficiency of the laccase–NHA and laccase–HBT systems in eucalyptus pulp bleaching. Biochem. Engin. J. 2010, 49, 401–407. [Google Scholar] [CrossRef]
  51. Gu, Y.; Yuan, L.; Jia, L.; Xue, P.; Yao, H. Recent developments of a co-immobilized laccase–mediator system: A review. RSC Adv. 2021, 11, 29498–29506. [Google Scholar] [CrossRef] [PubMed]
  52. Witayakran, S.; Ragauskas, A.J. Synthetic applications of laccase in green chemistry. Adv. Synth. Catal. 2009, 351, 1187–1209. [Google Scholar] [CrossRef]
  53. Johannes, C.; Majcherczyk, A. Natural mediators in the oxidation of polycyclic aromatic hydrocarbons by laccase mediator systems. Appl. Environ. Microbiol. 2000, 66, 524–528. [Google Scholar] [CrossRef] [Green Version]
  54. Matsumura, E.; Yamamoto, E.; Numata, A.; Kawano, T.; Shin, T.; Murao, S. Structures of the laccase-catalyzed oxidation products of hydroxybenzoic acids in the presence of ABTS [2,2′-azinodi-(3-ethylbenzothiazoline-6-sulfonic acid)]. Agric. Biol. Chem. 1986, 50, 1355–1357. [Google Scholar]
  55. Zhao, Q.; Huang, W.; Luo, Z.; Liu, L.; Lu, Y.; Li, Y.; Li, L.; Hu, J.; Ma, H.; Chen, J. High-capacity aqueous zinc batteries using sustainable quinone electrodes. Sci. Adv. 2018, 4, eaao1761. [Google Scholar] [CrossRef] [Green Version]
  56. Kudanga, T.; Nyanhongo, G.S.; Guebitz, G.M.; Burton, S. Potential applications of laccase-mediated coupling and grafting reactions: A review. Enzym. Microb. Technol. 2011, 48, 195–208. [Google Scholar] [CrossRef]
  57. Li, K.; Helm, R.F.; Eriksson, K.E.L. Mechanistic studies of the oxidation of a non-phenolic lignin model compound by the laccase/1-hydroxybenzotriazole redox system. Biotechnol. Appl. Biochem. 1998, 27, 239–243. [Google Scholar] [CrossRef]
  58. d’Acunzo, F.; Galli, C. First evidence of catalytic mediation by phenolic compounds in the laccase-induced oxidation of lignin models. Eur. J. Biochem. 2003, 270, 3634–3640. [Google Scholar] [CrossRef] [Green Version]
  59. Blanquez, A.; Rodriguez, J.; Brissos, V.; Mendes, S.; Martins, L.O.; Ball, A.S.; Arias, M.E.; Hernandez, M. Decolorization and detoxification of textile dyes using a versatile Streptomyces laccase-natural mediator system. Saudi J. Biol. Sci. 2019, 26, 913–920. [Google Scholar] [CrossRef] [PubMed]
  60. Vandertol-Vanier, H.; Vazquez-Duhalt, R.; Tinoco, R.; Pickard, M. Enhanced activity by poly (ethylene glycol) modification of Coriolopsis gallica laccase. J. Ind. Microbiol. Biotechnol. 2002, 29, 214–220. [Google Scholar] [CrossRef]
  61. Han, M.-J.; Park, H.-T.; Song, H.-G. Degradation of phenanthrene by Trametes versicolor and its laccase. J. Microbiol. 2004, 42, 94–98. [Google Scholar] [PubMed]
  62. Jeon, J.R.; Baldrian, P.; Murugesan, K.; Chang, Y.S. Laccase-catalysed oxidations of naturally occurring phenols: From in vivo biosynthetic pathways to green synthetic applications. Microb. Biotechnol. 2012, 5, 318–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Su, J.; Fu, J.; Wang, Q.; Silva, C.; Cavaco-Paulo, A. Laccase-catalysed oxidations of naturally occurring phenols: From in vivo biosynthetic pathways to green synthetic applications. Crit. Rev. Biotechnol. 2018, 38, 294–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Jeon, J.R.; Kim, E.J.; Murugesan, K.; Park, H.K.; Kim, Y.M.; Kwon, J.H.; Kim, W.G.; Lee, J.Y.; Chang, Y.S. Laccase-catalysed polymeric dye synthesis from plant-derived phenols for potential. Microb. Biotechnol. 2010, 3, 324–335. [Google Scholar] [CrossRef]
  65. Sun, K.; Li, S.; Si, Y.; Huang, Q. Advances in laccase-triggered anabolism for biotechnology applications. Crit. Rev. Biotechnol. 2021, 41, 969–993. [Google Scholar] [CrossRef]
  66. Bull, S.D.; Davies, S.G.; Fenton, G.; Mulvaney, A.W.; Prasad, R.S.; Smith, A.D. Chemoselective debenzylation of N-benzyl tertiary amines with ceric ammonium nitrate. J. Chem. Soc. Perkin Trans. 2000, 22, 3765–3774. [Google Scholar] [CrossRef]
  67. Bull, S.D.; Davies, S.G.; Mulvaney, A.W.; Prasad, R.S.; Smith, A.D.; Fenton, G. Chemoselective oxidative debenzylation of tertiary N-benzyl amines. Chem. Commun. 2000, 5, 337–338. [Google Scholar] [CrossRef]
  68. Cernova, M.; Cerna, I.; Pohl, R.; Hocek, M. Regioselective direct CH arylations of protected uracils. synthesis of 5- and 6-aryluracil bases. J. Org. Chem. 2011, 76, 5309–5319. [Google Scholar] [CrossRef]
  69. Chi, L.-P.; Li, X.-M.; Wan, Y.-P.; Li, X.; Wang, B.-G. Ophiobolin sesterterpenoids and farnesylated phthalide derivatives from the deep sea cold-seep-derived fungus aspergillus insuetus SD-512. J. Nat. Prod. 2020, 83, 3652–3660. [Google Scholar] [CrossRef]
  70. Davies, S.G.; Fletcher, A.M.; Hughes, D.G.; Lee, J.A.; Price, P.D.; Roberts, P.M.; Russell, A.J.; Smith, A.D.; Thomson, J.E.; Williams, O.M.H. Asymmetric synthesis of piperidines and octahydroindolizines using a one-pot ring-closure/N-debenzylation procedure. Tetrahedron 2011, 67, 9975–9992. [Google Scholar] [CrossRef]
  71. Sonobe, T.; Oisaki, K.; Kanai, M. Catalytic aerobic production of imines en route to mild, green, and concise derivatizations of amines. Chem. Sci. 2012, 3, 3249–3255. [Google Scholar] [CrossRef]
  72. Davies, S.G.; Fletcher, A.M.; Foster, E.M.; Houlsby, I.T.; Roberts, P.M.; Schofield, T.M.; Thomson, J.E. The asymmetric syntheses of pyrrolizidines, indolizidines and quinolizidines via two sequential tandem ring-closure/N-debenzylation processes. Org. Biomol. Chem. 2014, 12, 9223–9235. [Google Scholar] [CrossRef] [PubMed]
  73. Mateljak, I.; Monza, E.; Lucas, M.F.; Guallar, V.; Aleksejeva, O.; Ludwig, R.; Leech, D.; Shleev, S.; Alcalde, M. Increasing redox potential, redox mediator activity, and stability in a fungal laccase by computer-guided mutagenesis and directed evolution. ACS Catal. 2019, 9, 4561–4572. [Google Scholar] [CrossRef] [Green Version]
  74. Mateljak, I.; Alcalde, M. Engineering a highly thermostable high-redox potential laccase. ACS Sustain. Chem. Eng. 2021, 9, 9632–9637. [Google Scholar] [CrossRef]
  75. Gomez-Fernandez, B.J.; Risso, V.A.; Sanchez-Ruiz, J.M.; Alcalde, M. Consensus design of an evolved high-redox potential laccase. Front. Bioeng. Biotechnol. 2020, 8, 354. [Google Scholar] [CrossRef]
  76. Moreman, J.; Lee, O.; Trznadel, M.; David, A.; Kudoh, T.; Tyler, C.R. Acute toxicity, teratogenic, and estrogenic effects of bisphenol A and its alternative replacements bisphenol S, bisphenol F, and bisphenol AF in zebrafish embryo-larvae. Environ. Sci. Technol. 2017, 51, 12796–12805. [Google Scholar] [CrossRef]
  77. Mu, X.; Huang, Y.; Li, X.; Lei, Y.; Teng, M.; Li, X.; Wang, C.; Li, Y. Developmental effects and estrogenicity of bisphenol A alternatives in a zebrafish embryo model. Environ. Sci. Technol. 2018, 52, 3222–3231. [Google Scholar] [CrossRef]
  78. Lv, Y.-Z.; Yao, L.; Wang, L.; Liu, W.-R.; Zhao, J.-L.; He, L.-Y.; Ying, G.-G. Bioaccumulation, metabolism, and risk assessment of phenolic endocrine disrupting chemicals in specific tissues of wild fish. Chemosphere 2019, 226, 607–615. [Google Scholar] [CrossRef]
  79. Bhatnagar, A.; Anastopoulos, I. Adsorptive removal of bisphenol A (BPA) from aqueous solution: A review. Chemosphere 2017, 168, 885–902. [Google Scholar] [CrossRef] [PubMed]
  80. Lazim, Z.M.; Hadibarata, T.; Puteh, M.H.; Yusop, Z. Adsorption characteristics of bisphenol A onto low-cost modified phyto-waste material in aqueous solution. Water Air Soil Pollut. 2015, 226, 34. [Google Scholar] [CrossRef]
  81. Huang, Q.; Chai, K.; Zhou, L.; Ji, H. A phenyl-rich beta-cyclodextrin porous crosslinked polymer for efficient removal of aromatic pollutants: Insight into adsorption performance and mechanism. Chem. Eng. J. 2020, 387, 124020. [Google Scholar] [CrossRef]
  82. Hongyan, L.; Zexiong, Z.; Shiwei, X.; He, X.; Yinian, Z.; Haiyun, L.; Zhongsheng, Y. Study on transformation and degradation of bisphenol A by Trametes versicolor laccase and simulation of molecular docking. Chemosphere 2019, 224, 743–750. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, M.; Waigi, M.G.; Li, S.; Sun, K.; Si, Y. Fungal laccase-mediated humification of estrogens in aquatic ecosystems. Water Res. 2019, 166, 115040. [Google Scholar] [CrossRef] [PubMed]
  84. Uchida, H.; Fukuda, T.; Miyamoto, H.; Kawabata, T.; Suzuki, M.; Uwajima, T. Polymerization of bisphenol A by purified laccase from Trametes villosa. Biochem. Biophys. Res. Commun. 2001, 287, 355–358. [Google Scholar] [CrossRef] [PubMed]
  85. Nelson, A.A.; Anne, F.T. A novel acid-stable intracellular laccase from Aureobasidium pullulans: Purification, characterization and application in the removal of bisphenol A from solutions. Biocatal. Agric. Biotechnol. 2021, 33, 101966. [Google Scholar] [CrossRef]
  86. Daâssi, D.; Prieto, A.; Zouari-Mechichi, H.; Martínez, M.J.; Nasri, M.; Mechichi, T. Degradation of bisphenol A by different fungal laccases and identification of its degradation products. Int. Biodeterior. Biodegrad. 2016, 110, 181–188. [Google Scholar] [CrossRef] [Green Version]
  87. Zeng, S.; Zhao, J.; Xia, L. Simultaneous production of laccase and degradation of bisphenol A with Trametes versicolor cultivated on agricultural wastes. Bioprocess. Biosyst. Eng. 2017, 40, 1237–1245. [Google Scholar] [CrossRef] [Green Version]
  88. Brugnari, T.; Pereira, M.G.; Bubna, G.A.; de Freitas, E.N.; Contato, A.G.; Corrêa, R.C.G.; Castoldi, R.; de Souza, C.G.M.; de Moraes, M.d.L.T.; Bracht, A. A highly reusable MANAE-agarose-immobilized Pleurotus ostreatus laccase for degradation of bisphenol A. Sci. Total Environ. 2018, 634, 1346–1351. [Google Scholar] [CrossRef]
  89. Fujisawa, M.; Hirai, H.; Nishida, T. Degradation of polyethylene and nylon-66 by the laccase-mediator system. J. Polym. Environ. 2001, 9, 103–108. [Google Scholar] [CrossRef]
  90. Yao, C.; Xia, W.; Dou, M.; Du, Y.; Wu, J. Oxidative degradation of UV-irradiated polyethylene by laccase-mediator system. J. Hazard. Mater. 2022, 440, 129709. [Google Scholar] [CrossRef] [PubMed]
  91. Suzuki, K.; Hirai, H.; Murata, H.; Nishida, T. Removal of estrogenic activities of 17beta-estradiol and ethinylestradiol by ligninolytic enzymes from white rot fungi. Water Res. 2003, 37, 1972–1975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Zheng, F.; An, Q.; Meng, G.; Wu, X.J.; Dai, Y.C.; Si, J.; Cui, B.K. A novel laccase from white rot fungus Trametes orientalis: Purification, characterization, and application. Int. J. Biol. Macromol. 2017, 102, 758–770. [Google Scholar] [CrossRef] [PubMed]
  93. Nadaroglu, H.; Mosber, G.; Gungor, A.A.; Adıguzel, G.; Adiguzel, A. Biodegradation of some azo dyes from wastewater with laccase from Weissella viridescens LB37 immobilized on magnetic chitosan nanoparticles. J. Water Process Eng. 2019, 31, 100866. [Google Scholar] [CrossRef]
  94. Taghizadeh, T.; Talebian-Kiakalaieh, A.; Jahandar, H.; Amin, M.; Tarighi, S.; Faramarzi, M.A. Biodegradation of bisphenol A by the immobilized laccase on some synthesized and modified forms of zeolite Y. J. Hazard. Mater. 2020, 386, 121950. [Google Scholar] [CrossRef]
  95. Latif, A.; Maqbool, A.; Sun, K.; Si, Y. Immobilization of Trametes Versicolor laccase on Cu-alginate beads for biocatalytic degradation of bisphenol A in water: Optimized immobilization, degradation and toxicity assessment. J. Environ. Chem. Eng. 2022, 10, 107089. [Google Scholar] [CrossRef]
  96. Bilal, M.; Jing, Z.; Zhao, Y.; Iqbal, H.M.N. Immobilization of fungal laccase on glutaraldehyde cross-linked chitosan beads and its bio-catalytic potential to degrade bisphenol A. Biocatal. Agric. Biotechnol. 2019, 19, 101174. [Google Scholar] [CrossRef]
  97. Fu, M.; Xing, J.; Ge, Z. Preparation of laccase-loaded magnetic nanoflowers and their recycling for efficient degradation of bisphenol A. Sci. Total Environ. 2019, 651, 2857–2865. [Google Scholar] [CrossRef]
  98. Sadeghzadeh, S.; Ghobadi Nejad, Z.; Ghasemi, S.; Khafaji, M.; Borghei, S.M. Removal of bisphenol A in aqueous solution using magnetic cross-linked laccase aggregates from Trametes hirsuta. Bioresour. Technol. 2020, 306, 123169. [Google Scholar] [CrossRef]
  99. Dar, O.I.; Aslam, R.; Pan, D.; Sharma, S.; Andotra, M.; Kaur, A.; Jia, A.-Q.; Faggio, C. Source, bioaccumulation, degradability and toxicity of triclosan in aquatic environments: A review. Environ. Technol. Innov. 2022, 25, 102122. [Google Scholar] [CrossRef]
  100. Reiss, R.; Lewis, G.; Griffin, J. An ecological risk assessment for triclosan in the terrestrial environment. Environ. Toxicol. Chem. 2009, 28, 1546–1556. [Google Scholar] [CrossRef]
  101. Guo, J.; Iwata, H. Risk assessment of triclosan in the global environment using a probabilistic approach. Ecotoxicol. Environ. Saf. 2017, 143, 111–119. [Google Scholar] [CrossRef]
  102. Milanovic, M.; Duric, L.; Milosevic, N.; Milic, N. Comprehensive insight into triclosan—From widespread occurrence to health outcomes. Environ. Sci. Pollut. Res. Int. 2021, 30, 25119–25140. [Google Scholar] [CrossRef]
  103. Anger, C.T.; Sueper, C.; Blumentritt, D.J.; McNeill, K.; Engstrom, D.R.; Arnold, W.A. Quantification of triclosan, chlorinated triclosan derivatives, and their dioxin photoproducts in lacustrine sediment cores. Environ. Sci. Technol. 2013, 47, 1833–1843. [Google Scholar] [CrossRef] [Green Version]
  104. Singh, J.; Kumar, P.; Saharan, V.; Kapoor, R.K. Simultaneous laccase production and transformation of bisphenol-A and triclosan using Trametes versicolor. 3 Biotech 2019, 9, 129. [Google Scholar] [CrossRef] [PubMed]
  105. Dou, R.N.; Wang, J.H.; Chen, Y.C.; Hu, Y.Y. The transformation of triclosan by laccase: Effect of humic acid on the reaction kinetics, products and pathway. Environ. Pollut. 2018, 234, 88–95. [Google Scholar] [CrossRef] [PubMed]
  106. Asano, T.; Seto, Y.; Hashimoto, K.; Kurushima, H. Mini-review an insect-specific system for terrestrialization: Laccase mediated cuticle formation. Insect Biochem. Mol. Biol. 2019, 108, 61–70. [Google Scholar] [CrossRef]
  107. Kour, D.; Rana, K.L.; Thakur, S.; Sharma, S.; Yadav, N.; Rastegari, A.A.; Yadav, A.N.; Saxena, A.K. Disruption of protease genes in microbes for production of heterologous proteins. In New and Future Developments in Microbial Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2019; pp. 35–75. [Google Scholar] [CrossRef]
  108. Theerachat, M.; Guieysse, D.; Morel, S.; Remaud-Simeon, M.; Chulalaksananukul, W. Laccases from marine organisms and their applications in the biodegradation of toxic and environmental pollutants: A Review. Appl. Biochem. Biotechnol. 2019, 187, 583–611. [Google Scholar] [CrossRef]
  109. Ciecholewski, S.; Hammer, E.; Manda, K.; Bose, G.; Nguyen, V.T.; Langer, P.; Schauer, F. Laccase-catalyzed carbon–carbon bond formation: Oxidative dimerization of salicylic esters by air in aqueous solution. Tetrahedron 2005, 61, 4615–4619. [Google Scholar] [CrossRef]
  110. Abdel-Mohsen, H.; Conrad, J.; Harms, K.; Nohr, D.; Beifuss, U. Laccase-catalyzed green synthesis and cytotoxic activity of novel pyrimidobenzothiazoles and catechol thioethers. RSC Adv. 2017, 7, 17427–17441. [Google Scholar] [CrossRef] [Green Version]
  111. Osiadacz, J.; Al-Adhami, A.J.; Bajraszewska, D.; Fischer, P.; Peczyñska-Czoch, W. On the use of Trametes versicolor laccase for the conversion of 4-methyl-3-hydroxyanthranilic acid to actinocin chromophore. J. Biotechnol. 1999, 72, 141–149. [Google Scholar] [CrossRef]
  112. Pilati, S.; Malacarne, G.; Navarro-Payá, D.; Tomè, G.; Riscica, L.; Cavecchia, V.; Matus, J.T.; Moser, C.; Blanzieri, E. Vitis oneGenE: A causality-based approach to generate gene networks in Vitis vinifera sheds light on the laccase and dirigent gene families. Biomolecules 2021, 11, 1744. [Google Scholar] [CrossRef]
  113. Mogharabi, M.; Faramarzi, M.A. Laccase and laccase-mediated systems in the synthesis of organic compounds. Adv. Synth. Catal. 2014, 356, 897–927. [Google Scholar] [CrossRef]
  114. Bassanini, I.; Ferrandi, E.E.; Riva, S.; Monti, D. Biocatalysis with laccases: An Updated Overview. Catalysts 2020, 11, 26. [Google Scholar] [CrossRef]
  115. Hahn, V.; Mikolasch, A.; Manda, K.; Gördes, D.; Thurow, K.; Schauer, F. Derivatization of amino acids by fungal laccases: Comparison of enzymatic and chemical methods. J. Mol. Catal. B Enzym. 2009, 60, 76–81. [Google Scholar] [CrossRef]
  116. Rouhani, S.; Azizi, S.; Kibechu, R.W.; Mamba, B.B.; Msagati, T. Laccase immobilized Fe3O4-graphene oxide nanobiocatalyst improves stability and immobilization efficiency in the green preparation of sulfa drugs. Catalysts 2020, 10, 459. [Google Scholar] [CrossRef] [Green Version]
  117. Wencel-Delord, J.; Panossian, A.; Leroux, F.R.; Colobert, F. Recent advances and new concepts for the synthesis of axially stereoenriched biaryls. Chem. Soc. Rev. 2015, 44, 3418–3430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Emirdağ-Öztürk, S.; Hajdok, S.; Conrad, J.; Beifuss, U. Laccase-catalyzed reaction of 3-tert-butyl-1H-pyrazol-5 (4H)-one with substituted catechols using air as an oxidant. Tetrahedron 2013, 69, 3664–3668. [Google Scholar] [CrossRef]
  119. Wei, T.B.; Chen, Y.; Hu, Q.; Yang, J.K.; Huo, Z.Q.; Duan, R.F.; Wang, J.X.; Zeng, Y.P.; Li, J.M.; Liao, Y.X. Hydride vapor phase epitaxy of strain-reduced GaN film on nano-island template produced using self-assembled CsCl nanospheres. Mater. Lett. 2012, 68, 327. [Google Scholar] [CrossRef]
  120. Xie, H.; Guo, R.; Zhong, H.; Feng, Q.; Lan, Z.; Qin, B.; Ward, K.J.; Jackson, M.A.; Xia, Y.; Chen, X. Shotgun metagenomics of 250 adult twins reveals genetic and environmental impacts on the gut microbiome. Cell Syst. 2016, 3, 572–584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Yue, Q.; Yang, Y.; Zhao, J.; Zhang, L.; Xu, L.; Chu, X.; Liu, X.; Tian, J.; Wu, N. Identification of bacterial laccase CueO mutation from the metagenome of chemical plant sludge. Bioresour. Bioprocess. 2017, 4, 48. [Google Scholar] [CrossRef] [Green Version]
  122. Manda, K.; Gördes, D.; Mikolasch, A.; Hammer, E.; Schmidt, E.; Thurow, K.; Schauer, F. Carbon-oxygen bond formation by fungal laccases: Cross-coupling of 2, 5-dihydroxy-N-(2-hydroxyethyl)-benzamide with the solvents water, methanol, and other alcohols. Appl. Biochem. Microbiol. 2007, 76, 407–416. [Google Scholar] [CrossRef] [PubMed]
  123. Mikolasch, A.; Matthies, A.; Lalk, M.; Schauer, F. Laccase-induced C–N coupling of substituted p-hydroquinones with p-aminobenzoic acid in comparison with known chemical routes. Appl. Microbiol. Biotechnol. 2008, 80, 389–397. [Google Scholar] [CrossRef]
  124. Chong, J.; Poutaraud, A.; Hugueney, P. Metabolism and roles of stilbenes in plants. Plant Sci. 2009, 177, 143–155. [Google Scholar] [CrossRef]
  125. Jeon, J.-R.; Chang, Y.-S. Laccase-mediated oxidation of small organics: Bifunctional roles for versatile applications. Trends Biotechnol. 2013, 31, 335–341. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Crystal structure of the laccase (LccI) (PDB code: 1GYC) in Trametes versicolor. (A) The overall structure of LccI; (B) structure of the schematic representation of the four copper sites in LccI.
Figure 1. Crystal structure of the laccase (LccI) (PDB code: 1GYC) in Trametes versicolor. (A) The overall structure of LccI; (B) structure of the schematic representation of the four copper sites in LccI.
Catalysts 13 00750 g001
Figure 2. Oxidation of substrates by LMS.
Figure 2. Oxidation of substrates by LMS.
Catalysts 13 00750 g002
Figure 3. The chemical structures of several artificial and natural redox mediators in laccase-catalyzed oxidation reaction systems. (A) natural redox mediators; (B) artificial redox mediators.
Figure 3. The chemical structures of several artificial and natural redox mediators in laccase-catalyzed oxidation reaction systems. (A) natural redox mediators; (B) artificial redox mediators.
Catalysts 13 00750 g003
Figure 4. Oxidation of ABTS by Laccase.
Figure 4. Oxidation of ABTS by Laccase.
Catalysts 13 00750 g004
Scheme 1. N-PMP removal reaction by Laccase.
Scheme 1. N-PMP removal reaction by Laccase.
Catalysts 13 00750 sch001
Scheme 2. Debenzylation of N-protected secondary amine by laccase/TEMPO system.
Scheme 2. Debenzylation of N-protected secondary amine by laccase/TEMPO system.
Catalysts 13 00750 sch002
Scheme 3. Debenzylation of N-protected tertiary amine by laccase/TEMPO system.
Scheme 3. Debenzylation of N-protected tertiary amine by laccase/TEMPO system.
Catalysts 13 00750 sch003
Scheme 4. Free radical coupling and oxidative decomposition of BPA catalyzed by fungal laccase.
Scheme 4. Free radical coupling and oxidative decomposition of BPA catalyzed by fungal laccase.
Catalysts 13 00750 sch004
Scheme 5. Proposed triclosan transformation pathways in the absence and presence of HBT by laccase-mediated reaction systems.
Scheme 5. Proposed triclosan transformation pathways in the absence and presence of HBT by laccase-mediated reaction systems.
Catalysts 13 00750 sch005
Figure 5. Possible mechanism of C–N bond breakage catalyzed by laccase.
Figure 5. Possible mechanism of C–N bond breakage catalyzed by laccase.
Catalysts 13 00750 g005
Scheme 6. Preparation of actinocin by laccase catalyzed by 4-methyl-3-hydroxy-2-aminobenzoic acid.
Scheme 6. Preparation of actinocin by laccase catalyzed by 4-methyl-3-hydroxy-2-aminobenzoic acid.
Catalysts 13 00750 sch006
Scheme 7. Laccase-catalyzed synthesis of penicillin.
Scheme 7. Laccase-catalyzed synthesis of penicillin.
Catalysts 13 00750 sch007
Scheme 8. Synthesis of L-phenylalanine derivatives catalyzed by laccase.
Scheme 8. Synthesis of L-phenylalanine derivatives catalyzed by laccase.
Catalysts 13 00750 sch008
Scheme 9. Synthesis of L-tryptophan derivatives catalyzed by laccase.
Scheme 9. Synthesis of L-tryptophan derivatives catalyzed by laccase.
Catalysts 13 00750 sch009
Scheme 10. Carbon–oxygen bond formation by fungal laccases: cross-coupling of 2, 5-dihydroxy-N-(2-hydroxyethyl)-benzamide with the solvents water, methanol, and other alcohols.
Scheme 10. Carbon–oxygen bond formation by fungal laccases: cross-coupling of 2, 5-dihydroxy-N-(2-hydroxyethyl)-benzamide with the solvents water, methanol, and other alcohols.
Catalysts 13 00750 sch010
Scheme 11. Laccase catalyzes trans-resveratrol to form homologous dimers.
Scheme 11. Laccase catalyzes trans-resveratrol to form homologous dimers.
Catalysts 13 00750 sch011
Scheme 12. Laccase-catalyzed synthesis of actinocin via oxidative homocoupling.
Scheme 12. Laccase-catalyzed synthesis of actinocin via oxidative homocoupling.
Catalysts 13 00750 sch012
Table 1. Enzymatic properties of laccase from different sources.
Table 1. Enzymatic properties of laccase from different sources.
SourceMolecular MassOptimal pHOptimal Temperature (°C)Ref.
Fungal46–80 kD2.2–630–55[18,19,20,21]
Bacteria43–114 kD4.0–8.040–75[22,23,24,25,26]
Plant59.2–140 kD6.720[27,28,29]
Animal73–110 kD6.5–8.0-[30,31,32]
Table 2. Substrates catalyzed by Laccase for N-PMP removal.
Table 2. Substrates catalyzed by Laccase for N-PMP removal.
EntryPMP AmineLaccaseProduct and Yield
1
2
Catalysts 13 00750 i001
Substrate 1
T
AB
Catalysts 13 00750 i002
Product 1
0%
0%
3
4
Catalysts 13 00750 i003
Substrate 2
T
AB
Catalysts 13 00750 i004
Product 2
31%
56%
5
6
Catalysts 13 00750 i005
Substrate 3
T
AB
Catalysts 13 00750 i006
Product 3
67%
71%
7
8
Catalysts 13 00750 i007
Substrate 4
T
AB
Catalysts 13 00750 i008
Product 4
64%
74%
9
10
Catalysts 13 00750 i009
Substrate 5
T
AB
Catalysts 13 00750 i010
Product 5
89%
52%
11
12
Catalysts 13 00750 i011
Substrate 6
T
AB
Catalysts 13 00750 i012
Product 6
34%
41%
13
14
Catalysts 13 00750 i013
Substrate 7
T
AB
Catalysts 13 00750 i014
Product 7
30%
47%
AB: laccase from Agaricus bisporus, T: laccase from Trametes versicolor.
Table 3. Laccase-catalyzed secondary amine benzyl removal of substrate.
Table 3. Laccase-catalyzed secondary amine benzyl removal of substrate.
EntryAmineProduct and Yield
1Catalysts 13 00750 i015
1a
Catalysts 13 00750 i016
1b
7%
2Catalysts 13 00750 i017
2a
Catalysts 13 00750 i018
2b
97%
3Catalysts 13 00750 i019
3a
Catalysts 13 00750 i020
3b
99%
4Catalysts 13 00750 i021
4a
Catalysts 13 00750 i022
4b
99%
5Catalysts 13 00750 i023
5a
Catalysts 13 00750 i024
5b
97%
6Catalysts 13 00750 i025
6a
Catalysts 13 00750 i026
6b
99%
7Catalysts 13 00750 i027
7a
Catalysts 13 00750 i028
7b
99%
8Catalysts 13 00750 i029
8a
Catalysts 13 00750 i030
8b
<1%
Table 4. Laccase-catalyzed tertiary amine benzyl removal of substrate.
Table 4. Laccase-catalyzed tertiary amine benzyl removal of substrate.
EntryAmineProduct and Yield
1Catalysts 13 00750 i031
9a
Catalysts 13 00750 i032
9b
4%
2Catalysts 13 00750 i033
10a
Catalysts 13 00750 i034
10b
<1%
3Catalysts 13 00750 i035
11a
Catalysts 13 00750 i036
11b
<1%
Table 5. Overview of the oxidative homocoupling of salicylic esters conducted by Ciecholewski.
Table 5. Overview of the oxidative homocoupling of salicylic esters conducted by Ciecholewski.
Catalysts 13 00750 i037
R1R2R3R4Yield%Number of Products
HHHH00
MeHHH00
MeMeMeMe841
MeMe(CH2)2ClMe851
MeMe(CH2)2ClMe855
EtMe(CH2)2ClMe745
EtMe(CH2)2ClH0>7
MeHMeH783
MeMeHMe0>7
Table 6. The 3-tert-butyl-1H-pyrazol-5(4H)-one containing biaryl compounds obtained by laccase-catalyzed oxidation of catechol.
Table 6. The 3-tert-butyl-1H-pyrazol-5(4H)-one containing biaryl compounds obtained by laccase-catalyzed oxidation of catechol.
Catalysts 13 00750 i038
RRegioisomerRatioYield%
Ha-95
Mea,b91:998
OMea-80
Fa,b60:4098
Brb-96
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

Lin, H.; Yu, Z.; Wang, Q.; Liu, Y.; Jiang, L.; Xu, C.; Xian, M. Application of Laccase Catalysis in Bond Formation and Breakage: A Review. Catalysts 2023, 13, 750. https://doi.org/10.3390/catal13040750

AMA Style

Lin H, Yu Z, Wang Q, Liu Y, Jiang L, Xu C, Xian M. Application of Laccase Catalysis in Bond Formation and Breakage: A Review. Catalysts. 2023; 13(4):750. https://doi.org/10.3390/catal13040750

Chicago/Turabian Style

Lin, Huan, Zongjiang Yu, Qian Wang, Yaojie Liu, Long Jiang, Chao Xu, and Mo Xian. 2023. "Application of Laccase Catalysis in Bond Formation and Breakage: A Review" Catalysts 13, no. 4: 750. https://doi.org/10.3390/catal13040750

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