Next Article in Journal
The Human-Restricted Isoform of the α7 nAChR, CHRFAM7A: A Double-Edged Sword in Neurological and Inflammatory Disorders
Previous Article in Journal
LINE-1 Cargo and Reverse Transcriptase Activity Profiles in Extracellular Vesicles from Lung Cancer Cells and Human Plasma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Laccases and Tyrosinases in Organic Synthesis

by
Ludmila Martínková
1,*,
Barbora Křístková
1,2 and
Vladimír Křen
1
1
Institute of Microbiology of the Czech Academy of Sciences, Vídeňská 1083, CZ-142 20 Prague, Czech Republic
2
Faculty of Food and Biochemical Technology, University of Chemistry and Technology, Technická 5, CZ-166 28 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(7), 3462; https://doi.org/10.3390/ijms23073462
Submission received: 28 February 2022 / Revised: 17 March 2022 / Accepted: 18 March 2022 / Published: 22 March 2022
(This article belongs to the Section Bioactives and Nutraceuticals)

Abstract

:
Laccases (Lac) and tyrosinases (TYR) are mild oxidants with a great potential in research and industry. In this work, we review recent advances in their use in organic synthesis. We summarize recent examples of Lac-catalyzed oxidation, homocoupling and heterocoupling, and TYR-catalyzed ortho-hydroxylation of phenols. We highlight the combination of Lac and TYR with other enzymes or chemical catalysts. We also point out the biological and pharmaceutical potential of the products, such as dimers of piceid, lignols, isorhamnetin, rutin, caffeic acid, 4-hydroxychalcones, thiols, hybrid antibiotics, benzimidazoles, benzothiazoles, pyrimidine derivatives, hydroxytyrosols, alkylcatechols, halocatechols, or dihydrocaffeoyl esters, etc. These products include radical scavengers; antibacterial, antiviral, and antitumor compounds; and building blocks for bioactive compounds and drugs. We summarize the available enzyme sources and discuss the scalability of their use in organic synthesis. In conclusion, we assume that the intensive use of laccases and tyrosinases in organic synthesis will yield new bioactive compounds and, in the long-term, reduce the environmental impact of industrial organic chemistry.

Graphical Abstract

1. Introduction

Laccase (Lac; EC 1.10.3.2)—the “blue enzyme for green chemistry” [1] was first described in 1883 as a “diastatic matter” (Figure 1a) from the sap of the tree Toxicodendron vernicifluum (former name Rhus verniciflua; “Chinese lacquer tree”; Figure 1b). The main constituents of the sap are water, urushiol (a mixture of substituted benzene-1,2-diols—catechols), and polysaccharides. Laccase constitutes approximately 1% of the sap and catalyzes the polymerization of urushiol [2]. This sap has been used as varnish (“urushi”) in the traditional manufacture of highly regarded “urushi” goods [2,3,4]. Annual production of the sap was nearly four thousand tons in 2014 [2], but (semi)synthetic alternatives are being sought.
Later, the sources of Lacs were greatly expanded. Lac activities were found in other plants (poplar, sycamore maple, tulip tree, tobacco, maize, rice, oilseed rape, etc.) and in the two largest phyla of fungi (Basidiomycota, Ascomycota). A number of Lacs were also reported in the bacterial genera Azospirillum, Bacillus, Streptomyces, Escherichia, Pseudomonas, Thermus, Sinorhizobium, Oscillatoria, Haloferax, etc., and in various insects [5]. Lacs from different sources differ in their redox potential, which is higher in fungal Lacs (470–810 mV) than in bacterial Lacs or the already mentioned enzyme from the “lacquer tree” (approximately 400 mV) [6].
Lac is a copper protein with four copper atoms in its active site [1]. The copper atoms are classified as T1, T2, T3α, and T3β [7], and have different coordination environments [1,7]. The absorption of the T1 site at 605 nm (in the Lac from Trametes versicolor) is the cause of the blue color of Lac [8]. However, some laccases with an altered catalytic site structure may appear “yellow“ or “white“ [8,9,10]. We refer to previous studies [1,7,8,11,12] for an explanation of the catalytic mechanism and the structure-activity relationships in Lacs.
Figure 1. (a) The first report of a “diastatic matter”, i.e., laccase, in the “urushi” sap [13], with permission; here: page 483; (b) the sap producing tree Toxicodendron vernicifluum (iStock.com/Falombini); with permission.
Figure 1. (a) The first report of a “diastatic matter”, i.e., laccase, in the “urushi” sap [13], with permission; here: page 483; (b) the sap producing tree Toxicodendron vernicifluum (iStock.com/Falombini); with permission.
Ijms 23 03462 g001
Lac is widely recognized as a mild and environmentally friendly oxidation catalyst with atmospheric oxygen as a reactant and water as a byproduct. The variety of products formed by the use of Lacs is broad. The radicals formed during the reaction yield various products by coupling. In addition, mediators broaden the range of substrates. They allow alleviation of the limitations caused by the redox potential of Lacs and extend their use to non-phenol compounds. 2,2′-Azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) or 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) are widely used in Lac-mediator systems. For example, ABTS has improved the removal of both phenol and non-phenol synthetic dyes (bromophenol blue, xylene cyanol, Coommassie® Brilliant Blue R-250) by an immobilized Lac [14]. ABTS or a nature-derived mediator, ferulic acid, have been useful for the remediation of soil contaminated with polyaromatic hydrocarbons [15]. ABTS was used together with carbon nanotubes to mediate the oxidation of anthracene to the corresponding quinone. This allowed the construction of a new anthracene sensor with high sensitivity [16].
Thus, Lacs can be used in a variety of industries (paper, textile, food), materials engineering, bioremediation, biofuel cells, or biocatalysis. However, most of the proposed applications are still waiting to be used on a larger scale. In the last decade, the properties and uses of Lacs have been summarized in comprehensive reviews [17,18,19] and reviews focusing on subtopics such as applications of fungal Lacs [20], evolution and in vivo functions [5], biodegradation [21,22,23,24], biosensors [25], polymer synthesis [26] and grafting [27], biorefineries [28], food industry [29], wood composite production [30], immobilization [26,31,32,33,34,35,36], and biocatalysis in ionic liquids [37] or Lac mimics [38]. For background knowledge on Lac (structure, Lac-mediator reactions, applications) we refer to the review by Riva [1], and for an overview of laccases, their origin, and properties we refer to the review by Baldrian [39].
The common feature of tyrosinase (TYR; EC 1.14.18.1) and Lac is the involvement of copper in their catalytic cycles. However, TYR contains two copper atoms coordinated by His residues. The oxidation state of copper changes between Cu+ and Cu2+ within the catalytic cycle as in Lac. The reaction mechanism of TYR has been discussed previously [40,41].
The range of TYR applications is smaller compared to Lacs, as their substrate scope is not as broad. However, TYRs offer additional possibilities: In organic chemistry, they have been used to catalyze ortho-hydroxylations. They can also be useful in the modification of biopolymers [42] and biodegradation [21]. Most recent reviews on TYRs have focused on TYR inhibitors [43,44,45,46,47,48,49], which have been intensively studied because the enzyme plays a key role in melanogenesis. TYRs, particularly the readily available TYR from Agaricus bisporus, have often been used as models for testing new inhibitors with potential application in medicine and cosmetics [46,48]. One review addressed the kinetics of TYR [50].
Recent studies have suggested new uses for Lacs and TYRs in the production of bioactive compounds and building blocks. Thus, they open pathways to sustainable organic syntheses. We believe that this important area of biocatalysis should be summarized and its perspectives discussed. First, this work summarizes the synthetic applications of Lacs based on the literature since approximately 2016, which was not included in previous reviews on a similar topic [51,52]. Second, the applications of tyrosinases are summarized analogously, based on the literature of the past decade. This is because a proper review of tyrosinase biocatalysts has been lacking during this period.

2. Laccase-Catalyzed Reactions

The ability of Lac to form radicals that result in various types of products such as oligo- and polymers has long been known. However, the scope of Lac products has greatly expanded since the last reviews [51,52]. In the following, we categorize the applications as homocoupling, heterocoupling, and other oxidation reactions.

2.1. Homocoupling

The improved biological activities or solubilities of oligomers of natural compounds [52] has prompted further research in this area. This improvement in the product properties has been explained by the multiplication of functional (hydroxyl) groups [53]. Recently, these types of products were prepared from several natural compounds such as piceid [54], lignols [55], rutin [56], caffeic acid [57,58], 4-hydroxychalcons [59], or isorhamnetin [59].
Piceid is a β-glucopyranoside of resveratrol, a phytoalexin and well-known dietary supplement, although with uncertain bioactivity. Oxidation of piceid (≈18 mM) by Lac yielded a mixture of two dimeric glucosides, which were then partially or completely deglucosylated (Figure 2). The isolated yield of the resveratrol dimer (related to piceid) was ≈45%. In addition, its enantiomers, or the diastereomers of the glucosylated products, were separated by HPLC. Radical (2,2-diphenyl-1-picrylhydrazyl, DPPH) scavenging was weaker in the dimers than in piceid or resveratrol; nevertheless, the biological activities of the new dimers are of further interest [54].
Analogously, dimers were obtained from 50 mM glucosides of coumaryl and coniferyl alcohol (Figure 3). The bulkier diglycosylated (glucorhamnosyl) coniferyl alcohol was also a substrate [55]. The products of glycoside dimerization were prepared in acceptable (34–46%) yields in the tens to hundreds of mg scale [54,55]. The advantages of using glucosides as substrates are higher solubility and limitation of product diversity.
The low solubility of the substrate can also be addressed by medium engineering. Thus, the Lac-catalyzed reaction of the glycoside rutin was performed using an optimized (biphasic) system that consisted of ionic liquid (cholinium dihydrogen phosphate) and polyethylene glycol 600 [56]. The rutin oligomer, which is more water-soluble than rutin, was obtained with 95% conversion from a 3 g/L (≈5 mM) substrate. The conversion was still approximately 90% when using the catalyst two or three times. The reaction was also feasible at a substrate concentration of 10 g/L (≈16 mM), and the conversion was >93%.
Dimeric products were also obtained from flavonoid isorhamnetin [53] or caffeic acid [57,58]. The product of isorhamnetin (12.5 mM) was probably a C-C-linked dimer (Figure 4). It was obtained in a moderate yield (<30%) together with another (unidentified) product that was thought to contain a C-O-C bond. The DPPH scavenging activity of the first product and its ability to inhibit the growth of some bacteria (Listeria, Staphylococcus) was about twice that of the monomer.
Dimerization of caffeic acid (CFA; 10 mM) to phellinsin A (Figure 5) increased radical scavenging activity by 50% and 80%, respectively, as determined by DPPH and Trolox assays [57], respectively. The catalyst was a heterologously-produced bacterial Lac resistant to detergents [57]. Optionally, potato peel (blended and lyophilized) was the source of chlorogenic acid (CLA; ≈0.5 mg/g potato peel powder). CLA (≈1 mM) upon alkaline hydrolysis yielded the substrate CFA and the by-product quinic acid [58]. The total yield of the process was ≈33%.
Two types of dimers were obtained by the action of Lac on 25 mM chalcones (1,3-diphenylprop-2-en-1-ones; plant compounds, radical scavengers) synthesized from benzaldehyde (or its analogs) and acetophenone analogs (Figure 6). One of the dimers (a racemate) contained the 2,3-dihydrobenzofuran scaffold occurring in many bioactive compounds. The mixtures of dimers obtained at the ten-mg scale (15–26% yields) were separated at the analytical or preparative scale [59].
A variety of thiols were shown to undergo an analogous reaction (Figure 7). The resulting disulfides are useful in (bio)chemistry as oxidants and protein stabilizers. They were prepared from 100 mM substrates 75–95% yields on a ten-mg scale. The reactions were performed with an innovative Lac-mediator (4-phenyl urazol) system [60].

2.2. Heterocoupling

Recently, heterocoupling reactions were performed between β-lactam antibiotics [61], sulfanilamide, dapsone, or sulfamerazine [62] on one hand, and 2,5-dihydroxybenzoic acid derivatives on the other (Figure 8 and Figure 9). The latter were selected because a similar structure occurs in the antibacterial compound’s ganomycins. The products were obtained from 2 mM substrates in the ten-mg [62] to hundred-mg scale [61]. The yields were good to excellent (63–93%) except for the products from Dapson (22–28%). The hybrid antibiotics were active against some multi-drug resistant staphylococci [61,62].
Heterocoupling was also used to prepare a variety of building blocks for organic synthesis. Thus, 2-arylbenzimidazoles and 2-arylbenzothiazoles were prepared from benzaldehyde and o-phenylenediamine, or benzaldehyde and o-aminothiophenol (Figure 10a) and a great variety of analogs (Figure 10b). The benzimidazols and benzothiazols were obtained in 56–94% and 48–88% yields, respectively, from ≈0.5–0.9 M benzaldehydes. Benzimidazole is a structural motif of omeprazole and is similar to drugs used to cure gastric disorders (hyperacidity, ulcers, etc.). The compounds of this type are also precursors of antibacterial, antiviral, and antitumor drugs; antiallergens; and antihypertensives [63].
Heterocoupling of benzenediols (1,4-benzenediol or 3-substituted catechols) and benzenesulfinates led to diarylsulfones, known as structural motifs of the antibiotic dapsone, and various compounds with antifungal, antitumor, or antiviral activity. A set of these compounds were prepared in 75–95% yields on a hundred-mg scale [64] (Figure 11). Similarly, the corresponding catechol thioethers with cytostatic activity were prepared from 4-substituted catechols and pyrimidine analogs [65] (Figure 12). The concentrations of the diol substrates were ≈29 mM [65] and ≈67 mM [64]. Both types of hybrid products were prepared in good to excellent yields (75–95%) at the tens to hundred-mg scale.

2.3. Other Oxidation Reactions

Mild oxidation by Lac proved to be suitable for sensitive compounds such as 2-thiophenemethanol [66], propargyl alcohols [67], or secondary alcohols [68]. Thiophenemethanol (10 mM) was converted to 2-thiophenecarboxaldehyde by an immobilized Lac/TEMPO system on an analytical scale. Such thiocarbonyl compounds are important for the production of agricultural chemicals, pharmaceuticals, and dyes [66].
An enzyme cascade consisting of Lac and alcohol dehydrogenase (ADH) was used for the production of enantiomerically pure propargyl alcohols, which are important building blocks. Deracemization of racemic alcohols (50 mM) was performed in one pot on a semipreparative scale. Depending on the source of ADH, both enantiomers were obtained with largely excellent conversions at up to >99% enantiomeric excess (Figure 13) [67].
In addition, the synergy of Lac/TEMPO and organometallic compounds (RLi/RMgX) was used to transform a variety of secondary alcohols (0.73 M) into tertiary alcohols without isolating the intermediate ketones [68]. Under optimum conditions, the first step (oxidation by Lac) took place in an aqueous medium, while an organo-aqueous medium was suitable for the second (addition of RLi). The conversion was determined by GC and reached up to 80–91% (Figure 14).

2.4. Sources of Laccases

Most of the above reactions were performed with a Lac from Trametes. This enzyme is commercially available, e.g., from Sigma Aldrich and ASA Spezialenzyme. The latter company also produces several other Lacs from, e.g., A. bisporus. This enzyme was used in one of the above studies [65]. One of the Lacs from this company is a recombinantly produced enzyme recommended for organic synthesis (http://asa-enzyme.com/products/special-enzymes; accessed on 27 February 2022).
In addition, two Lacs from Myceliophthora thermophila (Suberase®; Denilite® II Base) were prepared by Novozymes, together with the Lac Novoprime Base 268. The last was evaluated as the best one in terms of chemoselectivity [63].
Commercial enzymes are not inexpensive. They have been used for tens to hundreds of mg-scale preparative reactions (see above), but scale-up can be costly. Some researchers have produced these enzymes themselves, such as laccase from Trametes pubescens. This extracellular enzyme was precipitated from the cultivation medium with ammonium sulfate [53]. Bacterial Lacs have rarely been used. An exception was a “small Lac“ (SLac) from Streptomyces coelicolor. The enzyme was produced in E. coli and purified [57,58].
There exist a number of other Lacs that have been studied for other applications but may also be useful for organic synthesis. For example, the cotA laccase from Bacillus subtilis is suitable for alkaline conditions. This was recently demonstrated by its use in the modification of Kraft lignins. The enzyme was produced in E. coli and purified in one step [69].
Fungi have been known as sources of Lacs for decades, but their potential has not yet been fully exploited. Namely, the understanding of Lac production in wild-type producers needs to be improved. Ganoderma lucidum, for example, contains several lac genes, whose transcription depends on pH. In the past, the focus has been on extracellular Lacs, but some fungi, including G. lucidum, may also produce intracellular Lacs. The study of these enzymes may open new perspectives in this research [70].
The heterologous expression systems suitable for Lacs were summarized in [18]. These include standard bacterial and yeast hosts but also insect cells. The order Agaricales has been less studied for Lacs than the order Polyporales [71]. A recently used system was based on Saccharomyces cerevisiae and the gene originated from the fungus Agrocybe pediades (Agaricales). A double mutant of this Lac is a promising catalyst in terms of redox potential and pH range (shifted to neutral values) [71].
Another way to improve Lacs is their immobilization. The general methodology and the uses of immobilized Lacs have been summarized previously [19]. Recent advances in this research include the use of cross-linked magnetic nanoparticles (NPs) [6] or hybrid Lac-copper phosphate and Lac-zinc phosphate catalysts [14,72]. These methods resulted in increased stability of the enzyme and allowed recycling of the catalyst. For example, the copper NPs (nanoflowers) were recycled ten times and retained over 92% activity [14]. The catalysts have been tested for the decolorization of synthetic dyes [14] or bisphenol A degradation [6], but other applications are also possible.

3. Tyrosinase-Catalyzed Reactions

3.1. Ortho-Hydroxylation of Phenols

The main application of TYR is the production of valuable benzene-1,2-diols (catechols) from the cheaper and readily available phenols. Catechols are often biologically active due to their ability to chelate metal ions and scavenge radicals. The synthesis of catechols consists of two coupled reactions—oxidation of phenols to ortho-quinones and the chemical reduction of the quinones. This one-pot pathway was established in 2001 for the production of hydroxytyrosol, a naturally occurring compound in olives [73]. Later this method and its variants were used to prepare a large number of catechols of diverse structures (see below).
The use of TYR for the preparation of hydroxytyrosol from tyrosol (16 mM) on an analytical scale was based on the combination of TYR as an oxidizing agent and ascorbic acid (30 mM) as a reducing agent [73]. The intermediate quinone is reduced to hydroxytyrosol (Figure 15a). The same principle has been used to convert 50 mM tyrosol and other phenols to the corresponding catechols on a preparative scale. The use of the soluble or immobilized TYR enabled 77–85% yields of hydroxytyrosol and almost quantitative yields of some other catechols [74]. Hydroxytyrosol is known as a radical scavenger and generally as a health-promoting compound. Hydroxytyrosol esters were also prepared from 50 mM tyrosol esters using sodium dithionite as a reducing agent, which was added after the completion of the phenol conversion (Figure 15b) [75]. In this case, the aqueous medium was replaced by an organo-aqueous medium (dichloromethane with ≈10% phosphate buffer). The enzyme had to be immobilized to work in this medium [75]. Recently, a new variant of this method was used to produce hydroxytyrosol. The reducing agent was NADH recycled using glucose dehydrogenase (GDH) (Figure 15c). This hydroxytyrosol production is competitive with, e.g., microbial hydrolysis of oleuropein or the use of cell factories (in terms of product yield and concentrations) [76].
Both the one-pot and cascade methods (Figure 15a,b) also proved to be useful for the ortho-hydroxylation of other phenol compounds (Figure 15d). These were, e.g., 4-hydroxyphenylacetic acid [80], 4-hydroxyphenylpropionic acid (phloretic acid) [80] and its esters [75], l-tyrosine [83,84] and its derivatives [77,78], 4-alkylphenols [74,80,81], 4-halophenols [74,80,82], other substituted phenols, and bisphenol A [74,80]. The targeted catechols were produced from 10–50 mM phenols in the mg to tens of mg range. The isolated yields were largely excellent and any byproducts (dimers, trihydroxyphenols) were minor [80]. Their production depended on the medium (presence of an organic solvent) and the enzyme form (soluble vs. immobilized). The organic solvent (dichloromethane) generally suppressed their formation. The cascade method was also used for the preparation of peptides, in which the tyrosine residue was converted to 3,4-dihydroxy-L-phenylalanine (DOPA) [80].
DOPA has been known for decades for its effect in the treatment of parkinsonism. The peptides from DOPA are expected to have a better effect in this sense. These peptides could also be useful in the treatment of atherosclerosis and in cosmetics [78]. The esters of hydroxytyrosol and 3,4-dihydroxyphenylpropionic (dihydrocaffeic) acid with lipophilic side chains attracted attention as potential antiviral agents [75,79]. Some alkylphenols with a short alkyl chain (e.g., methyl, ethyl) are promising as agents against oxidative stress [85]. This activity remains to be investigated in similar catechols with longer side chains.

3.2. Sources of Tyrosinases

The most commonly used TYR is derived from the common button mushroom, Agaricus bisporus, and is commercially available. Its immobilization was especially beneficial when the reaction was carried out in a largely organic medium. The catalyst was immobilized on Eupergit, and this preparation was then modified with a layer-by-layer (LbL) coating, i.e., it was covered with layers of electrolytes with opposite charges. This catalyst was recycled five times with 75% activity after the last run [74]. A variant of this method consisted of attaching TYR to carbon nanotubes followed by LbL coating [75]. It exhibited excellent stability during recycling; the yield of caffeic acid decreased from 98% to 91% in the sixth run. In another study, immobilization of a commercial TYR on a polyamide membrane enabled continuous production of DOPA in a laboratory-scale bioreactor [83].
The A. bisporus TYR can also be obtained in the laboratory from the fruiting bodies of the fungus. The methods for its production have already been discussed in relation to its potential for biodegradation [22]. This enzyme (partially purified by ammonium sulfate precipitation) proved useful for the synthesis of DOPA and DOPA peptides [78] or n-alkylcatechols [81]. It was also immobilized on Eupergit and coated by the LbL method [78], much like the commercial enzyme (see above). Determination of both commercial and crude TYR of A. bisporus indicated that the former has more than five times higher specific activity for L-tyrosine [80] or DOPA [81]. In addition, this TYR (in the form of its isoenzymes) and other fungal TYRs have been recombinantly produced [86,87,88,89,90,91,92]. However, this source is still underutilized in biocatalysis.
Bacterial TYRs were used for the production of halocatechols (TYR from Ralstonia solanacearum) [82] and hydroxytyrosol (TYR from Bacillus megaterium) [76]. Both enzymes were overproduced in E. coli and purified. In addition, artificial variants of the former enzyme were generated by single-point mutation and were found to be superior in terms of its kinetics for halophenols. The TYR of B. megaterium was used in a sol-gel immobilized form. It proved suitable for both continuous and repeated use (with an almost full conversion of tyrosol in the eighth run) [76].
There are also other TYRs that can be useful in organic synthesis. For example, the TYRs of Pholiota microspora [87,88], Streptomyces antibioticus [89] or Polyporus arcularius [90] were produced recombinantly. Although E. coli can be used, expression in this host results in an inactive pro-enzyme, which must be activated by proteolysis [87,88,90]. In contrast, the enzyme is processed in Aspergillus niger [91] or Komagataella phaffii (formerly Pichia pastoris) [92]. On the other hand, the presence of an active enzyme may be detrimental to the host, because the enzyme is capable of oxidizing Tyr residues. The TYRs of S. antibioticus, P. aucularius, and A. bisporus have been examined as models for inhibitor studies. However, they have also been shown to transform some (potentially bioactive) derivatives of aurons and may be useful for their modifications [89,90].

4. Major Challenges and Prospects

The above research addressed the major challenges of using Lac and TYR in organic synthesis. These are to prepare effective catalysts and to design selective reactions with high product yields. Some work also addressed the structure–activity relationships (SAR) of bioactive products.
A few Lacs and TYRs are readily available from commercial suppliers. These have been sufficient for small-scale syntheses and SAR studies. The immobilization of these enzymes, such as the Lac from T. toxicodendron [6,72] and T. versicolor [14], or the TYR from A. bisporus [75,78,79,80], was promising. The obtained catalysts showed improved properties in terms of thermal and pH stability resistance to solvents/detergents, shelf-life, and/or recyclability.
Some Lacs and TYRs can be prepared by simple methods for direct use. Lacs are available from the culture fluid of fungal cultures and can be used without purification [93]. However, the cultivation of fungi may be beyond the routine work of an organic chemist. In contrast, TYR can be prepared by simple extraction of fruiting bodies of the common button mushroom and, optionally, by ammonium sulfate precipitation [81].
The discovery of new Lacs and TYRs, including the hitherto poorly studied bacterial enzymes [57,58,76,82] suggests that the scope of these catalysts may be extended. Recombinant production of Lac and TYRs for organic synthesis is still poorly explored. The availability of commercial services such as gene synthesis and overexpression make this option accessible also to non-biologists.
Developing processes selected for upscaling requires overcoming additional challenges. Minimizing catalyst costs is one of the most important. Commercial enzymes are relatively expensive, although they are likely to be available in bulk quantities at reduced prices. In this context, the development of the stable immobilized catalysts mentioned above is an important advance that can be applied to other Lacs or TYRs. The combination of an overproduced thermostable enzyme with a suitable immobilization may be the best way to obtain low-cost catalysts.
The space-time yield (g product/L/h) must be optimized to make the process viable. This factor depends on the concentration and conversion of the substrate. Most of the above reactions were demonstrated with substrate concentrations of ≤50 mM. Those performed with higher substrate concentrations (0.1 M thiols [60], ≈0.5 M o-phenylenediamine, ≈0.7 M o-aminothiophenol [63], or ≈0.7 M secondary alcohol [68]) have been the exception. The space-time yield was calculated to be 0.16 g/L/h and 0.69 g/L/h for 50 mM and 10 mM substrate, respectively, for hydrotyrosol production. Thus, it was relatively low, especially at the higher substrate concentrations. This was partly due to the inhibition of the enzyme by the product. Therefore, product removal by adsorption, for example, was suggested [76].
The performance of other reactions used in the process must be taken into account, and reaction conditions must be adjusted, especially for one-pot reactions. Many of the reactions demonstrated were two- or multi-step. Virtually all phenol-catechol reactions combined oxidation by TYR with chemical reduction by ascorbic acid [73,74,81], dithionite [75,78,80] or NADH. The recycling of NADH is necessary for a viable process, and it has been achieved using an additional enzyme—GDH. Moreover, both TYR and GDH were immobilized to allow recycling [76].
The multi- and chemoenzymatic reactions were also used to increase selectivity (dimerization of glucoside followed by deglycosylation [54,55]), to enable deracemization (one-pot reaction catalyzed by Lac and ADH [67]) or to utilize a cheap substrate (dimerization of CFA obtained from waste material [58]). We assume that these processes are particularly promising, as they maximize and valorize the synthetic potential of Lacs and TYRs.

5. Conclusions

The use of Lacs and TYRs opens up simple solutions for the synthesis of valuable (hetero) aromatic compounds. The alternative chemical routes are often complex and involve protection and deprotection steps, strong acids or bases, expensive or toxic oxidants, or high temperatures. Several Lacs and the TYR from A. bisporus are available from commercial companies. This has allowed organic chemists to directly develop new “green” syntheses that yielded compounds promising as dietary supplements, pharmaceuticals, and building blocks for fine chemicals. Known bioactive compounds were modified by dimerization, oligomerization, or heterocoupling, and new ones were proposed. The enzymes have been very useful for the production of small amounts of the compounds for SAR studies. The biological tests of these products allowed the identification of structural motifs important for biological activities in some cases. The applications of the enzymes in the fine-chemical industry are plausible but it is essential to solve demanding tasks connected to scale-up.

Author Contributions

Conceptualization, L.M. and V.K.; writing—original draft preparation, L.M.; writing—review and editing, L.M. and V.K.; data curation, B.K.; project administration, L.M. and V.K.; funding acquisition, L.M. and V.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education, Youth and Sports of the Czech Republic (grant number LTC19037), Czech Science Foundation grant No. GA21-01799S, and Czech Academy of Sciences, grant number RVO61388971. The contribution of COST Action LignoCOST (CA17128 Pan-European network on sustainable lignin valorization) supported by COST (European Cooperation in Science and Technology), in promoting interaction, the exchange of knowledge, and collaborations in the field of lignin valorization is gratefully acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Petr Novotný (Institute of Microbiology) for his technical assistance with preparing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Riva, S. Laccases: Blue enzymes for green chemistry. Trends Biotechnol. 2006, 24, 219–226. [Google Scholar] [CrossRef]
  2. Chang, C.W.; Liao, J.Y.; Lu, K.T. Syntheses and characteristics of urushiol-based waterborne UV-cured wood coatings. Polymers 2021, 13, 4005. [Google Scholar] [CrossRef]
  3. Pezzella, C.; Guarino, L.; Piscitelli, A. How to enjoy laccases. Cell. Mol. Life Sci. 2015, 72, 923–940. [Google Scholar] [CrossRef]
  4. Uyama, H. Functional polymers from renewable plant oils. Polym. J. 2018, 50, 1003–1011. [Google Scholar] [CrossRef]
  5. 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]
  6. Patel, S.K.S.; Gupta, R.K.; Kim, S.Y.; Kim, I.W.; Kalia, V.C.; Lee, J.K. Rhus vernicifera laccase immobilization on magnetic nanoparticles to improve stability and its potential application in bisphenol A degradation. Indian J. Microbiol. 2021, 61, 45–54. [Google Scholar] [CrossRef] [PubMed]
  7. Mehra, R.; Muschiol, J.; Meyer, A.S.; Kepp, K.P. A structural-chemical explanation of fungal laccase activity. Sci. Rep. 2018, 8, 17285. [Google Scholar] [CrossRef]
  8. Agrawal, K.; Verma, P. Multicopper oxidase laccases with distinguished spectral properties: A new outlook. Heliyon 2020, 6, e03972. [Google Scholar] [CrossRef] [PubMed]
  9. Radveikiené, I.; Vidžiúnaité, R.; Meškiené, R.; Meškys, R.; Časaité, V. Characterization of a yellow laccase from Botrytis cinerea 241. J. Fungi 2021, 7, 143. [Google Scholar] [CrossRef]
  10. Mot, A.C.; Coman, C.; Hadade, N.; Damian, G.; Silaghi-Dumitrescu, R.; Heering, H. “Yellow” laccase from Sclerotinia sclerotiorum is a blue laccase that enhances its substrate affinity by forming a reversible tyrosyl-product adduct. PLoS ONE 2020, 15, e0225530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Mehra, R.; Kepp, K.P. Contribution of substrate reorganization energies of electron transfer to laccase activity. Phys. Chem. Chem. Phys. 2019, 21, 15805–15814. [Google Scholar] [CrossRef]
  12. Mehra, R.; Meyer, A.S.; Kepp, K.P. Molecular dynamics derived life times of active substrate binding poses explain KM of laccase mutants. RSC Adv. 2018, 8, 36915–36926. [Google Scholar] [CrossRef] [Green Version]
  13. Yoshida, H. Chemistry of lacquer (urushi). J. Chem. Soc. 1883, 43, 472–486. [Google Scholar] [CrossRef] [Green Version]
  14. Patel, S.K.S.; Otari, S.V.; Li, J.; Kim, D.R.; Kim, S.C.; Cho, B.-K.; Kalia, V.C.; Kang, Y.C.; Lee, J.-K. Synthesis of cross-linked protein-metal hybrid nanoflowers and its application in repeated batch decolorization of synthetic dyes. J. Hazard. Mater. 2018, 347, 442–450. [Google Scholar] [CrossRef]
  15. Vipotnik, Z.; Michelin, M.; Tavares, T. Rehabilitation of a historically contaminated soil by different laccases and laccase-mediator system. J. Soils Sediments 2022, 1–9. [Google Scholar] [CrossRef]
  16. Sorrentino, I.; Carriere, M.; Jamet, H.; Stanzione, I.; Piscitelli, A.; Giardina, P.; Le Goff, A. The laccase mediator system at carbon nanotubes for anthracene oxidation and femtomolar electrochemical biosensing. Analyst 2022, 147, 897–904. [Google Scholar] [CrossRef] [PubMed]
  17. 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] [Green Version]
  18. Debnath, R.; Saha, T. An insight into the production strategies and applications of the ligninolytic enzyme laccase from bacteria and fungi. Biocatal. Agric. Biotechnol. 2020, 26, 101645. [Google Scholar] [CrossRef]
  19. Khatami, S.H.; Vakili, O.; Movahedpour, A.; Ghesmati, Z.; Ghasemi, H.; Taheri-Anganeh, M. Laccase: Various types and applications. Biotechnol. Appl. Biochem. 2022. [Google Scholar] [CrossRef]
  20. Senthivelan, T.; Kanagaraj, J.; Panda, R.C. Recent trends in fungal laccase for various industrial applications: An eco-friendly approach—A review. Biotechnol. Bioproc. Eng. 2016, 21, 19–38. [Google Scholar] [CrossRef]
  21. Rostami, A.; Abdelrasoul, A.; Shokri, Z.; Shirvandi, Z. Applications and mechanisms of free and immobilized laccase in detoxification of phenolic compounds—A review. Korean J. Chem. Eng. 2022, 1–12. [Google Scholar] [CrossRef]
  22. Martínková, L.; Kotik, M.; Marková, E.; Homolka, L. Biodegradation of phenolic compounds by Basidiomycota and its phenol oxidases: A review. Chemosphere 2016, 149, 373–382. [Google Scholar] [CrossRef] [PubMed]
  23. Bilal, M.; Rasheed, T.; Nabeel, F.; Iqbal, H.M.N.; Zhao, Y.P. Hazardous contaminants in the environment and their laccase-assisted degradation—A review. J. Environ. Manag. 2019, 234, 253–264. [Google Scholar] [CrossRef] [PubMed]
  24. Barrios-Estrada, C.; Rostro-Alanis, M.D.; Muñoz-Gutiérrez, B.D.; Iqbal, H.M.N.; Kannan, S.; Parra-Saldívar, R. Emergent contaminants: Endocrine disruptors and their laccase-assisted degradation—A review. Sci. Total Environ. 2018, 612, 1516–1531. [Google Scholar] [CrossRef]
  25. Yashas, S.R.; Shivakumara, B.P.; Udayashankara, T.H.; Krishna, B.M. Laccase biosensor: Green technique for quantification of phenols in wastewater (a review). Orient. J. Chem. 2018, 34, 631–637. [Google Scholar] [CrossRef] [Green Version]
  26. Sun, K.; Li, S.Y.; Si, Y.B.; Huang, Q.G. Advances in laccase-triggered anabolism for biotechnology applications. Crit. Rev. Biotechnol. 2021, 41, 969–993. [Google Scholar] [CrossRef]
  27. Slagman, S.; Zuilhof, H.; Franssen, M.C.R. Laccase-mediated grafting on biopolymers and synthetic polymers: A critical review. ChemBioChem 2018, 19, 288–311. [Google Scholar] [CrossRef] [Green Version]
  28. Liu, Y.; Luo, G.; Ngo, H.H.; Guo, W.S.; Zhang, S.C. Advances in thermostable laccase and its current application in lignin-first biorefinery: A review. Bioresour. Technol. 2020, 298, 122511. [Google Scholar] [CrossRef]
  29. Mayolo-Deloisa, K.; González-González, M.; Rito-Palomares, M. Laccases in food industry: Bioprocessing, potential industrial and biotechnological applications. Front. Bioeng. Biotechnol. 2020, 8, 222. [Google Scholar] [CrossRef] [Green Version]
  30. Nasir, M.; Hashim, R.; Sulaiman, O.; Nordin, N.A.; Lamaming, J.; Asim, M. Laccase, an emerging tool to fabricate green composites: A review. Bioresources 2015, 10, 6262–6284. [Google Scholar] [CrossRef] [Green Version]
  31. Adamian, Y.; Lonappan, L.; Alokpa, K.; Agathos, S.N.; Cabana, H. Recent developments in the immobilization of laccase on carbonaceous supports for environmental applications—A critical review. Front. Bioeng. Biotechnol. 2021, 9. [Google Scholar] [CrossRef] [PubMed]
  32. Vieira, Y.A.; Gurgel, D.; Henriques, R.O.; Machado, R.A.F.; de Oliveira, D.; Oechsler, B.F.; Furigo, A. A perspective review on the application of polyacrylonitrile-based supports for laccase immobilization. Chem. Rec. 2021, e202100215. [Google Scholar] [CrossRef] [PubMed]
  33. Gu, Y.H.; Yuan, L.; Jia, L.N.; Xue, P.; Yao, H.Q. Recent developments of a co-immobilized laccase-mediator system: A review. RSC Adv. 2021, 11, 29498–29506. [Google Scholar] [CrossRef]
  34. Zhou, W.T.; Zhang, W.X.; Cai, Y.P. Laccase immobilization for water purification: A comprehensive review. Chem. Eng. J. 2021, 403, 126272. [Google Scholar] [CrossRef]
  35. Daronch, N.A.; Kelbert, M.; Pereira, C.S.; de Araújo, P.H.H.; de Oliveira, D. Elucidating the choice for a precise matrix for laccase immobilization: A review. Chem. Eng. J. 2020, 397, 125506. [Google Scholar] [CrossRef]
  36. Ren, D.J.; Wang, Z.B.; Jiang, S.; Yu, H.Y.; Zhang, S.Q.; Zhang, X.Q.; Taylor; Francis, L. Recent environmental applications of and development prospects for immobilized laccase: A review. Biotechnol. Genet. Eng. Rev. 2020, 36, 81–131. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, H.; Wu, X.; Sun, J.L.; Chen, S.C. Stimulation of laccase biocatalysis in ionic liquids: A review on recent progress. Curr. Prot. Pept. Sci. 2018, 19, 100–111. [Google Scholar] [CrossRef]
  38. Lei, L.L.; Yang, X.Y.; Song, Y.D.; Huang, H.; Li, Y.X. Current research progress on laccase-like nanomaterials. New J. Chem. 2022, 46, 3541–3550. [Google Scholar] [CrossRef]
  39. Baldrian, P. Fungal laccases-occurrence and properties. FEMS Microbiol. Rev. 2006, 30, 215–242. [Google Scholar] [CrossRef] [Green Version]
  40. Ramsden, C.A.; Riley, P.A. Tyrosinase: The four oxidation states of the active site and their relevance to enzymatic activation, oxidation and inactivation. Bioorg. Med. Chem. Lett. 2014, 22, 2388–2395. [Google Scholar] [CrossRef] [PubMed]
  41. Ramsden, C.A.; Riley, P.A. Mechanistic aspects of the tyrosinase oxidation of hydroquinone. Bioorg. Med. Chem. Lett. 2014, 24, 2463–2464. [Google Scholar] [CrossRef]
  42. Halaouli, S.; Asther, M.; Sigoillot, J.C.; Hamdi, M.; Lomascolo, A. Fungal tyrosinases: New prospects in molecular characteristics, bioengineering and biotechnological applications. J. Appl. Microbiol. 2006, 100, 219–232. [Google Scholar] [CrossRef] [PubMed]
  43. Ullah, S.; Son, S.; Yun, H.Y.; Kim, D.H.; Chun, P.; Moon, H.R. Tyrosinase inhibitors: A patent review (2011–2015). Expert Opin. Ther. Pat. 2016, 26, 347–362. [Google Scholar] [CrossRef]
  44. Burlando, B.; Clericuzio, M.; Cornara, L. Moraceae plants with tyrosinase inhibitory activity: A review. Mini-Rev. Med. Chem. 2017, 17, 108–121. [Google Scholar] [CrossRef]
  45. Zolghadri, S.; Bahrami, A.; Khan, M.T.H.; Muñoz-Muñoz, J.; García-Molina, F.; García-Cánovas, F.; Saboury, A.A. A comprehensive review on tyrosinase inhibitors. J. Enzym. Inhibit. Med. Chem. 2019, 34, 279–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Obaid, R.J.; Mughal, E.U.; Naeem, N.; Sadiq, A.; Alsantali, R.I.; Jassas, R.S.; Moussa, Z.; Ahmed, S.A. Natural and synthetic flavonoid derivatives as new potential tyrosinase inhibitors: A systematic review. RSC Adv. 2021, 11, 22159–22198. [Google Scholar] [CrossRef]
  47. Opperman, L.; De Kock, M.; Klaasen, J.; Rahiman, F. Tyrosinase and melanogenesis inhibition by indigenous african plants: A review. Cosmetics 2020, 7, 60. [Google Scholar] [CrossRef]
  48. Peng, Z.Y.; Wang, G.C.; Zeng, Q.H.; Li, Y.F.; Liu, H.Q.; Wang, J.J.; Zhao, Y. A systematic review of synthetic tyrosinase inhibitors and their structure-activity relationship. Crit. Rev. Food Sci. Nutr. 2021, 1–42. [Google Scholar] [CrossRef]
  49. Fernandes, M.S.; Kerkar, S. Microorganisms as a source of tyrosinase inhibitors: A review. Ann. Microbiol. 2017, 67, 343–358. [Google Scholar] [CrossRef]
  50. García-Molina, P.; García-Molina, F.; Teruel-Puche, J.A.; Rodríguez-López, J.N.; García-Cánovas, F.; Muñoz-Muñoz, J.L. Considerations about the kinetic mechanism of tyrosinase in its action on monophenols: A review. Mol. Catal. 2022, 518, 112072. [Google Scholar] [CrossRef]
  51. 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]
  52. Kudanga, T.; Nemadziva, B.; Le Roes-Hill, M. Laccase catalysis for the synthesis of bioactive compounds. Appl. Microbiol. Biotechnol. 2017, 101, 13–33. [Google Scholar] [CrossRef] [PubMed]
  53. Aruwa, C.E.; Amoo, S.O.; Koorbanally, N.; Kudanga, T. Laccase-mediated modification of isorhamnetin improves antioxidant and antibacterial activities. Process Biochem. 2022, 112, 53–61. [Google Scholar] [CrossRef]
  54. Gavezzotti, P.; Bertacchi, F.; Fronza, G.; Křen, V.; Monti, D.; Riva, S. Laccase-catalyzed dimerization of piceid, a resveratrol glucoside, and its further enzymatic elaboration. Adv. Synth. Catal. 2015, 357, 1831–1839. [Google Scholar] [CrossRef]
  55. Bassanini, I.; Gavezzotti, P.; Monti, D.; Krejzova, J.; Kren, V.; Riva, S. Laccase-catalyzed dimerization of glycosylated lignols. J. Mol. Catal. B-Enzym. 2016, 134, 295–301. [Google Scholar] [CrossRef]
  56. Muñiz-Mouro, A.; Ferreira, A.M.; Coutinho, J.A.P.; Freire, M.G.; Tavares, A.P.M.; Gullón, P.; González-García, S.; Eibes, G. Integrated biocatalytic platform based on aqueous biphasic systems for the sustainable oligomerization of rutin. ACS Sustain. Chem. Eng. 2021, 9, 9941–9950. [Google Scholar] [CrossRef]
  57. Nemadziva, B.; Le Roes-Hill, M.; Koorbanally, N.; Kudanga, T. Small laccase-catalyzed synthesis of a caffeic acid dimer with high antioxidant capacity. Process Biochem. 2018, 69, 99–105. [Google Scholar] [CrossRef]
  58. Nemadziva, B.; Ngubane, S.; Ruzengwe, F.M.; Kasumbwe, K.; Kudanga, T. Potato peels as feedstock for laccase-catalysed synthesis of phellinsin A. Biomass Convers. Biorefinery 2022. [Google Scholar] [CrossRef]
  59. Grosso, S.; Radaelli, F.; Fronza, G.; Passarella, D.; Monti, D.; Riva, S. Studies on the laccase-catalyzed oxidation of 4-hydroxy-chalcones. Adv. Synth. Catal. 2019, 361, 2696–2705. [Google Scholar] [CrossRef] [Green Version]
  60. Khaledian, D.; Rostami, A.; Zarei, S.A. Laccase-catalyzed in situ generation and regeneration of N-phenyltriazolinedione for the aerobic oxidative homo-coupling of thiols to disulfides. Catal. Commun. 2018, 114, 75–78. [Google Scholar] [CrossRef]
  61. Mikolasch, A.; Hammer, E.; Witt, S.; Lindequist, U. Laccase-catalyzed derivatization of 6-aminopenicillanic, 7-aminocephalosporanic and 7-aminodesacetoxycephalosporanic acid. AMB Express 2020, 10, 177. [Google Scholar] [CrossRef]
  62. Mikolasch, A.; Hahn, V. Laccase-catalyzed derivatization of antibiotics with sulfonamide or sulfone structures. Microorganisms 2021, 9, 2199. [Google Scholar] [CrossRef]
  63. Maphupha, M.; Juma, W.P.; de Koning, C.B.; Brady, D. A modern and practical laccase-catalysed route suitable for the synthesis of 2-arylbenzimidazoles and 2-arylbenzothiazoles. RSC Adv. 2018, 8, 39496–39510. [Google Scholar] [CrossRef] [Green Version]
  64. Habibi, D.; Rahimi, A.; Rostami, A.; Moradi, S. Green and mild laccase-catalyzed aerobic oxidative coupling of benzenediol derivatives with various sodium benzenesulfinates. Tetrahedron Lett. 2017, 58, 289–293. [Google Scholar] [CrossRef]
  65. Abdel-Mohsen, H.T.; 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]
  66. Benny, L.; Cherian, A.R.; Varghese, A.; Sangwan, N.; Avti, P.K.; Hegde, G. A novel laccase-based biocatalyst for selective electro-oxidation of 2-thiophene methanol. Mol. Catal. 2021, 516, 111999. [Google Scholar] [CrossRef]
  67. González-Granda, S.; Méndez-Sánchez, D.; Lavandera, I.; Gotor-Fernández, V. Laccase-mediated oxidations of propargylic alcohols. application in the deracemization of 1-arylprop-2-yn-1-ols in combination with alcohol dehydrogenases. ChemCatChem 2020, 12, 520–527. [Google Scholar] [CrossRef]
  68. Ramos-Martín, M.; Lecuna, R.; Cicco, L.; Vitale, P.; Capriati, V.; Rios-Lombardía, N.; González-Sabin, J.; Soto, A.P.; García-Álvarez, J. A one-pot two-step synthesis of tertiary alcohols combining the biocatalytic laccase/TEMPO oxidation system with organolithium reagents in aerobic aqueous media at room temperature. Chem. Commun. 2021, 57, 13534–13537. [Google Scholar] [CrossRef]
  69. Mayr, S.A.; Subagia, R.; Weiss, R.; Schwaiger, N.; Weber, H.K.; Leitner, J.; Ribitsch, D.; Nyanhongo, G.S.; Guebitz, G.M. Oxidation of various kraft lignins with a bacterial laccase enzyme. Int. J. Mol. Sci. 2021, 22, 13161. [Google Scholar] [CrossRef]
  70. Zhu, J.; Song, S.Q.; Lian, L.D.; Shi, L.; Ren, A.; Zhao, M.W. Improvement of laccase activity by silencing PacC in Ganoderma lucidum. World J. Microbiol. Biotechnol. 2022, 38, 32. [Google Scholar] [CrossRef]
  71. Aza, P.; Molpeceres, G.; Ruiz-Dueñas, F.J.; Camarero, S. Heterologous expression, engineering and characterization of a novel laccase of Agrocybe pediades with promising properties as biocatalyst. J. Fungi 2021, 7, 359. [Google Scholar] [CrossRef] [PubMed]
  72. Patel, S.K.S.; Kalia, V.C.; Kim, S.Y.; Lee, J.K.; Kim, I.W. Immobilization of laccase through inorganic-protein hybrids using various metal ions. Indian J. Microbiol. 2022, 1–5. [Google Scholar] [CrossRef]
  73. Espín, J.C.; Soler-Rivas, C.; Cantos, E.; Tomas-Barberán, F.A.; Wichers, H.J. Synthesis of the antioxidant hydroxytyrosol using tyrosinase as biocatalyst. J. Agric. Food Chem. 2001, 49, 1187–1193. [Google Scholar] [CrossRef] [PubMed]
  74. Guazzaroni, M.; Crestini, C.; Saladino, R. Layer-by-Layer coated tyrosinase: An efficient and selective synthesis of catechols. Bioorg. Med. Chem. 2012, 20, 157–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Botta, G.; Bizzarri, B.M.; Garozzo, A.; Timpanaro, R.; Bisignano, B.; Amatore, D.; Palamara, A.T.; Nencioni, L.; Saladino, R. Carbon nanotubes supported tyrosinase in the synthesis of lipophilic hydroxytyrosol and dihydrocaffeoyl catechols with antiviral activity against DNA and RNA viruses. Bioorg. Med. Chem. 2015, 23, 5345–5351. [Google Scholar] [CrossRef]
  76. Deri-Zenaty, B.; Bachar, S.; Rebros, M.; Fishman, A. A coupled enzymatic reaction of tyrosinase and glucose dehydrogenase for the production of hydroxytyrosol. Appl. Microbiol. Biotechnol. 2020, 104, 4945–4955. [Google Scholar] [CrossRef]
  77. Bizzarri, B.M.; Pieri, C.; Botta, G.; Arabuli, L.; Mosesso, P.; Cinelli, S.; Schinoppi, A.; Saladino, R. Synthesis and antioxidant activity of DOPA peptidomimetics by a novel IBX mediated aromatic oxidative functionalization. RSC Adv. 2015, 5, 60354–60364. [Google Scholar] [CrossRef]
  78. Botta, G.; Delfino, M.; Guazzaroni, M.; Crestini, C.; Onofri, S.; Saladino, R. Selective synthesis of DOPA and DOPA peptides by native and immobilized tyrosinase in organic solvent. ChemPlusChem 2013, 78, 325–330. [Google Scholar] [CrossRef] [Green Version]
  79. Bozzini, T.; Botta, G.; Delfino, M.; Onofri, S.; Saladino, R.; Amatore, D.; Sgarbanti, R.; Nencioni, L.; Palamara, A.T. Tyrosinase and Layer-by-Layer supported tyrosinases in the synthesis of lipophilic catechols with antiinfluenza activity. Bioorg. Med. Chem. 2013, 21, 7699–7708. [Google Scholar] [CrossRef]
  80. Guazzaroni, M.; Pasqualini, M.; Botta, G.; Saladino, R. A Novel synthesis of bioactive catechols by layer-by-layer immobilized tyrosinase in an organic solvent medium. ChemCatChem 2012, 4, 89–99. [Google Scholar] [CrossRef]
  81. Martínková, L.; Příhodová, R.; Kulik, N.; Pelantová, H.; Křístková, B.; Petrásková, L.; Biedermann, D. Biocatalyzed reactions towards functional food components 4-alkylcatechols and their analogues. Catalysts 2020, 10, 1077. [Google Scholar] [CrossRef]
  82. Davis, R.; Molloy, S.; Quigley, B.; Nikodinovic-Runic, J.; Solano, F.; O’Connor, K.E. Biocatalytic versatility of engineered and wild-type tyrosinase from R. solanacearum for the synthesis of 4-halocatechols. Appl. Microbiol. Biotechnol. 2018, 102, 5121–5131. [Google Scholar] [CrossRef] [PubMed]
  83. Algieri, C.; Donato, L.; Bonacci, P.; Giorno, L. Tyrosinase immobilised on polyamide tubular membrane for the l-DOPA production: Total recycle and continuous reactor study. Biochem. Eng. J. 2012, 66, 14–19. [Google Scholar] [CrossRef]
  84. Donato, L.; Algieri, C.; Rizzi, A.; Giorno, L. Kinetic study of tyrosinase immobilized on polymeric membrane. J. Membr. Sci. 2014, 454, 346–350. [Google Scholar] [CrossRef]
  85. Senger, D.R.; Li, D.; Jaminet, S.C.; Cao, S.G. Activation of the Nrf2 cell defense pathway by ancient foods: Disease prevention by important molecules and microbes lost from the modern western diet. PLoS ONE 2016, 11, e0148042. [Google Scholar] [CrossRef]
  86. Lezzi, C.; Bleve, G.; Spagnolo, S.; Perrotta, C.; Grieco, F. Production of recombinant Agaricus bisporus tyrosinase in Saccharomyces cerevisiae cells. J. Ind. Microbiol. Biotechnol. 2012, 39, 1875–1880. [Google Scholar] [CrossRef]
  87. Kawamura-Konishi, Y.; Maekawa, S.; Tsuji, M.; Goto, H. C-terminal processing of tyrosinase is responsible for activation of Pholiota microspora proenzyme. Appl. Microbiol. Biotechnol. 2011, 90, 227–234. [Google Scholar] [CrossRef]
  88. Moe, L.L.; Maekawa, S.; Kawamura-Konishi, Y. The pro-enzyme C-terminal processing domain of Pholiota nameko tyrosinase is responsible for folding of the N-terminal catalytic domain. Appl. Microbiol. Biotechnol. 2015, 99, 5499–5510. [Google Scholar] [CrossRef]
  89. Haudecoeur, R.; Gouron, A.; Dubois, C.; Jamet, H.; Lightbody, M.; Hardré, R.; Milet, A.; Bergantino, E.; Bubacco, L.; Belle, C.; et al. Investigation of binding-site homology between mushroom and bacterial tyrosinases by using aurones as effectors. ChemBioChem. 2014, 15, 1325–1333. [Google Scholar] [CrossRef]
  90. Marková, E.; Kotik, M.; Křenková, A.; Man, P.; Haudecoeur, R.; Boumendjel, A.; Hardré, R.; Mekmouche, Y.; Courvoisier-Dezord, E.; Réglier, M.; et al. Recombinant tyrosinase from Polyporus arcularius: Overproduction in escherichia coli, characterization, and use in a study of aurones as tyrosinase effectors. J. Agric. Food Chem. 2016, 64, 2925–2931. [Google Scholar] [CrossRef]
  91. Halaouli, S.; Record, E.; Casalot, L.; Hamdi, M.; Sigoillot, J.C.; Asther, M.; Lomascolo, A. Cloning and characterization of a tyrosinase gene from the white-rot fungus Pycnoporus sanguineus, and overproduction of the recombinant protein in Aspergillus niger. Appl. Microbiol. Biotechnol. 2006, 70, 580–589. [Google Scholar] [CrossRef] [PubMed]
  92. Westerholm-Parvinen, A.; Selinheimo, E.; Boer, H.; Kalkkinen, N.; Mattinen, M.; Saloheimo, M. Expression of the Trichoderma reesei tyrosinase 2 in Pichia pastoris: Isotopic labeling and physicochemical characterization. Protein Expr. Purif. 2007, 55, 147–158. [Google Scholar] [CrossRef] [PubMed]
  93. Uhnáková, B.; Ludwig, R.; Pěknicová, J.; Homolka, L.; Lisá, L.; Šulc, M.; Petříčková, A.; Elzeinová, F.; Pelantová, H.; Monti, D.; et al. Biodegradation of tetrabromobisphenol A by oxidases in basidiomycetous fungi and estrogenic activity of the biotransformation products. Bioresour. Technol. 2011, 102, 9409–9415. [Google Scholar] [CrossRef] [PubMed]
Figure 2. Dimerization of piceid (2.56 mmol) by laccase (18 mg; 86 U) from Trametes versicolor followed by deglucosylation of the dimer with β-glucosidase from almonds, or cellulase from Trichoderma viride [54]. Conditions: (i) acetate buffer (pH 5.0): MeOH, 71:29; 25 °C, 3.5 h (laccase); (ii) acetate buffer (pH 5.0): MeOH, 88:12, 25 °C (β-glucosidase), or acetate buffer (pH 5.0): MeOH, 93:7; 30 °C (cellulase).
Figure 2. Dimerization of piceid (2.56 mmol) by laccase (18 mg; 86 U) from Trametes versicolor followed by deglucosylation of the dimer with β-glucosidase from almonds, or cellulase from Trichoderma viride [54]. Conditions: (i) acetate buffer (pH 5.0): MeOH, 71:29; 25 °C, 3.5 h (laccase); (ii) acetate buffer (pH 5.0): MeOH, 88:12, 25 °C (β-glucosidase), or acetate buffer (pH 5.0): MeOH, 93:7; 30 °C (cellulase).
Ijms 23 03462 g002
Figure 3. Dimerization of d-glucosides of the lignols coumaryl alcohol (R = H; 0.2 mmol) and coniferyl alcohol (R = OMe; 1.4 mmol) by laccase (0.32 and 2.35 U, respectively) from Trametes versicolor. l-Glucoside (0.89 mmol) or rutinoside (0.46 mmol) of coniferyl alcohol reacted analogously [55]. Conditions: acetate buffer (pH 5.0), 30 °C, 6–8 h.
Figure 3. Dimerization of d-glucosides of the lignols coumaryl alcohol (R = H; 0.2 mmol) and coniferyl alcohol (R = OMe; 1.4 mmol) by laccase (0.32 and 2.35 U, respectively) from Trametes versicolor. l-Glucoside (0.89 mmol) or rutinoside (0.46 mmol) of coniferyl alcohol reacted analogously [55]. Conditions: acetate buffer (pH 5.0), 30 °C, 6–8 h.
Ijms 23 03462 g003
Figure 4. Dimerization of isorhamnetin (12.5 mM) by laccase from Trametes pubescens [53]. Conditions: acetate buffer (pH 5.0): EtOH, 1:1; 37 °C, 3 h.
Figure 4. Dimerization of isorhamnetin (12.5 mM) by laccase from Trametes pubescens [53]. Conditions: acetate buffer (pH 5.0): EtOH, 1:1; 37 °C, 3 h.
Ijms 23 03462 g004
Figure 5. Production of phellinsin A, i.e., caffeic acid (CFA) dimer, from chlorogenic acid (CLA) [58]. Conditions: (i) 1.8 M NaOH, 0.1% ascorbic acid, 10 mM EDTA, r.t., 20 min; (ii) extraction of CLA with ethyl acetate (pH 3.0, saturation with NaCl); (iii) laccase (2 × 0.64 U/mL, phosphate buffer (pH 7.5): ethyl acetate, 1:4, 25 °C, 3 h. Optionally, CLA was obtained from potato peel. The yield of phellinsin A is related to CLA.
Figure 5. Production of phellinsin A, i.e., caffeic acid (CFA) dimer, from chlorogenic acid (CLA) [58]. Conditions: (i) 1.8 M NaOH, 0.1% ascorbic acid, 10 mM EDTA, r.t., 20 min; (ii) extraction of CLA with ethyl acetate (pH 3.0, saturation with NaCl); (iii) laccase (2 × 0.64 U/mL, phosphate buffer (pH 7.5): ethyl acetate, 1:4, 25 °C, 3 h. Optionally, CLA was obtained from potato peel. The yield of phellinsin A is related to CLA.
Ijms 23 03462 g005
Figure 6. (a) Synthesis of 4-hydroxychalcones from acetophenone derivative and 4-hydroxybenzaldehyde or its derivative (2 mmol each) and (b) homocoupling of selected 4-hydroxychalcone (0.89 mmol) by laccase (20 mg, 90 U) from Trametes versicolor [59]. Conditions: (a) ≈0.6 M NaOH, EtOH; r.t., ≈20 h, or ≈0.5 M H2SO4, MeOH, 80 °C, 48 h; (b) acetate buffer (pH 4.5): acetone, 1:1; 27 °C, 6 h.
Figure 6. (a) Synthesis of 4-hydroxychalcones from acetophenone derivative and 4-hydroxybenzaldehyde or its derivative (2 mmol each) and (b) homocoupling of selected 4-hydroxychalcone (0.89 mmol) by laccase (20 mg, 90 U) from Trametes versicolor [59]. Conditions: (a) ≈0.6 M NaOH, EtOH; r.t., ≈20 h, or ≈0.5 M H2SO4, MeOH, 80 °C, 48 h; (b) acetate buffer (pH 4.5): acetone, 1:1; 27 °C, 6 h.
Ijms 23 03462 g006
Figure 7. (a) Dimerization of thiophenols (1 mmol) by laccase from Trametes versicolor (46 mg, 40 U) [60]. (b) Other thiols react analogously. Conditions: phosphate buffer (pH 5.0), MeCN, 4:1, 4-phenyl urazole (10% mol., catalyst), r.t., 3–15 h.
Figure 7. (a) Dimerization of thiophenols (1 mmol) by laccase from Trametes versicolor (46 mg, 40 U) [60]. (b) Other thiols react analogously. Conditions: phosphate buffer (pH 5.0), MeCN, 4:1, 4-phenyl urazole (10% mol., catalyst), r.t., 3–15 h.
Ijms 23 03462 g007
Figure 8. Heterocoupling of (a) 6-aminopenicillanic acid (6-APA) or (b) 7-aminocephalosporanic acid (7-ACA) and 7-aminodesacetoxycephalosporanic acid (7-ADCA) with 2,5-dihydroxybenzoic acid derivatives (1.2 mmol each) [61] by laccase (528 U) from Trametes sp. (Spezialenzyme, Wolfenbüttel, Germany). Conditions: acetate buffer (pH 5.6), r.t., 3 h.
Figure 8. Heterocoupling of (a) 6-aminopenicillanic acid (6-APA) or (b) 7-aminocephalosporanic acid (7-ACA) and 7-aminodesacetoxycephalosporanic acid (7-ADCA) with 2,5-dihydroxybenzoic acid derivatives (1.2 mmol each) [61] by laccase (528 U) from Trametes sp. (Spezialenzyme, Wolfenbüttel, Germany). Conditions: acetate buffer (pH 5.6), r.t., 3 h.
Ijms 23 03462 g008
Figure 9. Heterocoupling of (a) sulfanilamide, dapsone, or (b) sulfamerazine with 2,5-dihydroxybenzoic acid derivatives (0.12 mmol each) [62] by laccase (53 U) from Trametes sp. (Spezialenzyme, Wolfenbüttel, Germany). Conditions: acetate buffer (pH 5.6), r.t., 6 h.
Figure 9. Heterocoupling of (a) sulfanilamide, dapsone, or (b) sulfamerazine with 2,5-dihydroxybenzoic acid derivatives (0.12 mmol each) [62] by laccase (53 U) from Trametes sp. (Spezialenzyme, Wolfenbüttel, Germany). Conditions: acetate buffer (pH 5.6), r.t., 6 h.
Ijms 23 03462 g009
Figure 10. (a) Heterocoupling of benzaldehyde (20 mmol) and o-phenylenediamine (10 mmol) by laccase Novoprime Base 268 (0.105 g), or benzaldehyde (10 mmol) and o-aminothiophenol (15 mmol) by laccase Suberase® (2 mL) [63]; (b) Some derivatives of benzaldehyde and o-phenylenediamine react analogously. Conditions: acetate buffer (pH 4.0): MeCN, 1:1; r.t., 2–24 h.
Figure 10. (a) Heterocoupling of benzaldehyde (20 mmol) and o-phenylenediamine (10 mmol) by laccase Novoprime Base 268 (0.105 g), or benzaldehyde (10 mmol) and o-aminothiophenol (15 mmol) by laccase Suberase® (2 mL) [63]; (b) Some derivatives of benzaldehyde and o-phenylenediamine react analogously. Conditions: acetate buffer (pH 4.0): MeCN, 1:1; r.t., 2–24 h.
Ijms 23 03462 g010
Figure 11. Heterocoupling of benzenediols and benzenesulfinates and (a) substituted catechols or (b) 1,4-benzenediol (1 mmol each) by laccase from Trametes versicolor (75.5 mg, 40 U) [64]. Conditions: phosphate buffer (pH 5.0); r.t., 18 h.
Figure 11. Heterocoupling of benzenediols and benzenesulfinates and (a) substituted catechols or (b) 1,4-benzenediol (1 mmol each) by laccase from Trametes versicolor (75.5 mg, 40 U) [64]. Conditions: phosphate buffer (pH 5.0); r.t., 18 h.
Ijms 23 03462 g011
Figure 12. Hetero-coupling of benzenediols (0.58 mmol) and (a) 2,3-dihydro-2-thioxopyrimidin-4(1H)-ones (0.5 mmol) or (b) 2,3,6,7-tetrahydro-2-thioxo-1H-cyclopenta[d]pyrimidin-4(5H)-ones (0.5 mmol) by laccase (10 mg, 12 U) from Agaricus bisporus [65]. Conditions: phosphate buffer (pH 6.0), 10% EtOH; r.t., 12–20 h.
Figure 12. Hetero-coupling of benzenediols (0.58 mmol) and (a) 2,3-dihydro-2-thioxopyrimidin-4(1H)-ones (0.5 mmol) or (b) 2,3,6,7-tetrahydro-2-thioxo-1H-cyclopenta[d]pyrimidin-4(5H)-ones (0.5 mmol) by laccase (10 mg, 12 U) from Agaricus bisporus [65]. Conditions: phosphate buffer (pH 6.0), 10% EtOH; r.t., 12–20 h.
Ijms 23 03462 g012
Figure 13. Oxidation of sensitive alcohols by laccase: (a) Oxidation of propargyl alcohols (0.05 mmol) by laccase from Trametes versicolor (10 mg, 5U) and subsequent reduction of the products by alcohol dehydrogenase (ADH) RasADH from Ralstonia sp. or evo-1.1.200 (Evoxx technologies, Düsseldorf, Germany) [67]. (b) Substrates undergoing the same reaction. Conditions: (i) Citrate buffer (pH 5.0), TEMPO (0.015 mmol), 10% tert-butyl methyl ether, 30 °C, 16 h; (ii) ADH, NAD(P)H, glucose dehydrogenase GCH-105 (Codexis Inc. Redwood, CA, USA), pH 7.5; 30 °C, 24 h.
Figure 13. Oxidation of sensitive alcohols by laccase: (a) Oxidation of propargyl alcohols (0.05 mmol) by laccase from Trametes versicolor (10 mg, 5U) and subsequent reduction of the products by alcohol dehydrogenase (ADH) RasADH from Ralstonia sp. or evo-1.1.200 (Evoxx technologies, Düsseldorf, Germany) [67]. (b) Substrates undergoing the same reaction. Conditions: (i) Citrate buffer (pH 5.0), TEMPO (0.015 mmol), 10% tert-butyl methyl ether, 30 °C, 16 h; (ii) ADH, NAD(P)H, glucose dehydrogenase GCH-105 (Codexis Inc. Redwood, CA, USA), pH 7.5; 30 °C, 24 h.
Ijms 23 03462 g013
Figure 14. One-pot cascade from secondary to tertiary alcohols. (a) Conversion of 1-phenylpropan-1-ol (0.365 or 0.73 mmol) to tertiary alcohols using laccase from Trametes versicolor (140 U) and RLi reagents [68]. Conditions: (i) water, TEMPO (10 mol%), r.t., 24 h; (ii) water: cyclopentyl methyl ether (CPME), 1:1 (v/v), RLi (3 eq.), r.t., 3 s; (b) Secondary alcohols undergoing an analogous reaction with PhLi.
Figure 14. One-pot cascade from secondary to tertiary alcohols. (a) Conversion of 1-phenylpropan-1-ol (0.365 or 0.73 mmol) to tertiary alcohols using laccase from Trametes versicolor (140 U) and RLi reagents [68]. Conditions: (i) water, TEMPO (10 mol%), r.t., 24 h; (ii) water: cyclopentyl methyl ether (CPME), 1:1 (v/v), RLi (3 eq.), r.t., 3 s; (b) Secondary alcohols undergoing an analogous reaction with PhLi.
Ijms 23 03462 g014
Figure 15. ortho-Hydroxylation of tyrosol by tyrosinase. (a) One-pot reaction of tyrosol (16 mM) with tyrosinase (30 μg/mL) from Agaricus bisporus (Sigma) [73]. Conditions: phosphate buffer (pH 6.5), 30 mM ascorbic acid (AA); DHAA, dehydroascorbic acid. (b) Cascade reaction of tyrosol esters (0.05 mmol) using immobilized tyrosinase (240 U) from A. bisporus. Conditions: (i) Buffer (pH 7.0): CH2Cl2, 1:10, 25 °C, 24 h; (ii) Na2S2O4 (in THF:H2O, 1:1) [75]. (c) One-pot reaction of tyrosol (1 mM, 0.35 mmol) with immobilized tyrosinase (0.7 g) from Bacillus megaterium. Conditions: Buffer (pH 7.5), 7 g immobilized glucose dehydrogenase (GDH), 0.4 mM Cu2+, 100 mM d-glucose, 5 mM NAD+; 25 °C, ≈2 h [76]. (d) Phenol compounds undergoing analogous reactions according to (a) and/or (b) [74,75,77,78,79,80,81,82].
Figure 15. ortho-Hydroxylation of tyrosol by tyrosinase. (a) One-pot reaction of tyrosol (16 mM) with tyrosinase (30 μg/mL) from Agaricus bisporus (Sigma) [73]. Conditions: phosphate buffer (pH 6.5), 30 mM ascorbic acid (AA); DHAA, dehydroascorbic acid. (b) Cascade reaction of tyrosol esters (0.05 mmol) using immobilized tyrosinase (240 U) from A. bisporus. Conditions: (i) Buffer (pH 7.0): CH2Cl2, 1:10, 25 °C, 24 h; (ii) Na2S2O4 (in THF:H2O, 1:1) [75]. (c) One-pot reaction of tyrosol (1 mM, 0.35 mmol) with immobilized tyrosinase (0.7 g) from Bacillus megaterium. Conditions: Buffer (pH 7.5), 7 g immobilized glucose dehydrogenase (GDH), 0.4 mM Cu2+, 100 mM d-glucose, 5 mM NAD+; 25 °C, ≈2 h [76]. (d) Phenol compounds undergoing analogous reactions according to (a) and/or (b) [74,75,77,78,79,80,81,82].
Ijms 23 03462 g015
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martínková, L.; Křístková, B.; Křen, V. Laccases and Tyrosinases in Organic Synthesis. Int. J. Mol. Sci. 2022, 23, 3462. https://doi.org/10.3390/ijms23073462

AMA Style

Martínková L, Křístková B, Křen V. Laccases and Tyrosinases in Organic Synthesis. International Journal of Molecular Sciences. 2022; 23(7):3462. https://doi.org/10.3390/ijms23073462

Chicago/Turabian Style

Martínková, Ludmila, Barbora Křístková, and Vladimír Křen. 2022. "Laccases and Tyrosinases in Organic Synthesis" International Journal of Molecular Sciences 23, no. 7: 3462. https://doi.org/10.3390/ijms23073462

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