Heterologous Expression and Production of Oxidoreductase Enzymes: Comparison
Please note this is a comparison between Version 2 by Camila Xu and Version 1 by Radivoje Prodanović.

Enzymes are biocatalysts with complex structures and specific catalytic mechanisms that determine their distinctive properties, such as high catalytic activity and selectivity of specific substrates. Oxidoreductase (OXR) enzymes are in high demand for biocatalytic applications in the food industry and cosmetics (glucose oxidase (GOx) and cellobiose dehydrogenase (CDH)), bioremediations (horseradish peroxidase (HRP) and laccase (LAC)), and medicine for biosensors and miniature biofuel cells (GOx, CDH, LAC, and HRP). Therefore, scientists are still trying to find optimal fermentation formulas and, most recently, also using protein engineering and directed evolution for an additional increase in the yield of recombinant enzyme production. 

  • recombinant
  • oxidoreductase
  • expression
  • yeasts
  • directed evolution
  • high-throughput screening

1. Introduction

Enzymes are biocatalysts with complex structures and specific catalytic mechanisms that determine their distinctive properties, such as high catalytic activity and selectivity of specific substrates. According to the BRENDA database, oxidation–reduction reactions constitute at least thirty percent of all enzymatic reactions; given that fact, oxidoreductase (OXR) occupies a special place among biocatalysts [1,2,3][1][2][3]. These enzymes catalyze the transfer of electrons from an electron donor to an electron acceptor molecule. Different cofactors, such as heme, flavin, and metal ions, are necessary for OXR catalytic activity [4]. That usually complicates their expression. The abundance of these enzymes is versatile, and the source of OXR defines their biological functions. These enzymes act as efficient biocatalysts in various processes and fields of biotechnology and have a wide range of applications in the degradation of xenobiotic compounds, the design of biosensors for environmental or medical purposes, the food and textile industry, and other fields [3].

1.1. Glucose Oxidase (GOx)

Glucose oxidase (EC 1.1.3.4) is an enzyme belonging to the OXR group. This flavoprotein uses molecular oxygen as an electron acceptor in a two-step reaction to catalyze the oxidation of β-D-glucose to D-glucono-δ-lactone and H2O2. In the first half of the reaction, namely reduction, GOx catalyzes the oxidation of β-D-glucose to D-glucono-δ-lactone, which is then hydrolyzed to gluconic acid. During the first phase cofactor of GOx, flavine adenine dinucleotide (FAD) is reduced to FADH2. In the second phase, the oxidative half-reaction, oxygen reoxidizes the reduced GOx to produce H2O2 and FADH2 oxidizes to FAD [5]. In 1999, the structure of glucose oxidase from Aspergillus was confirmed, which determined that GOx consists of two uniform subunits and that both subunits contain two separate domains: one is not covalently bound with FAD and the second attaches to the substrate. Structurally, the first domain is mainly a β-sheet and the second domain consists of four α-helices and an antiparallel β-sheet [6,7][6][7]. GOx is an enzyme that is considered an “ideal enzyme”. It is often called an oxidase “Ferrari” because it has high activity, stability, and specificity and can be used in various biotechnological, medical, and industrial applications [8,9][8][9].

1.2. Cellobiose Dehydrogenase (CDH)

Cellobiose dehydrogenase (CDH; EC1.1.99.18; CAZy AA3_1) belongs to the OXR enzyme family [10]. Many species of wood-decaying fungi produce this glycosylated enzyme involved in lignocellulose degradation [11]. CDH is usually monomeric and consists of two domains, a C-terminal cytochrome-binding fragment (CYT) and a catalytic flavin-containing dehydrogenase domain (DH), which are connected via a linear papain-sensitive linker peptide [12]. The catalytic mechanism of CDH is comprised of an oxidative and reductive half-reaction. The substrate is converted into appropriate lactone during the oxidative half-reaction while FAD is reduced to FADH2. Subsequently, electrons are transferred to the electron acceptor [13]. Intensive investigation has been carried out in the last decade on this enzyme because of its possible application in many fields like biosensors, biofuel cells, bioremediation, and clinical application [10,14][10][14].

1.3. Horseradish Peroxidase (HRP)

Horseradish peroxidase (EC 1.11.1.7) belongs to the group of OXR. This glycoprotein, with 44 kDa, contains at least 15 distinct isoenzyme forms, the most common and, thus, most studied of which is isoenzyme C1A [15]. It was determined that the tertiary structure of HRP includes two domains formed by ten α-helices, four disulfide bonds, two Ca2+ ions per molecule, and the prosthetic heme group [16,17,18][16][17][18]. The reaction between H2O2 and Fe(III) in the active center is the first phase of the catalytic cycle. As a result of this phase, a high oxidation state intermediate consisting of a Fe(IV) oxoferryl center and porphyrin-based cation radical, called Compound I, is generated. Compound II, which represents Fe(IV) oxoferryl species, is generated in the first one-electron reduction step and reduced in the second step by producing an enzyme resting state. The excess hydrogen peroxide reacts with the resting state enzyme and, as a product, compound III is obtained [19]. Due to the increasing possibilities for the applications of this enzyme in biotechnology and other branches of industry, solving the problem of recombinant production of this enzyme represents a significant challenge in science.

1.4. Laccase (LAC)

Laccases (benzenediol: oxygen oxidoreductase; p-diphenol oxidase EC 1.10.3.2) belong to a large family of multicopper oxidases. Like all multicopper oxidases, LAC also possess a relatively uncomplicated 3D structure, mainly comprising beta sheets and thurns, including a small cupredoxin-like domain. They are glycoproteins with four Cu atoms per monomer. These four Cu atoms form the catalytic core, which helps the enzyme catalyze the redox reaction. LAC couples the four single electron oxidations of a reducing substrate to the four-electron reduction cleavage of the dioxygen bond using four Cu atoms [20]. LACs are some of the oldest and most widespread enzymes in nature, and they can be found in fungi, bacteria, plants, and animals, and their function depends on the biological source. The reactions performed by LAC include the rupture of alkyl-aryl bonds, the oxidation of benzyl alcohol, and the rupture of aromatic rings that generate a wide variety of oxidized phenolic compounds. In addition, in vitro studies have shown that LACs are capable of polymerization, depolymerization, methylation, and demethylation reactions, as well as oxidation of o- and p-diphenols, aminophenols, polyphenols, polyamines, aryl-amines, and several other phenolic compounds [21,22,23,24][21][22][23][24]. The scope of laccase-catalyzed reactions can be expanded using mediators. The wide substrate spectrum and molecular oxygen as a final electron acceptor with the water molecule as the only by-product make LAC “eco-friendly” and attractive for various biotechnological industries. These biocatalysts have several bioremediation and biodegradation applications in numerous industries (food, cosmetics, nanobiotechnological, textile, woodworking, and pulp/paper) [20,25,26][20][25][26].
The production of oxidoreductases from native sources cannot meet the high market demand due to low yields and the incompatibility of the standard industrial fermentation processes with the conditions required for the growth of many microorganisms [6]. Recombinant technologies can be used to achieve higher yields of these enzymes. The diversity and upscaling possibilities of heterologous protein expression opened new commercial opportunities for their industrial uses [6].

2. Heterologous Expression

Various expression systems exist for recombinant protein production in bacteria, yeasts, insects, and mammalian cells. All of them have their drawbacks and advantages. Yeasts are interesting and versatile hosts caused by benefits such as growth speed, simple genetic manipulations, secretory expression, post-translational modification, scalable fermentation, high biomass concentrations, and safe, pathogen-free production [27,28][27][28]. There are two large yeast expression systems: methylotroph and non-methylotroph. Typical examples of non-methylotroph and methylotroph yeasts are Saccharomyces cerevisiae and Pichia pastoris.

2.1. Saccharomyces Cerevisiae

Saccharomyces cerevisiae, known as baker’s yeast, was the first eukaryotic organism with a completely sequenced genome [29]. As said before, S. cerevisiae belongs to the non-methylotroph yeast group. It was initially developed as a replacement host for producing a recombinant protein that could not be expressed in bacterial cells [27,30][27][30]. The native resistance of S. cerevisae to low pH, high osmolality, and numerous inhibitors (acetic acid, furfural, vanillin, etc.) allows low-cost and facile fermentation procedures with high biomass concentrations under aerobic and anaerobic conditions, thereby enabling a higher yield of recombinant proteins [31]. Plentiful genetic tools are developed for expression in S. cerevisiae, such as recombinant protein expression controlled by strong constitutive promoters like TEF1 and GAP or inducible promoters like galactose-inducible GAL1 [32]. Unfortunately, this expression host has some disadvantages such as a low level of protein secretion, hyperglycosylation [33], and proteolytic degradation of expressed proteins. From a biotechnological point of view, this expression system is “generally recognized as safe” (GRAS) because it is nonpathogenic and has a history in the food and pharmaceutical industry [27]. With the intention of achieving a higher expression level, a few approaches were used, such as various fermentation conditions, optimization of the codon, strong promotors and terminators, and a multi-copy expression vector—Table 1 [27,34,35][27][34][35].
Table 1. Overview of the fermentation experiments and parameters used to optimize recombinant GOx, CDH, LAC, and HRP expression in S. cerevisiae, including the fermentation yield of the enzyme.
45]. A recent study examined a few factors affecting LAC expression in S. cerevisiae. It concluded that expression depends on the gene source, used construct, temperature, pH value of cultivation media, and copper concentration. This research has provided evidence for adding alanine to maintain the pH value of cultivation media because it greatly impacts laccase expression. Also, they proved that lowering the temperature during expression from 30 to 20 °C yielded more than double the laccase activity in the identical cultivation period [49]. Several studies suggest that the secretion signal could greatly impact LAC expression in S. cerevisiae, so Aza et al. successfully used an improved α9H2 signal sequence in their work [47,52][47][52]. As was the case in other enzymes expressed in yeasts, the structure of the carbohydrate component of the enzyme synthesized by yeast differs from the structure of carbohydrate component of the native enzyme. Usually, hyperglycosylation in yeast, especially in S. cerevisiae, occurs (up to 50% per mass) that can increase stability but also decrease the specific activity of the recombinant enzyme.

2.2. Pichia Pastoris

Pichia pastoris (recently renamed as Komagataella phaffii) has gained an important role in recombinant protein production in the past several decades. P. pastoris can use methanol as a sole carbon source, given the fact it belongs to methylotroph yeasts. Undemanding genetic manipulation, high cell culture density, the capability of secreting recombinant proteins into the cell culture medium, unexacting purification of secreted proteins, eukaryotic post-translational modification, and stability of genetic constructs make this host convenient for the recombinant expression of proteins [53,54][53][54]. Promoters that control the heterologous expression of proteins could be constitutive (glyceraldehyde-3-phosphate dehydrogenase and PGAP) or inducible (alcohol oxidase, AOX1, and AOX2) [55]. Considering the existence of PAOX1 and PAOX2 in this host genome, there are three possible phenotypes of P. pastoris, specifically Mut + (wild-type methanol utilization; both of the alcohol oxidase enzymes, Aox1p and Aox2p, are functional), Muts (slow methanol utilization; disrupted Aox1p and functional Aox2p), and Mut- (no methanol utilization; both Aox1p and Aox2p are disrupted) [29]. According to Walsh and Walsh (2022), P. pastoris takes a superior position regarding S. cerevisiae in the field of recombinant protein production; the reason for that probably lies in the tightly regulated expression of both the intracellular and extracellular recombinant protein, achieving high cell density caused by the aerobic process of respiration and glycosylation like in eukaryotic cells [28,56][28][56]. P. pastoris is “generally recognized as safe” (GRAS) by the Food and Drug Administration (FDA) and suitable for diverse biotechnological applications. Several strategies can be applied to achieve a high yield of recombinant proteins. The strategies cover codon optimization, choice of suitable host strains and expression vectors, gene copy number, insertion of the gene of interest under the control of the strong promoter and appropriate signal sequence, and optimization of fermentation conditions (temperature, incubation period, agitation, carbon source, and concentration of inducer)—Table 2 [28,57,58][28][57][58]. The stability of integrated genes within yeast genomes without selection pressures is high, but this can still lead to 1% transformant loss after each cell division cycle.
Table 2. Overview of the fermentation experiments and parameters used for the optimization of recombinant Gox, CDH, LAC, and HRP expression in P. pastoris, including the fermentation yield of the enzyme.

References

  1. Tikhonov, B.B.; Sulman, E.M.; Stadol’nikova, P.Y.; Sulman, A.M.; Golikova, E.P.; Sidorov, A.I.; Matveeva, V.G. Immobilized Enzymes from the Class of Oxidoreductases in Technological Processes: A Review. Catal. Ind. 2019, 11, 251–263.
  2. Robinson, P.K. Enzymes: Principles and Biotechnological Applications. Essays Biochem. 2015, 59, 1–41.
  3. Cárdenas-Moreno, Y.; González-Bacerio, J.; García Arellano, H.; del Monte-Martínez, A. Oxidoreductase Enzymes: Characteristics, Applications, and Challenges as a Biocatalyst. Biotechnol. Appl. Biochem. 2023, 70, 2108–2135.
  4. Martínez, A.T.; Ruiz-Dueñas, F.J.; Camarero, S.; Serrano, A.; Linde, D.; Lund, H.; Vind, J.; Tovborg, M.; Herold-Majumdar, O.M.; Hofrichter, M.; et al. Oxidoreductases on Their Way to Industrial Biotransformations. Biotechnol. Adv. 2017, 35, 815–831.
  5. Stanišić, M.D.; Popović Kokar, N.; Ristić, P.; Balaž, A.M.; Senćanski, M.; Ognjanović, M.; Đokić, V.R.; Prodanović, R.; Todorović, T.R. Chemical Modification of Glycoproteins’ Carbohydrate Moiety as a General Strategy for the Synthesis of Efficient Biocatalysts by Biomimetic Mineralization: The Case of Glucose Oxidase. Polymers 2021, 13, 3875.
  6. Bankar, S.B.; Bule, M.V.; Singhal, R.S.; Ananthanarayan, L. Glucose Oxidase—An Overview. Biotechnol. Adv. 2009, 27, 489–501.
  7. Wang, F.; Chen, X.; Wang, Y.; Li, X.; Wan, M.; Zhang, G.; Leng, F.; Zhang, H. Insights into the Structures, Inhibitors, and Improvement Strategies of Glucose Oxidase. Int. J. Mol. Sci. 2022, 23, 9841.
  8. Bauer, J.A.; Zámocká, M.; Majtán, J.; Bauerová-Hlinková, V. Glucose Oxidase, an Enzyme “Ferrari”: Its Structure, Function, Production and Properties in the Light of Various Industrial and Biotechnological Applications. Biomolecules 2022, 12, 472.
  9. Khatami, S.H.; Vakili, O.; Ahmadi, N.; Soltani Fard, E.; Mousavi, P.; Khalvati, B.; Maleksabet, A.; Savardashtaki, A.; Taheri-Anganeh, M.; Movahedpour, A. Glucose Oxidase: Applications, Sources, and Recombinant Production. Biotechnol. Appl. Biochem. 2022, 69, 939–950.
  10. Csarman, F.; Wohlschlager, L.; Ludwig, R. Cellobiose Dehydrogenase. In the Enzymes; Chaiyen, P., Tamanoi, F., Eds.; Academic Press: Cambridge, CA, USA, 2020; Volume 47, pp. 457–489.
  11. Scheiblbrandner, S.; Csarman, F.; Ludwig, R. Cellobiose Dehydrogenase in Biofuel Cells. Curr. Opin. Biotechnol. 2022, 73, 205–212.
  12. Kracher, D.; Ludwig, R. Cellobiose Dehydrogenase: An Essential Enzyme for Lignocellulose Degradation in Nature—A Review/Cellobiosedehydrogenase: Ein Essentielles Enzym Für Den Lignozelluloseabbau in Der Natur–Eine Übersicht. Die Bodenkult. J. Land. Manag. Food Environ. 2016, 67, 145–163.
  13. Balaz, A.; Blazic, M.; Popovic, N.; Prodanovic, O.; Ostafe, R.; Fischer, R.; Prodanovic, R. Expression, Purification and Characterization of Cellobiose Dehydrogenase Mutants from Phanerochaete chrysosporium in Pichia pastoris KM71H Strain. J. Serbian Chem. Soc. 2020, 85, 25–35.
  14. Sulej, J.; Jaszek, M.; Osińska-Jaroszuk, M.; Matuszewska, A.; Bancerz, R.; Janczarek, M. Natural Microbial Polysaccharides as Effective Factors for Modification of the Catalytic Properties of Fungal Cellobiose Dehydrogenase. Arch. Microbiol. 2021, 203, 4433–4448.
  15. Stanišić, M.D.; Popović Kokar, N.; Ristić, P.; Balaž, A.M.; Ognjanović, M.; Đokić, V.R.; Prodanović, R.; Todorović, T.R. The Influence of Isoenzyme Composition and Chemical Modification on Horseradish Peroxidase@ZIF-8 Biocomposite Performance. Polymers 2022, 14, 4834.
  16. Grigorenko, V.; Chubar, T.; Kapeliuch, Y.; Börchers, T.; Spener, F.; Egorova, A. New Approaches for Functional Expression of Recombinant Horseradish Peroxidase C in Escherichia coli. Biocatal. Biotransformation 1999, 17, 359–379.
  17. Spadiut, O.; Herwig, C. Production and Purification of the Multifunctional Enzyme Horseradish Peroxidase. Pharm. Bioprocess. 2013, 1, 283–295.
  18. Pantić, N.; Spasojević, M.; Stojanović, Ž.; Veljović, Đ.; Krstić, J.; Balaž, A.M.; Prodanović, R.; Prodanović, O. Immobilization of Horseradish Peroxidase on Macroporous Glycidyl-Based Copolymers with Different Surface Characteristics for the Removal of Phenol. J. Polym. Environ. 2022, 30, 3005–3020.
  19. Veitch, N.C. Horseradish Peroxidase: A Modern View of a Classic Enzyme. Phytochemistry 2004, 65, 249–259.
  20. Piscitelli, A.; Pezzella, C.; Giardina, P.; Faraco, V.; Sannia, G. Heterologous Laccase Production and Its Role in Industrial Applications. Bioeng. Bugs 2010, 1, 254–264.
  21. Bourbonnais, R.; Paice, M.G. Demethylation and Delignification of Kraft Pulp by Trametes versicolor Laccase in the Presence of 2,2?-Azinobis-(3-Ethylbenzthiazoline-6-Sulphonate). Appl. Microbiol. Biotechnol. 1992, 36, 823–827.
  22. Kawai, S.; Umezawa, T.; Higuchi, T. Degradation Mechanisms of Phenolic β-1 Lignin Substructure Model Compounds by Laccase of Coriolus versicolor. Arch. Biochem. Biophys. 1988, 262, 99–110.
  23. Eggert, C.; Temp, U.; Dean, J.F.D.; 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.
  24. Necochea, R.; Valderrama, B.; Dãaz-Sandoval, S.; Folch-Mallol, J.L.; Vázquez-Duhalt, R.; Iturriaga, G. Phylogenetic and Biochemical Characterisation of a Recombinant Laccase from Trametes versicolor. FEMS Microbiol. Lett. 2005, 244, 235–241.
  25. Janusz, G.; Pawlik, A.; Świderska-Burek, U.; Polak, J.; Sulej, J.; Jarosz-Wilkołazka, A.; Paszczyński, A. Laccase Properties, Physiological Functions, and Evolution. Int. J. Mol. Sci. 2020, 21, 966.
  26. Khatami, S.H.; Vakili, O.; Movahedpour, A.; Ghesmati, Z.; Ghasemi, H.; Taheri-Anganeh, M. Laccase: Various Types and Applications. Biotechnol. Appl. Biochem. 2022, 69, 2658–2672.
  27. Baghban, R.; Farajnia, S.; Rajabibazl, M.; Ghasemi, Y.; Mafi, A.; Hoseinpoor, R.; Rahbarnia, L.; Aria, M. Yeast Expression Systems: Overview and Recent Advances. Mol. Biotechnol. 2019, 61, 365–384.
  28. Karbalaei, M.; Rezaee, S.A.; Farsiani, H. Pichia Pastoris: A Highly Successful Expression System for Optimal Synthesis of Heterologous Proteins. J. Cell Physiol. 2020, 235, 5867–5881.
  29. Gündüz Ergün, B.; Hüccetoğulları, D.; Öztürk, S.; Çelik, E.; Çalık, P. Established and Upcoming Yeast Expression Systems. In Recombinant Protein Production in Yeast; Gasser, B., Mattanovich, D., Eds.; Humana Press: Totowa, NJ, USA, 2019; Volume 1923, pp. 1–74.
  30. Huang, C.-J.; Lowe, A.J.; Batt, C.A. Recombinant Immunotherapeutics: Current State and Perspectives Regarding the Feasibility and Market. Appl. Microbiol. Biotechnol. 2010, 87, 401–410.
  31. Hasunuma, T.; Ishii, J.; Kondo, A. Rational Design and Evolutional Fine Tuning of Saccharomyces cerevisiae for Biomass Breakdown. Curr. Opin. Chem. Biol. 2015, 29, 1–9.
  32. Partow, S.; Siewers, V.; Bjørn, S.; Nielsen, J.; Maury, J. Characterization of Different Promoters for Designing a New Expression Vector in Saccharomyces cerevisiae. Yeast 2010, 27, 955–964.
  33. Kulagina, N.; Besseau, S.; Godon, C.; Goldman, G.H.; Papon, N.; Courdavault, V. Yeasts as Biopharmaceutical Production Platforms. Front. Fungal Biol. 2021, 2, 733492.
  34. Kingsman, S.M.; Kingsman, A.J.; Dobson, M.J.; Mellor, J.; Roberts, N.A. Heterologous Gene Expression in Saccharomyces cerevisiae. Biotechnol. Genet. Eng. Rev. 1985, 3, 377–416.
  35. Karaoğlan, M.; Erden-Karaoğlan, F. Effect of Codon Optimization and Promoter Choice on Recombinant Endo-Polygalacturonase Production in Pichia pastoris. Enzym. Microb. Technol. 2020, 139, 109589.
  36. Park, E.-H.; Shin, Y.-M.; Lim, Y.-Y.; Kwon, T.-H.; Kim, D.-H.; Yang, M.-S. Expression of Glucose Oxidase by Using Recombinant Yeast. J. Biotechnol. 2000, 81, 35–44.
  37. Kapat, A.; Jung, J.-K.; Park, Y.-H. Enhancement of Extracellular Glucose Oxidase Production in PH-Stat Feed-Back Controlled Fed-Batch Culture of Recombinant Saccharomyces cerevisiae. Biotechnol. Lett. 1998, 20, 683–686.
  38. De Baetselier, A. A New Production Method for Glucose Oxidase. J. Biotechnol. 1992, 24, 141–148.
  39. Sygmund, C.; Santner, P.; Krondorfer, I.; Peterbauer, C.K.; Alcalde, M.; Nyanhongo, G.S.; Guebitz, G.M.; Ludwig, R. Semi-Rational Engineering of Cellobiose Dehydrogenase for Improved Hydrogen Peroxide Production. Microb. Cell Fact. 2013, 12, 38.
  40. Blažić, M.; Balaž, A.M.; Tadić, V.; Draganić, B.; Ostafe, R.; Fischer, R.; Prodanović, R. Protein Engineering of Cellobiose Dehydrogenase from Phanerochaete chrysosporium in Yeast Saccharomyces cerevisiae InvSc1 for Increased Activity and Stability. Biochem. Eng. J. 2019, 146, 179–185.
  41. Banerjee, S.; Roy, A. Molecular Cloning, Characterisation and Expression of a Gene Encoding Cellobiose Dehydrogenase from Termitomyces clypeatus. Gene Rep. 2021, 23, 101063.
  42. Zhao, X.; Yu, H.; Liang, Q.; Zhou, J.; Li, J.; Du, G.; Chen, J. Stepwise Optimization of Inducible Expression System for the Functional Secretion of Horseradish Peroxidase in Saccharomyces cerevisiae. J. Agric. Food Chem. 2023, 71, 4059–4068.
  43. Morawski, B.; Lin, Z.; Cirino, P.; Joo, H.; Bandara, G.; Arnold, F.H. Functional Expression of Horseradish Peroxidase in Saccharomyces cerevisiae and Pichia pastoris. Protein Eng. Des. Sel. 2000, 13, 377–384.
  44. Bulter, T.; Alcalde, M.; Sieber, V.; Meinhold, P.; Schlachtbauer, C.; Arnold, F.H. Functional Expression of a Fungal Laccase in Saccharomyces cerevisiae by Directed Evolution. Appl. Environ. Microbiol. 2003, 69, 987–995.
  45. Iimura, Y.; Sonoki, T.; Habe, H. Heterologous Expression of Trametes versicolor Laccase in Saccharomyces cerevisiae. Protein Expr. Purif. 2018, 141, 39–43.
  46. Kurose, T.; Saito, Y.; Kimata, K.; Nakagawa, Y.; Yano, A.; Ito, K.; Kawarasaki, Y. Secretory Expression of Lentinula Edodes Intracellular Laccase by Yeast High-Cell-Density System: Sub-Milligram Production of Difficult-to-Express Secretory Protein. J. Biosci. Bioeng. 2014, 117, 659–663.
  47. 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.
  48. Klonowska, A.; Gaudin, C.; Asso, M.; Fournel, A.; Réglier, M.; Tron, T. LAC3, a New Low Redox Potential Laccase from Trametes sp. Strain C30 Obtained as a Recombinant Protein in Yeast. Enzym. Microb. Technol. 2005, 36, 34–41.
  49. Antošová, Z.; Herkommerová, K.; Pichová, I.; Sychrová, H. Efficient Secretion of Three Fungal Laccases from Saccharomyces cerevisiae and Their Potential for Decolorization of Textile Industry Effluent—A Comparative Study. Biotechnol. Prog. 2018, 34, 69–80.
  50. Kojima, Y.; Tsukuda, Y.; Kawai, Y.; Tsukamoto, A.; Sugiura, J.; Sakaino, M.; Kita, Y. Cloning, Sequence Analysis, and Expression of Ligninolytic Phenoloxidase Genes of the White-Rot Basidiomycete Coriolus hirsutus. J. Biol. Chem. 1990, 265, 15224–15230.
  51. Cassland, P.; Jönsson, L.J. Characterization of a Gene Encoding Trametes Versicolor Laccase A and Improved Heterologous Expression in Saccharomyces cerevisiae by Decreased Cultivation Temperature. Appl. Microbiol. Biotechnol. 1999, 52, 393–400.
  52. Aza, P.; Molpeceres, G.; de Salas, F.; Camarero, S. Design of an Improved Universal Signal Peptide Based on the α-Factor Mating Secretion Signal for Enzyme Production in Yeast. Cell. Mol. Life Sci. 2021, 78, 3691–3707.
  53. Ogata, K.; Nishikawa, H.; Ohsugi, M. A Yeast Capable of Utilizing Methanol. Agric. Biol. Chem. 1969, 33, 1519–1520.
  54. Vijayakumar, V.E.; Venkataraman, K. A Systematic Review of the Potential of Pichia pastoris (Komagataella phaffii) as an Alternative Host for Biologics Production. Mol. Biotechnol. 2023, 1–19.
  55. Çalık, P.; Ata, Ö.; Güneş, H.; Massahi, A.; Boy, E.; Keskin, A.; Öztürk, S.; Zerze, G.H.; Özdamar, T.H. Recombinant Protein Production in Pichia pastoris under Glyceraldehyde-3-Phosphate Dehydrogenase Promoter: From Carbon Source Metabolism to Bioreactor Operation Parameters. Biochem. Eng. J. 2015, 95, 20–36.
  56. Walsh, G.; Walsh, E. Biopharmaceutical Benchmarks 2022. Nat. Biotechnol. 2022, 40, 1722–1760.
  57. Ahmad, M.; Hirz, M.; Pichler, H.; Schwab, H. Protein Expression in Pichia pastoris: Recent Achievements and Perspectives for Heterologous Protein Production. Appl. Microbiol. Biotechnol. 2014, 98, 5301–5317.
  58. Juturu, V.; Wu, J.C. Heterologous Protein Expression in Pichia pastoris: Latest Research Progress and Applications. ChemBioChem 2018, 19, 7–21.
  59. Qiu, Z.; Guo, Y.; Bao, X.; Hao, J.; Sun, G.; Peng, B.; Bi, W. Expression of Aspergillus niger Glucose Oxidase in Yeast Pichia pastoris SMD1168. Biotechnol. Biotechnol. Equip. 2016, 30, 998–1005.
  60. Martínez-Mora, E.; González-González, M.D.R.; Zarate, X.; Carranza-Rosales, P.; Ramírez-Cabrera, M.A.; Balderas-Rentería, I.; Arredondo-Espinoza, E. Enhanced in Vitro Anticancer Activity of Yeast Expressed Recombinant Glucose Oxidase versus Commercial Enzyme. Appl. Microbiol. Biotechnol. 2021, 105, 2377–2384.
  61. Belyad, F.; Karkhanei, A.A.; Raheb, J. Expression, Characterization and One Step Purification of Heterologous Glucose Oxidase Gene from Aspergillus niger ATCC 9029 in Pichia pastoris. EuPA Open Proteom. 2018, 19, 1–5.
  62. Kovačević, G.; Blažić, M.; Draganić, B.; Ostafe, R.; Gavrović-Jankulović, M.; Fischer, R.; Prodanović, R. Cloning, Heterologous Expression, Purification and Characterization of M12 Mutant of Aspergillus niger Glucose Oxidase in Yeast Pichia pastoris KM71H. Mol. Biotechnol. 2014, 56, 305–311.
  63. Bey, M.; Berrin, J.-G.; Poidevin, L.; Sigoillot, J.-C. Heterologous Expression of Pycnoporus cinnabarinus Cellobiose Dehydrogenase in Pichia pastoris and Involvement in Saccharification Processes. Microb. Cell Fact. 2011, 10, 113.
  64. Zhang, R.; Fan, Z.; Kasuga, T. Expression of Cellobiose Dehydrogenase from Neurospora crassa in Pichia pastoris and Its Purification and Characterization. Protein Expr. Purif. 2011, 75, 63–69.
  65. Acar, M.; Abul, N.; Yildiz, S.; Taskesenligil, E.D.; Gerni, S.; Unver, Y.; Kalin, R.; Ozdemir, H. Affinity-Based and in a Single Step Purification of Recombinant Horseradish Peroxidase A2A Isoenzyme Produced by Pichia pastoris. Bioprocess. Biosyst. Eng. 2023, 46, 523–534.
  66. Krainer, F.W.; Darnhofer, B.; Birner-Gruenberger, R.; Glieder, A. Recombinant Production of a Peroxidase-Protein G Fusion Protein in Pichia pastoris. J. Biotechnol. 2016, 219, 24–27.
  67. Kontro, J.; Lyra, C.; Koponen, M.; Kuuskeri, J.; Kähkönen, M.A.; Wallenius, J.; Wan, X.; Sipilä, J.; Mäkelä, M.R.; Nousiainen, P.; et al. Production of Recombinant Laccase From Coprinopsis cinerea and Its Effect in Mediator Promoted Lignin Oxidation at Neutral PH. Front. Bioeng. Biotechnol. 2021, 9, 767139.
  68. Ardila-Leal, L.D.; Albarracín-Pardo, D.A.; Rivera-Hoyos, C.M.; Morales-Álvarez, E.D.; Poutou-Piñales, R.A.; Cardozo-Bernal, A.M.; Quevedo-Hidalgo, B.E.; Pedroza-Rodríguez, A.M.; Díaz-Rincón, D.J.; Rodríguez-López, A.; et al. Media Improvement for 10 L Bioreactor Production of RPOXA 1B Laccase by P. pastoris. 3 Biotech. 2019, 9, 447.
  69. Hong, F.; Meinander, N.Q.; Jönsson, L.J. Fermentation Strategies for Improved Heterologous Expression of Laccase in Pichia pastoris. Biotechnol. Bioeng. 2002, 79, 438–449.
  70. O’Callaghan, J.; O’Brien, M.; McClean, K.; Dobson, A. Optimisation of the Expression of a Trametes versicolor Laccase Gene in Pichia pastoris. J. Ind. Microbiol. Biotechnol. 2002, 29, 55–59.
  71. Li, Q.; Pei, J.; Zhao, L.; Xie, J.; Cao, F.; Wang, G. Overexpression and Characterization of Laccase from Trametes versicolor in Pichia pastoris. Appl. Biochem. Microbiol. 2014, 50, 140–147.
  72. Avelar, M.; Olvera, C.; Aceves-Zamudio, D.; Folch, J.L.; Ayala, M. Recombinant Expression of a Laccase from Coriolopsis gallica in Pichia pastoris Using a Modified α-Factor Preproleader. Protein Expr. Purif. 2017, 136, 14–19.
  73. Fan, F.; Zhuo, R.; Sun, S.; Wan, X.; Jiang, M.; Zhang, X.; Yang, Y. Cloning and Functional Analysis of a New Laccase Gene from Trametes sp. 48424 Which Had the High Yield of Laccase and Strong Ability for Decolorizing Different Dyes. Bioresour. Technol. 2011, 102, 3126–3137.
  74. Xu, G.; Wang, J.; Yin, Q.; Fang, W.; Xiao, Y.; Fang, Z. Expression of a Thermo- and Alkali-Philic Fungal Laccase in Pichia pastoris and Its Application. Protein Expr. Purif. 2019, 154, 16–24.
  75. Yang, Z.; Zhang, Z. Engineering Strategies for Enhanced Production of Protein and Bio-Products in Pichia pastoris: A Review. Biotechnol. Adv. 2018, 36, 182–195.
  76. Gao, Z.; Li, Z.; Zhang, Y.; Huang, H.; Li, M.; Zhou, L.; Tang, Y.; Yao, B.; Zhang, W. High-Level Expression of the Penicillium notatum Glucose Oxidase Gene in Pichia pastoris Using Codon Optimization. Biotechnol. Lett. 2012, 34, 507–514.
  77. Yu, S.; Miao, L.; Huang, H.; Li, Y.; Zhu, T. High-Level Production of Glucose Oxidase in Pichia pastoris: Effects of Hac1p Overexpression on Cell Physiology and Enzyme Expression. Enzym. Microb. Technol. 2020, 141, 109671.
  78. Dietzsch, C.; Spadiut, O.; Herwig, C. A Dynamic Method Based on the Specific Substrate Uptake Rate to Set up a Feeding Strategy for Pichia pastoris. Microb. Cell Fact. 2011, 10, 14.
  79. Dietzsch, C.; Spadiut, O.; Herwig, C. A Fast Approach to Determine a Fed Batch Feeding Profile for Recombinant Pichia pastoris Strains. Microb. Cell Fact. 2011, 10, 85.
  80. Antošová, Z.; Sychrová, H. Yeast Hosts for the Production of Recombinant Laccases: A Review. Mol. Biotechnol. 2016, 58, 93–116.
  81. Jönsson, L.J.; Saloheimo, M.; Penttilä, M. Laccase from the White-Rot Fungus Trametes versicolor: CDNA Cloning of Lcc1 and Expression in Pichia pastoris. Curr. Genet. 1997, 32, 425–430.
  82. Markova, E.A.; Shaw, R.E.; Reynolds, C.R. Prediction of Strain Engineerings That Amplify Recombinant Protein Secretion through the Machine Learning Approach MaLPHAS. Eng. Biol. 2022, 6, 82–90.
  83. Waltenspühl, Y.; Jeliazkov, J.R.; Kummer, L.; Plückthun, A. Directed Evolution for High Functional Production and Stability of a Challenging G Protein-Coupled Receptor. Sci. Rep. 2021, 11, 8630.
  84. Wang, Y.; Yu, L.; Shao, J.; Zhu, Z.; Zhang, L. Structure-Driven Protein Engineering for Production of Valuable Natural Products. Trends Plant Sci. 2023, 28, 460–470.
  85. Iizuka, R.; Tahara, K.; Matsueda, A.; Tsuda, S.; Yoon, D.H.; Sekiguchi, T.; Shoji, S.; Funatsu, T. Selection of Green Fluorescent Proteins by in vitro Compartmentalization Using Microbead-Display Libraries. Biochem. Eng. J. 2022, 187, 108627.
  86. Hillson, N.; Caddick, M.; Cai, Y.; Carrasco, J.A.; Chang, M.W.; Curach, N.C.; Bell, D.J.; Le Feuvre, R.; Friedman, D.C.; Fu, X.; et al. Building a Global Alliance of Biofoundries. Nat. Commun. 2019, 10, 2040.
  87. Zrimec, J.; Börlin, C.S.; Buric, F.; Muhammad, A.S.; Chen, R.; Siewers, V.; Verendel, V.; Nielsen, J.; Töpel, M.; Zelezniak, A. Deep Learning Suggests That Gene Expression Is Encoded in All Parts of a Co-Evolving Interacting Gene Regulatory Structure. Nat. Commun. 2020, 11, 6141.
  88. Avsec, Ž.; Agarwal, V.; Visentin, D.; Ledsam, J.R.; Grabska-Barwinska, A.; Taylor, K.R.; Assael, Y.; Jumper, J.; Kohli, P.; Kelley, D.R. Effective Gene Expression Prediction from Sequence by Integrating Long-Range Interactions. Nat. Methods 2021, 18, 1196–1203.
  89. Ji, Y.; Zhou, Z.; Liu, H.; Davuluri, R. V DNABERT: Pre-Trained Bidirectional Encoder Representations from Transformers Model for DNA-Language in Genome. Bioinformatics 2021, 37, 2112–2120.
  90. Lu, H.; Li, F.; Sánchez, B.J.; Zhu, Z.; Li, G.; Domenzain, I.; Marcišauskas, S.; Anton, P.M.; Lappa, D.; Lieven, C.; et al. A Consensus S. Cerevisiae Metabolic Model Yeast8 and Its Ecosystem for Comprehensively Probing Cellular Metabolism. Nat. Commun. 2019, 10, 3586.
  91. Ito, Y.; Ishigami, M.; Terai, G.; Nakamura, Y.; Hashiba, N.; Nishi, T.; Nakazawa, H.; Hasunuma, T.; Asai, K.; Umetsu, M.; et al. A Streamlined Strain Engineering Workflow with Genome-Wide Screening Detects Enhanced Protein Secretion in Komagataella phaffii. Commun. Biol. 2022, 5, 561.
  92. Munro, L.J.; Kell, D.B. Intelligent Host Engineering for Metabolic Flux Optimisation in Biotechnology. Biochem. J. 2021, 478, 3685–3721.
  93. Sandberg, T.E.; Salazar, M.J.; Weng, L.L.; Palsson, B.O.; Feist, A.M. The Emergence of Adaptive Laboratory Evolution as an Efficient Tool for Biological Discovery and Industrial Biotechnology. Metab. Eng. 2019, 56, 1–16.
  94. Aza, P.; De Salas, F.; Molpeceres, G.; Rodríguez-Escribano, D.; De La Fuente, I.; Camarero, S. Protein Engineering Approaches to Enhance Fungal Laccase Production in S. cerevisiae. Int. J. Mol. Sci. 2021, 22, 1157.
  95. Zhou, H.; Zhang, W.; Qian, J. Hypersecretory Production of Glucose Oxidase in Pichia pastoris through Combinatorial Engineering of Protein Properties, Synthesis, and Secretion. Biotechnol. Bioeng. 2023, 121, 735–748.
  96. Shen, Q.; Wu, M.; Wang, H.-B.; Naranmandura, H.; Chen, S.-Q. The Effect of Gene Copy Number and Co-Expression of Chaperone on Production of Albumin Fusion Proteins in Pichia pastoris. Appl. Microbiol. Biotechnol. 2012, 96, 763–772.
  97. Duan, G.; Ding, L.; Wei, D.; Zhou, H.; Chu, J.; Zhang, S.; Qian, J. Screening Endogenous Signal Peptides and Protein Folding Factors to Promote the Secretory Expression of Heterologous Proteins in Pichia pastoris. J. Biotechnol. 2019, 306, 193–202.
  98. Ito, Y.; Terai, G.; Ishigami, M.; Hashiba, N.; Nakamura, Y.; Bamba, T.; Kumokita, R.; Hasunuma, T.; Asai, K.; Ishii, J.; et al. Exchange of Endogenous and Heterogeneous Yeast Terminators in Pichia pastoris to Tune MRNA Stability and Gene Expression. Nucleic Acids Res. 2020, 48, 13000–13012.
  99. Wang, G.; Björk, S.M.; Huang, M.; Liu, Q.; Campbell, K.; Nielsen, J.; Joensson, H.N.; Petranovic, D. RNAi Expression Tuning, Microfluidic Screening, and Genome Recombineering for Improved Protein Production in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 2019, 116, 9324–9332.
  100. Pekarsky, A.; Veiter, L.; Rajamanickam, V.; Herwig, C.; Grünwald-Gruber, C.; Altmann, F.; Spadiut, O. Production of a Recombinant Peroxidase in Different Glyco-Engineered Pichia pastoris Strains: A Morphological and Physiological Comparison. Microb. Cell Fact. 2018, 17, 183.
  101. Hyka, P.; Züllig, T.; Ruth, C.; Looser, V.; Meier, C.; Klein, J.; Melzoch, K.; Meyer, H.-P.; Glieder, A.; Kovar, K. Combined Use of Fluorescent Dyes and Flow Cytometry To Quantify the Physiological State of Pichia pastoris during the Production of Heterologous Proteins in High-Cell-Density Fed-Batch Cultures. Appl. Environ. Microbiol. 2010, 76, 4486–4496.
  102. Alfasi, S.; Sevastsyanovich, Y.; Zaffaroni, L.; Griffiths, L.; Hall, R.; Cole, J. Use of GFP Fusions for the Isolation of Escherichia coli Strains for Improved Production of Different Target Recombinant Proteins. J. Biotechnol. 2011, 156, 11–21.
  103. Totaro, D.; Radoman, B.; Schmelzer, B.; Rothbauer, M.; Steiger, M.G.; Mayr, T.; Sauer, M.; Ertl, P.; Mattanovich, D. Microscale Perfusion-Based Cultivation for Pichia pastoris Clone Screening Enables Accelerated and Optimized Recombinant Protein Production Processes. Biotechnol. J. 2021, 16, e2000215.
  104. Li, Q.; Lu, J.; Liu, J.; Li, J.; Zhang, G.; Du, G.; Chen, J. High-Throughput Droplet Microfluidics Screening and Genome Sequencing Analysis for Improved Amylase-Producing Aspergillus oryzae. Biotechnol. Biofuels Bioprod. 2023, 16, 185.
  105. Scheele, R.A.; Lindenburg, L.H.; Petek, M.; Schober, M.; Dalby, K.N.; Hollfelder, F. Droplet-Based Screening of Phosphate Transfer Catalysis Reveals How Epistasis Shapes MAP Kinase Interactions with Substrates. Nat. Commun. 2022, 13, 844.
  106. Femmer, C.; Bechtold, M.; Panke, S. Semi-rational Engineering of an Amino Acid Racemase That Is Stabilized in Aqueous/Organic Solvent Mixtures. Biotechnol. Bioeng. 2020, 117, 2683–2693.
  107. Napiorkowska, M.; Pestalozzi, L.; Panke, S.; Held, M.; Schmitt, S. High-Throughput Optimization of Recombinant Protein Production in Microfluidic Gel Beads. Small 2021, 17, e2005523.
  108. Ito, Y.; Sasaki, R.; Asari, S.; Yasuda, T.; Ueda, H.; Kitaguchi, T. Efficient Microfluidic Screening Method Using a Fluorescent Immunosensor for Recombinant Protein Secretions. Small 2023, 19, e2207943.
More