Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2967 2023-10-09 18:35:32 |
2 format change Meta information modification 2967 2023-10-10 08:09:47 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Gutiérrez-Corona, J.F.; González-Hernández, G.A.; Padilla-Guerrero, I.E.; Olmedo-Monfil, V.; Martínez-Rocha, A.L.; Patiño-Medina, J.A.; Meza-Carmen, V.; Torres-Guzmán, J.C. Physiological Role of Alcohol Dehydrogenases in Fungal Cells. Encyclopedia. Available online: https://encyclopedia.pub/entry/50003 (accessed on 05 September 2024).
Gutiérrez-Corona JF, González-Hernández GA, Padilla-Guerrero IE, Olmedo-Monfil V, Martínez-Rocha AL, Patiño-Medina JA, et al. Physiological Role of Alcohol Dehydrogenases in Fungal Cells. Encyclopedia. Available at: https://encyclopedia.pub/entry/50003. Accessed September 05, 2024.
Gutiérrez-Corona, J. Félix, Gloria Angélica González-Hernández, Israel Enrique Padilla-Guerrero, Vianey Olmedo-Monfil, Ana Lilia Martínez-Rocha, J. Alberto Patiño-Medina, Víctor Meza-Carmen, Juan Carlos Torres-Guzmán. "Physiological Role of Alcohol Dehydrogenases in Fungal Cells" Encyclopedia, https://encyclopedia.pub/entry/50003 (accessed September 05, 2024).
Gutiérrez-Corona, J.F., González-Hernández, G.A., Padilla-Guerrero, I.E., Olmedo-Monfil, V., Martínez-Rocha, A.L., Patiño-Medina, J.A., Meza-Carmen, V., & Torres-Guzmán, J.C. (2023, October 09). Physiological Role of Alcohol Dehydrogenases in Fungal Cells. In Encyclopedia. https://encyclopedia.pub/entry/50003
Gutiérrez-Corona, J. Félix, et al. "Physiological Role of Alcohol Dehydrogenases in Fungal Cells." Encyclopedia. Web. 09 October, 2023.
Physiological Role of Alcohol Dehydrogenases in Fungal Cells
Edit

Alcohol dehydrogenases (ADHs) (EC 1.1.1.1) are oxidoreductases that catalyze the interconversion of alcohols, aldehydes, and ketones. Oxidoreductases have been classified into three main categories: (1) NAD- or NADP-dependent dehydrogenases; (2) NAD(P)-independent enzymes that use pyrroloquinoline quinone, haem, or F420 as a cofactor; and (3) oxidases that catalyze essentially irreversible oxidation of alcohols.

alcohol dehydrogenase biological interactions fungi

1. Introduction

Alcohol dehydrogenases (ADHs) from category 1 have been divided into the following: group I long-chain (approximately 350 amino acid residues) zinc-dependent enzymes; group II short-chain (about 250 residues) zinc-independent enzymes; and group III “iron-activated” enzymes that generally contain approximately 385 amino acid residues [1][2]. ADHs’ functions are crucial to the fungal life cycle, such as growing under aerobic or anaerobic conditions, living inside the host, or establishing symbiotic interactions. With the increase in genome sequencing of fungal isolates, ADH systems in fungi belonging to different taxonomic groups have been revealed. However, progress in understanding the molecular structure and physiological role of the various ADH enzymes has been made in very few organisms. The most well-studied ADHs for their role in growth and energy metabolism are ADHs from category 1, group I (Zn-dependent), and group II (Zn-independent proteins). 

2. Role of Zn-Dependent ADHs in Growth and Energy Metabolism in Yeasts

The yeast Saccharomyces cerevisiae is the organism in which the most progress has been made in relation to the identification and study of the function of ADH proteins so that the proteins encoded by the ADH1 to ADH7 genes have been identified and biochemically and physiologically characterized. S. cerevisiae is a Crabtree-positive organism in which the availability of oxygen is irrelevant to fermentative metabolism. Glucose generates high levels of glycolytic enzymes and represses respiration, which leads to ethanol production [3]. In this yeast, both in the presence and absence of oxygen, the cytosolic enzyme Adh1p produces ethanol and NAD+ using glucose as a carbon source. Under aerobic conditions, after glucose depletion, the ethanol produced is oxidized to acetaldehyde via the cytosolic enzyme Adh2p; this last compound can be metabolized via the tricarboxylic acid cycle or as an intermediate metabolite in gluconeogenesis [4][5]. In S. cerevisiae, the Adh3p enzyme is localized in the mitochondria and was proposed to be involved in the transport of mitochondrial-reducing equivalents to the cytosol [6]. By using quadruple mutants in ADH genes, in which only a single functional ADH was left, it was shown that the enzyme Adh3p produces ethanol from acetaldehyde with a capacity no greater than that shown by Adh2p. In the absence of glucose, Adh3p enables growth and ethanol assimilation, which indicates that the product of the ADH3 gene replaces the Adh1p and Adh2p enzymes in its function, which occurs to a limited extent and is only noticeable when these enzymes are not active [5]. The Adh4p protein is not similar in amino acid sequence or structurally with the enzymes Adh1p, Adh2p, and Adh3p; it is activated by zinc and resembles Adh1p in its kinetic parameters [7]. Subsequent studies, using a quadruple mutant that only has Adh4p as a functional enzyme, showed that this enzyme does not allow the production of ethanol via the reduction of acetaldehyde during growth in glucose medium. A similar experimental approach showed that the enzyme Adh5p does not support growth in ethanol as the sole carbon source, indicating that this enzyme is not appreciably involved in the oxidation of ethanol to acetaldehyde [5]. The ADH6 gene from S. cerevisiae encodes a zinc-containing ADH. The enzyme showed specificity to NADPH and activity with aldehydes and primary, aliphatic (linear and branched chain), and aromatic alcohols. The interruption of the ADH6 gene makes S. cerevisiae sensitive to toxic concentrations of veratraldehyde, and the overexpression of the ADH6 gene enables the yeast to grow under these conditions [8]. The ADH7 gene of S. cerevisiae codes for an NADP(H)-dependent ADH, like the Adh6p enzyme. The purified Adh7p enzyme is a homodimer that exhibits broad substrate specificity similar to Adh6p. The deletion of the ADH7 gene does not affect growth, although with the mutant deleted in this gene, no experiments were carried out on the phenotype of sensitivity to compounds such as veratraldehyde and anisaldehyde, which are produced during ligninolysis and substrates for the enzymes Adh6p and Adh7p [8].
In the galactose-fermenting yeast Kluyveromyces lactis, four ADH genes have been identified that encode cytosolic (KlAdh1p and KlAdh2p) and mitochondrial isozymes (KlAdh3p and KlAdh4p) [9][10][11]. The physiological role of K. lactis ADH enzymes was investigated by generating triple-mutant strains in ADH genes, in which only a single functional ADH was left [12]. The analysis of these mutants showed that each ADH activity enables growth in ethanol as the sole carbon source; only the strain carrying all four mutant genes showed an inability to grow in ethanol. Regarding the accumulation of ethanol produced in a culture medium containing glucose, different amounts were observed, KlAdh1p being the highest producer, followed by KlAdh2p, and finally, KlAdh3p and KlAdh4p produced ethanol at a very low level [12].
The yeast Pichia stipitis (currently known as Scheffersomyces stipitis) is one of the few organisms capable of fermenting the pentose xylose, although it is also capable of using glucose; the yeast shows the regulation of the transition between the respiratory and fermentative processes [13]. S. stipitis is a Crabtree-negative organism, so oxygen limitation, rather than the presence of glucose or xylose, induces fermentation [14][15]. Using gene-disruption experiments, the physiological role of the ADH-gene-encoded products was studied. The loss of SsADH1 negatively affects growth and ethanol production in a xylose medium under oxygen limitation [16]. On the other hand, the loss of SsADH2 enzyme function did not cause the aforementioned effects. The double-mutant strain SsADH1 SsADH2 could not grow in xylose and produced very little ethanol. These observations suggested that the SsADH2 isoenzyme is also involved in xylose fermentation [16]. Like S. stipitis, Candida maltosa is capable of fermenting glucose or xylose. In this yeast, three ADH genes, CmADH1, CmADH2A, and CmADH2B, were cloned and sequenced [17]. In a mutant of C. maltose isolated as resistant to allyl alcohol, it was observed that the activity of the CmADH2A enzyme was affected and that this deficiency caused a dramatic decrease in ethanol production in a medium with xylose under oxygen limitation; however, the above-mentioned deficiency of CmADH2A enzyme activity had no effect on ethanol production from glucose under aerobic conditions [17]. According to the expression pattern of CmADH isozymes and fermentation studies of the CmADH2A-deficient mutant, it was found that CmADH1 and CmADH2A were both expressed and involved in ethanol formation during xylose metabolism [17]. Candida utilis has also been known as a yeast capable of metabolizing xylose; the enzyme ADH1 has been described in this organism. Inspection of the amino acid sequence of C. utilis ADH1 and enzyme characterization with different aliphatic and branched alcohols indicated that it could be a primary alcohol dehydrogenase that requires zinc ions for catalytic reactions; the specific ADH activity is nine times higher for NAD+ than that for NADP+, indicating that the enzyme preferred NAD+ to NADP+ as a cofactor [18].
The yeast Yarrowia lipolytica can utilize ethanol as a carbon source but is not capable of producing ethanol or growing under anaerobic conditions. Eight ADH genes were identified in this yeast, all of which were deleted [19][20]. The characterization of combinations of mutations in the ADH genes showed that the joint alteration of the ADH1, ADH2, and ADH3 genes results in the phenotype of an inability to grow in ethanol, indicating that the corresponding proteins participate in the oxidation of the alcohol and that it is enough that one of them is active for this process to occur. Additionally, it was observed that for growth in ethanol, in addition to the three ADH genes mentioned, the ACS1 gene is required, which encodes an acetyl-CoA synthetase [21]. On the other hand, another study found that alcohol dehydrogenase genes ADH1 and ADH3 and a fatty alcohol oxidase gene, FAO1, play an important role in the oxidation of exogenous fatty alcohols but play less prominent roles in the oxidation of fatty alcohols derived from n-alkanes [20].

3. Role of Zn-Dependent ADHs in Growth and Energy Metabolism in Filamentous Fungi

Neurospora crassa was one of the first filamentous fungi in which the ability to produce appreciable amounts of ethanol was demonstrated in cultures incubated in a glucose medium [22]. Through zymography, it was shown that the fungus produces two different ADHs. Fermentative ADH, which showed a single band of protein with activity, is produced when the fungus is cultured in sucrose, while oxidative ADH, which consists of at least two electromorphs, is produced in a medium with ethanol as a carbon source; the latter enzyme is repressed by glucose or by sucrose. The enzymes also differ in substrate specificity, in the ratio of the forward (glucose as substrate) and reverse reactions (ethanol as substrate), and in their thermostability [23]. It has been shown that N. crassa can produce ethanol in cultures with different hexoses and pentoses, as well as with cellulose polymers and with lignocellulosic agroindustrial residues, such as straw, wood, and various agricultural and wood processing waste products [22]. Genes encoding ADH1 and ADH3 enzymes have been identified and expressed in E. coli to carry out the biochemical characterization of the enzymes [24]. However, no alteration experiments have been carried out on these genes to demonstrate in vivo the function of the enzymes they encode.
In Aspergillus nidulans, three genes that code for ADH enzymes have been identified. The alcA gene codes for the enzyme ADHI, which is the enzyme required to utilize ethanol as the sole carbon source [25][26][27]. The enzyme alcohol dehydrogenase II (ADHII) has been detected, by its activity, in zymograms, although its physiological function is unknown. This enzyme is encoded by the alcB gene [28]. The gene alcC encoding the enzyme alcohol dehydrogenase III (ADHIII) was identified as a cDNA sequence capable of complementing the ADHI mutation in S. cerevisiae [29]. To date, there is no known physiological function of the enzyme.
In Aspergillus flavus Furukawa et al., [30] reported that ethanol and 2-propanol at low concentrations (<1% and <0.6%, respectively) increase aflatoxin production. Both alcohols are incorporated into the aflatoxin biosynthesis via acetyl-CoA. In this context, A. flavus utilizes ethanol and 2-propanol as carbon sources for aflatoxin biosynthesis, and the adh1 gene and probably other putative alcohol dehydrogenase controls this aflatoxin production, balancing ethanol production and catabolism.
Mucor circinelloides f. lusitanicus (currently known as Mucor lusitanicus) is a dimorphic fungus belonging to the subphylum Mucoromycotina. It has been described that M. lusitanicus is a Crabtree-positive organism since, when grown in a high-glucose medium under aerobic conditions, it produces appreciable amounts of ethanol [31][32]. In this organism, the adh1 gene was identified, which encodes a zinc-dependent enzyme that uses NAD+ as a cofactor [33]. A mutant of M. lusitanicus altered in the adh1 gene showed, in comparison with the wild-type strain R7B and with the mutant complemented with the wild-type adh+ allele, alteration to several physiological characteristics, including an inability to grow under anaerobic conditions in a glucose medium, production of low levels of ethanol in a medium with glucose in cultures carried out under aerobic conditions, a notable reduction in growth in a medium with ethanol as the only carbon source under aerobic conditions, and very low levels of alcohol dehydrogenase activity in the cytosolic fraction of the mycelium obtained under aerobic conditions in a glucose medium. These observations indicate that in M. lusitanicus, the product of the adh1 gene acts to mediate the Crabtree effect and can act as a fermentative or oxidative enzyme, depending on the nutritional conditions [34].
Shah et al. [32] suggests that the adh1 gene from M. circinelloides is also involved in lipid biosynthesis. In the M. lusitanicus strain WJ11, the knocking out of the adh1 gene reduced the ethanol production by 85–90%, and the lipid and fatty acid content was decreased. These lipid and fatty acid contents were restored when the fermentation media was supplemented with 0.5% external ethanol.
In the phytopathogenic fungus Fusarium oxysporum f. sp. lycopersici, mutants deficient in the ADH1 gene were isolated by obtaining spontaneous allyl alcohol-resistant mutants in a glucose medium. The characterization of one of these strains showed that the insertion into the ADH1 gene of an incomplete transposable element of F. oxysporum occurred. This mutation caused an inability to use ethanol as a carbon source under aerobic conditions, impaired growth under hypoxic conditions with glucose as a carbon source, and decreased ethanol production in the glucose medium. Complementation with the wild-type ADH1 allele restored all defects produced by the mutation, indicating that depending on the culture conditions, the ADH1 gene product has fermentative and oxidative enzymatic functions [35].
Botrytis cinerea is a necrotrophic fungus that infects many types of plant species. A Bcadh1 gene was identified in this organism, encoding a zinc-dependent enzyme. A mutant strain deleted in the referred gene was obtained and characterized [36]. It was observed that, compared to the wild-type strain, the Δbcadh1 mutant showed several phenotypic alterations, including a reduced growth capacity in low oxygen tension, greater tolerance to ethanol, and sensitivity to reactive oxygen species. The increased ethanol tolerance of the Δbcadh1 mutant indicated that the enzyme is involved in the fermentation of glucose to ethanol. However, the phenotypic characteristic that best explains this physiological role of the enzyme is the lower growth capacity in low oxygen tension [36].
Table 1 summarizes the substrates, cofactors, oligomeric state, cellular location, and function of fungal Zn-dependent ADH enzymes.
Table 1. Features of Zn-dependent fungal ADHs.

4. Role of Non-Zn-Dependent ADHs in Growth and Energy Metabolism

The white rot basidiomycete fungus Phanerochaete chrysosporium can degrade lignin to carbon dioxide, which requires enzymes that carry out the reduction of aryl-aldehydes to the corresponding alcohols. In this organism, the PcAAD1 gene encoding a 386-amino-acid aryl alcohol dehydrogenase (AAD) was identified. The PcAAD1 gene is expressed in a higher proportion under conditions of nitrogen limitation [43][45]. Analysis of the protein PcAad1p indicates that it belongs to the AKR9A subfamily of the AKR (aldo-keto reductase) superfamily, in which there is a conserved NADPH-binding domain. The purified enzyme is active in the reduction of more than 20 different aliphatic and aryl aldehydes, although it shows the highest activity with veratryl alcohol, vanillyl alcohol, and the corresponding aldehydes [43]. A similar AAD enzyme is overexpressed in the white rot basidiomycete Coriolus versicolor exposed to either 4-methyldibenzothiophene-5-oxide (4MDBTO) or dibenzothiophene-5-oxide (DBTO) [46]. It is proposed that in these organisms, aryl alcohol dehydrogenase is part of the lignin degradation machinery and the response to chemical stress [43][46].
Anaerobic chytridiomycete fungi, such as Piromyces and Neocallimastix, are important symbionts in the gastrointestinal tracts of herbivorous mammals, where they contribute to the degradation of plant polymers that form an important part of the diets of ruminants [47]. It was shown that Piromyces and Neocallimastix have an ADHE-type gene in their genome and that the ADHE protein was located in the cytosol [47]. This enzyme is homologous to the ADHE protein of eubacteria and that from parabasal flagellate anaerobes Trichomonas vaginalis and Tritrichomonas fetus [48]. Piromyces sp. E2 conducts the final steps of its carbohydrate catabolism via PFL and ADHE enzymes, producing metabolites of a bacterial-type mixed-acid fermentation. Similar to bacterial ADHE, the Piromyces ADHE enzyme is anticipated to be iron-dependent for catalysis and uses NAD as a cofactor [49].

References

  1. Reid, M.F.; Fewson, C.A. Molecular characterization of microbial alcohol dehydrogenases. Crit. Rev. Microbiol. 1994, 20, 13–56.
  2. Gaona-López, C.; Julián-Sánchez, A.; Riveros-Rosas, H. Diversity and Evolutionary Analysis of Iron-Containing (Type-III) Alcohol Dehydrogenases in Eukaryotes. PLoS ONE 2016, 11, e0166851.
  3. Bennetzen, J.L.; Hall, B.D. The primary structure of the Saccharomyces cerevisiae gene for alcohol dehydrogenase. J. Biol. Chem. 1982, 257, 3018–3025.
  4. De Smidt, O.; du Preez, J.C.; Albertyn, J. The alcohol dehydrogenases of Saccharomyces cerevisiae: A comprehensive review. FEMS Yeast Res. 2008, 8, 967–978.
  5. De Smidt, O.; du Preez, J.C.; Albertyn, J. Molecular and physiological aspects of alcohol dehydrogenases in the ethanol metabolism of Saccharomyces cerevisiae. FEMS Yeast Res. 2012, 12, 33–47.
  6. Bakker, B.M.; Bro, C.; Kötter, P.; Luttik, M.A.; van Dijken, J.P.; Pronk, J.T. The mitochondrial alcohol dehydrogenase Adh3p is involved in a redox shuttle in Saccharomyces cerevisiae. J. Bacteriol. 2000, 182, 4730–4737.
  7. Drewke, C.; Ciriacy, M. Overexpression, purification and properties of alcohol dehydrogenase IV from Saccharomyces cerevisiae. Biochim. Biophys. Acta 1988, 950, 54–60.
  8. Larroy, C.; Parés, X.; Biosca, J.A. Characterization of a Saccharomyces cerevisiae NADP(H)-dependent alcohol dehydrogenase (ADHVII), a member of the cinnamyl alcohol dehydrogenase family. Eur. J. Biochem. 2002, 269, 5738–5745.
  9. Saliola, M.; Shuster, J.R.; Falcone, C. The alcohol dehydrogenase system in the yeast, Kluyveromyces lactis. Yeast 1990, 6, 193–204.
  10. Saliola, M.; Gonnella, R.; Mazzoni, C.; Falcone, C. Two genes encoding putative mitochondrial alcohol dehydrogenases are present in the yeast Kluyveromyces lactis. Yeast 1991, 7, 391–400.
  11. Shain, D.H.; Salvadore, C.; Denis, C.L. Evolution of the alcohol dehydrogenase (ADH) genes in yeast: Characterization of a fourth ADH in Kluyveromyces lactis. Mol. Gen. Genet. 1992, 232, 479–488.
  12. Saliola, M.; Bellardi, S.; Marta, I.; Falcone, C. Glucose metabolism and ethanol production in adh multiple and null mutants of Kluyveromyces lactis. Yeast 1994, 10, 1133–1140.
  13. Jeffries, T.W. Utilization of xylose by bacteria, yeasts, and fungi. Adv. Biochem. Eng. Biotechnol. 1983, 27, 1–32.
  14. Passoth, V.; Zimmermann, M.; Klinner, U. Peculiarities of the regulation of fermentation and respiration in the crabtree-negative, xylose-fermenting yeast Pichia stipitis. Appl. Biochem. Biotechnol. 1996, 57–58, 201–212.
  15. Van Urk, H.; Voll, W.S.; Scheffers, W.A.; Van Dijken, J.P. Transient-state analysis of metabolic fluxes in crabtree-positive and crabtree-negative yeasts. Appl. Environ. Microbiol. 1990, 56, 281–287.
  16. Cho, J.Y.; Jeffries, T.W. Pichia stipitis genes for alcohol dehydrogenase with fermentative and respiratory functions. Appl. Environ. Microbiol. 1998, 64, 1350–1358.
  17. Lin, Y.; He, P.; Wang, Q.; Lu, D.; Li, Z.; Wu, C.; Jiang, N. The alcohol dehydrogenase system in the xylose-fermenting yeast Candida maltosa. PLoS ONE 2010, 5, e11752.
  18. Park, Y.C.; Yun, N.R.; San, K.Y.; Bennett, G.N. Molecular cloning and characterization of the alcohol dehydrogenase ADH1 gene of Candida utilis ATCC 9950. J. Ind. Microbiol. Biotechnol. 2006, 33, 1032–1036.
  19. Gatter, M.; Förster, A.; Bär, K.; Winter, M.; Otto, C.; Petzsch, P.; Ježková, M.; Bahr, K.; Pfeiffer, M.; Matthäus, F.; et al. A newly identified fatty alcohol oxidase gene is mainly responsible for the oxidation of long-chain ω-hydroxy fatty acids in Yarrowia lipolytica. FEMS Yeast Res. 2014, 14, 858–872.
  20. Iwama, R.; Kobayashi, S.; Ohta, A.; Horiuchi, H.; Fukuda, R. Alcohol dehydrogenases and an alcohol oxidase involved in the assimilation of exogenous fatty alcohols in Yarrowia lipolytica. FEMS Yeast Res. 2015, 15, fov014.
  21. Gatter, M.; Ottlik, S.; Kövesi, Z.; Bauer, B.; Matthäus, F.; Barth, G. Three alcohol dehydrogenase genes and one acetyl-CoA synthetase gene are responsible for ethanol utilization in Yarrowia lipolytica. Fungal Genet. Biol. 2016, 95, 30–38.
  22. Dogaris, I.; Mamma, D.; Kekos, D. Biotechnological production of ethanol from renewable resources by Neurospora crassa: An alternative to conventional yeast fermentations? Appl. Microbiol. Biotechnol. 2013, 97, 1457–1473.
  23. Zink, M.W. Regulation of NAD-specific alcohol dehydrogenases in Neurospora crassa. Can. J. Microbiol. 1969, 15, 265–271.
  24. Park, Y.C.; San, K.Y.; Bennett, G.N. Characterization of alcohol dehydrogenase 1 and 3 from Neurospora crassa FGSC2489. Appl. Microbiol. Biotechnol. 2007, 76, 349–356.
  25. Pateman, J.A.; Doy, C.H.; Olsen, J.E.; Norris, U.; Creaser, E.H.; Hynes, M. Regulation of alcohol dehydrogenase (ADH) and aldehyde dehydrogenase (AldDH) in Aspergillus nidulans. Proc. R Soc. Lond. B Biol. Sci. 1983, 217, 243–264.
  26. Lockington, R.A.; Sealy-Lewis, H.M.; Scazzocchio, C.; Davies, R.W. Cloning and characterization of the ethanol utilization regulon in Aspergillus nidulans. Gene 1985, 33, 137–149.
  27. Doy, C.H.; Pateman, J.A.; Olsen, J.E.; Kane, H.J.; Creaser, E.H. Genomic clones of Aspergillus nidulans containing alcA, the structural gene for alcohol dehydrogenase and alcR, a regulatory gene for ethanol metabolism. DNA 1985, 4, 105–114.
  28. Sealy-Lewis, H.M.; Lockington, R.A. Regulation of two alcohol dehydrogenases in Aspergillus nidulans. Curr. Genet. 1984, 8, 253–259.
  29. McKnight, G.L.; Kato, H.; Upshall, A.; Parker, M.D.; Saari, G.; O’Hara, P.J. Identification and molecular analysis of a third Aspergillus nidulans alcohol dehydrogenase gene. EMBO J. 1985, 4, 2093–2099.
  30. Furukawa, T.; Kushiro, M.; Nakagawa, H.; Enomoto, H.; Sakuda, S. Low-dose ethanol increases aflatoxin production due to the adh1-dependent incorporation of ethanol into aflatoxin biosynthesis. iScience 2023, 26, 106051.
  31. McIntyre, M.; Breum, J.; Arnau, J.; Nielsen, J. Growth physiology and dimorphism of Mucor circinelloides (syn. racemosus) during submerged batch cultivation. Appl. Microbiol. Biotechnol. 2002, 58, 495–502.
  32. Shah, A.M.; Mohamed, H.; Fazili, A.B.A.; Yang, W.; Song, Y. Investigating the Effect of Alcohol Dehydrogenase Gene Knockout on Lipid Accumulation in Mucor circinelloides WJ11. J. Fungi 2022, 8, 917.
  33. Rangel-Porras, R.A.; Meza-Carmen, V.; Martinez-Cadena, G.; Torres-Guzmán, J.C.; González-Hernández, G.A.; Arnau, J.; Gutiérrez-Corona, J.F. Molecular analysis of an NAD-dependent alcohol dehydrogenase from the zygomycete Mucor circinelloides. Mol. Genet. Genom. 2005, 274, 354–363.
  34. Rangel-Porras, R.A.; Díaz-Pérez, S.P.; Mendoza-Hernández, J.M.; Romo-Rodríguez, P.; Alejandre-Castañeda, V.; Valle-Maldonado, M.I.; Torres-Guzmán, J.C.; González-Hernández, G.A.; Campos-Garcia, J.; Arnau, J.; et al. Alcohol dehydrogenase 1 participates in the Crabtree effect and connects fermentative and oxidative metabolism in the Zygomycete Mucor circinelloides. J. Microbiol. 2019, 57, 606–617.
  35. Corrales Escobosa, A.R.; Rangel Porras, R.A.; Meza Carmen, V.; Gonzalez Hernandez, G.A.; Torres Guzman, J.C.; Wrobel, K.; Wrobel, K.; Roncero, M.I.; Gutierrez Corona, J.F. Fusarium oxysporum Adh1 has dual fermentative and oxidative functions and is involved in fungal virulence in tomato plants. Fungal Genet. Biol. 2011, 48, 886–895.
  36. DafaAlla, T.; Abdalla, M.; El-Arabey, A.A.; Eltayb, W.A.; Mohapatra, R.K. Botrytis cinerea alcohol dehydrogenase mediates fungal development, environmental adaptation and pathogenicity. J. Biomol. Struct. Dyn. 2022, 40, 12426–12438.
  37. Creaser, E.H.; Porter, R.L.; Britt, K.A.; Pateman, J.A.; Doy, C.H. Purification and preliminary characterization of alcohol dehydrogenase from Aspergillus nidulans. Biochem. J. 1985, 225, 449–454.
  38. Jones, I.G.; Fairhurst, V.; Sealy-Lewis, H.M. ADHII in Aspergillus nidulans is induced by carbon starvation stress. Fungal Genet. Biol. 2001, 32, 33–43.
  39. Kelly, J.M.; Drysdale, M.R.; Sealy-Lewis, H.M.; Jones, I.G.; Lockington, R.A. Alcohol dehydrogenase III in Aspergillus nidulans is anaerobically induced and post-transcriptionally regulated. Mol. Gen. Genet. 1990, 222, 323–328.
  40. Zhang, E.; Cao, Y.; Xia, Y. Ethanol Dehydrogenase I Contributes to Growth and Sporulation Under Low Oxygen Condition via Detoxification of Acetaldehyde in Metarhizium acridum. Front. Microbiol. 2018, 9, 1932.
  41. Callejas-Negrete, O.A.; Torres-Guzmán, J.C.; Padilla-Guerrero, I.E.; Esquivel-Naranjo, U.; Padilla-Ballesteros, M.F.; García-Tapia, A.; Schrank, A.; Salazar-Solís, E.; Gutiérrez-Corona, F.; González-Hernández, G.A. The Adh1 gene of the fungus Metarhizium anisopliae is expressed during insect colonization and required for full virulence. Microbiol. Res. 2015, 172, 57–67.
  42. Matsuzaki, F.; Shimizu, M.; Wariishi, H. Proteomic and metabolomic analyses of the white-rot fungus Phanerochaete chrysosporium exposed to exogenous benzoic acid. J. Proteome Res. 2008, 7, 2342–2350.
  43. Yang, D.D.; François, J.M.; de Billerbeck, G.M. Cloning, expression and characterization of an aryl-alcohol dehydrogenase from the white-rot fungus Phanerochaete chrysosporium strain BKM-F-1767. BMC Microbiol. 2012, 12, 126.
  44. Raj, S.B.; Ramaswamy, S.; Plapp, B.V. Yeast alcohol dehydrogenase structure and catalysis. Biochemistry 2014, 53, 5791–5803.
  45. Reiser, J.; Muheim, A.; Hardegger, M.; Frank, G.; Fiechter, A. Aryl-alcohol dehydrogenase from the white-rot fungus Phanerochaete chrysosporium. Gene cloning, sequence analysis, expression, and purification of the recombinant enzyme. J. Biol. Chem. 1994, 269, 28152–28159.
  46. Ichinose, H.; Wariishi, H.; Tanaka, H. Molecular analysis of arylalcohol dehydrogenase of Coriolus versicolor expressed against exogenous addition of dibenzothiophene derivatives. J. Basic Microbiol. 2002, 42, 327–336.
  47. Trinci, A.P.J.; Davies, D.R.; Gull, K.; Lawrence, M.I.; Bonde Nielsen, B.; Rickers, A.; Theodorou, M.K. Anaerobic fungi in herbivorous animals. Mycol. Res. 1994, 98, 129–152.
  48. Hackstein, J.H.; Akhmanova, A.; Boxma, B.; Harhangi, H.R.; Voncken, F.G. Hydrogenosomes: Eukaryotic adaptations to anaerobic environments. Trends Microbiol. 1999, 7, 441–447.
  49. Boxma, B.; Voncken, F.; Jannink, S.; van Alen, T.; Akhmanova, A.; van Weelden, S.W.; van Hellemond, J.J.; Ricard, G.; Huynen, M.; Tielens, A.G.; et al. The anaerobic chytridiomycete fungus Piromyces sp. E2 produces ethanol via pyruvate:formate lyase and an alcohol dehydrogenase E. Mol. Microbiol. 2004, 51, 1389–1399.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , ,
View Times: 254
Revisions: 2 times (View History)
Update Date: 10 Oct 2023
1000/1000
ScholarVision Creations