1887

Abstract

DCL14 assimilates all stereoisomers of carveol and dihydrocarveol as sole source of carbon and energy. Induction experiments with carveol- or dihydrocarveol-grown cells showed high oxygen consumption rates with these two compounds and with carvone and dihydrocarvone. (Dihydro)carveol-grown cells of DCL14 contained the following enzymic activities involved in the carveol and dihydrocarveol degradation pathways of this micro-organism: (dihydro)carveol dehydrogenase (both NAD- and dichlorophenolindophenol-dependent activities), an unknown cofactor-dependent carvone reductase, (iso-)dihydrocarvone isomerase activity, NADPH-dependent dihydrocarvone monooxygenase (Baeyer–Villiger monooxygenase), ϵ-lactone hydrolase and an NAD-dependent 6-hydroxy-3-isopropenylheptanoate dehydrogenase. Product accumulation studies identified (4)-carvone, (1,4)-dihydrocarvone, (4,7)-4-isopropenyl-7-methyl-2-oxo-oxepanone, (3)-6-hydroxy-3-isopropenylheptanoate, (3)-3-isopropenyl-6-oxoheptanoate, (3,6)-6-isopropenyl-3-methyl-2-oxo-oxepanone and (5)-6-hydroxy-5-isopropenyl-2-methylhexanoate as intermediates in the (4)-carveol degradation pathway. The opposite stereoisomers of these compounds were identified in the (4)-carveol degradation pathway. With dihydrocarveol, the same intermediates are involved except that carvone was absent. These results show that DCL14 metabolizes all four diastereomers of carveol via oxidation to carvone, which is subsequently stereospecifically reduced to (1)-(iso-)dihydrocarvone. At this point also dihydrocarveol enters the pathway, and this compound is directly oxidized to (iso-)dihydrocarvone. Cell extracts contained both (1)-(iso-)dihydrocarvone 1,2-monooxygenase and (1)-(iso)-dihydrocarvone 2,3-monooxygenase activity, resulting in a branch point of the degradation pathway; (1)-(iso-)dihydrocarvone was converted to 4-isopropenyl-7-methyl-2-oxo-oxepanone, while (1)-(iso)-dihydrocarvone, which is isomerized to (1)-(iso-)dihydrocarvone, was converted to 6-isopropenyl-3-methyl-2-oxo-oxepanone. 4-Isopropenyl-7-methyl-2-oxo-oxepanone is hydrolysed to 6-hydroxy-3-isopropenylheptanoate, which is subsequently oxidized to 3-isopropenyl-6-oxoheptanoate, thereby linking the (dihydro)carveol degradation pathways to the limonene degradation pathway of this micro-organism. 6-Isopropenyl-3-methyl-2-oxo-oxepanone is, , hydrolysed to 6-hydroxy-5-isopropenyl-2-methylhexanoate, which is thought to be a dead-end metabolite.

Loading

Article metrics loading...

/content/journal/micro/10.1099/00221287-146-5-1129
2000-05-01
2024-04-20
Loading full text...

Full text loading...

/deliver/fulltext/micro/146/5/1461129a.html?itemId=/content/journal/micro/10.1099/00221287-146-5-1129&mimeType=html&fmt=ahah

References

  1. Alphand V., Furstoss R. 1992; Microbiological transformations. 23. A surprising regioselectivity of microbiological Baeyer–Villiger oxidations of menthone and dihydrocarvone. Tetrahedron: Asymmetry 3:379–382 [CrossRef]
    [Google Scholar]
  2. Banthorpe D. V. 1994; Terpenoids. In Natural Products: Their Chemistry and Biological Significance pp. 289–359Edited by Mann J., Davidson R. S., Hobbs J. B., Banthorpe D. V., Harborne J. B. Harlow, UK: Longman;
    [Google Scholar]
  3. Barbirato F., Verdoes J. C., de Bont J. A. M., van der Werf M. J. 1998; The Rhodococcus erythropolis DCL14 limonene-1,2-epoxide hydrolase gene encodes an enzyme belonging to a novel class of epoxide hydrolases. FEBS Lett 438:293–296 [CrossRef]
    [Google Scholar]
  4. Bradford M. M. 1976; A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein–dye binding. Anal Biochem 72:248–254 [CrossRef]
    [Google Scholar]
  5. Buckingham J. 1998 In Dictionary of Natural Products on CD-ROM Version 6·1 London: Chapman & Hall;
    [Google Scholar]
  6. Burdock G. A. 1995 Fenaroli’s Handbook of Flavor Ingredients, 3rd edn. Boca Raton, FL: CRC Press;
    [Google Scholar]
  7. van Dyk M. S., van Rensburg E., Rensburg I. P. B., Moleleki N. 1998; Biotransformation of monoterpenoid ketones by yeasts and yeast-like fungi. J Mol Catal B (Enzymatic) 5:149–154 [CrossRef]
    [Google Scholar]
  8. Guenther A., Hewitt C. N., Erickson D.13 other authors 1995; A global model of natural volatile organic compound emissions. J Geophys Res 100:8873–8892 [CrossRef]
    [Google Scholar]
  9. Harborne J. B. 1991; Recent advances in the ecological chemistry of plant terpenoids. In Ecological Chemistry and Biochemistry of Plant Terpenoids pp. 399–426Edited by Harborne J. B., Tomas-Barberan F. A. Oxford: Clarendon Press;
    [Google Scholar]
  10. Hartmans S., Smits J. P., van der Werf M. J., Volkering F., de Bont J. A. M. 1989; Metabolism of styrene oxide and 2-phenylethanol in the styrene-degrading Xanthobacter strain 124X. Appl Environ Microbiol 55:2850–2855
    [Google Scholar]
  11. Hirata T., Hamada H., Aoki T., Suga T. 1982; Stereoselectivity of the reduction of carvone and dihydrocarvone by suspensions of cells of Nicotiana tabacum. Phytochemistry 21:2209–2212 [CrossRef]
    [Google Scholar]
  12. McCaskill D., Croteau R. 1997; Prospects for the bioengineering of isoprenoid biosynthesis. Adv Biochem Eng Biotechnol 55:107–146
    [Google Scholar]
  13. Mironowicz A., Siewiński A. 1986; Biotransformations. XIX. Reduction of some terpenic ketones by means of immobilized cells of Rhodotorula mucilaginosa. Acta Biotechnol 6:141–146 [CrossRef]
    [Google Scholar]
  14. Mironowicz A., Raczkowska J., Siewiński A., Szykuła J., Zabża A. 1982; Microbiological transformations. Part XIV. Microbiological reduction of (+)-carvone and eucarvone by Rhodotorula mucilaginosa. Pol J Chem 56:735–739
    [Google Scholar]
  15. Nishimura H., Hiramoto S., Mizutani J., Noma Y., Furusaki A., Matsumoto T. 1983; Structure and biological activity of bottrospicatol, a novel monoterpene produced by microbial transformation of (−)-cis-carveol. Agric Biol Chem 47:2697–2699 [CrossRef]
    [Google Scholar]
  16. Noma Y. 1980; Conversion of (−)-carvone by strains of Streptomyces, A-5-1, and Nocardia, 1-3-11. Agric Biol Chem 44:807–812 [CrossRef]
    [Google Scholar]
  17. Noma Y., Asakawa Y. 1992; Enantio- and diastereoselectivity in the biotransformation of carveols by Euglena gracilis Z. Phytochemistry 31:2009–2011 [CrossRef]
    [Google Scholar]
  18. Noma Y., Nishimura H. 1987; Bottrospicatols, novel monoterpenes produced on conversion of (−)- and (+)-cis-carveol by Streptomyces. Agric Biol Chem 51:1845–1849 [CrossRef]
    [Google Scholar]
  19. Noma Y., Nonomura S. 1974; Conversion of (−)-carvone and (+)-carvone by a strain of Aspergillus niger. Agric Biol Chem 38:741–744 [CrossRef]
    [Google Scholar]
  20. Noma Y., Nonomura S., Ueda H., Tatsumi C. 1974; Conversion of (+)-carvone by Pseudomonas ovalis, strain 6-1. Agric Biol Chem 38:735–740 [CrossRef]
    [Google Scholar]
  21. Noma Y., Nonomura S., Sakai H. 1975; Epimerization of (−)-isodihydrocarvone to (−)-dihydrocarvone by Pseudomonas fragi IFO 3458. Agric Biol Chem 39:437–441 [CrossRef]
    [Google Scholar]
  22. Noma Y., Nishimura H., Hiramoto S., Iwami M., Tatsumi C. 1982; A new compound, (4R,6R)-(+)-6,8-oxidomenth-1-en-9-ol produced by microbial conversion of (−)-cis-carveol. Agric Biol Chem 46:2871–2872 [CrossRef]
    [Google Scholar]
  23. Ravid U., Putievsky E., Katzir I., Weinstein V., Ikan R. 1992; Chiral GC analysis of (S)(+)- and (R)(−)-carvone with high enantiomeric purity in caraway, dill and spearmint oils. Flavour Fragrance J 7:289–292 [CrossRef]
    [Google Scholar]
  24. Sauers R. R., Ahearn G. P. 1961; The importance of steric effects in the Baeyer–Villiger oxidation. J Am Chem Soc 83:2759–2762 [CrossRef]
    [Google Scholar]
  25. Shimoda K., Hirata T., Noma Y. 1998; Stereochemistry in the reduction of enones by the reductase from Euglena gracilis Z. Phytochemistry 49:49–53 [CrossRef]
    [Google Scholar]
  26. Trudgill P. W. 1986; Terpenoid metabolism by Pseudomonas. In The Bacteria. A Treatise on Structure and Function vol. XThe Biology of Pseudomonas pp. 483–525Edited by Sokatch J. R. Orlando, FL: Academic Press;
    [Google Scholar]
  27. Trudgill P. W. 1990; Microbial metabolism of monoterpenes – recent developments. Biodegradation 1:93–105 [CrossRef]
    [Google Scholar]
  28. Trudgill P. W. 1994; Microbial metabolism and transformation of selected monoterpenes. In Biochemistry of Microbial Degradation pp. 33–61Edited by Ratledge C. Dordrecht: Kluwer;
    [Google Scholar]
  29. Varie D. L., Brennan J., Briggs B., Cronin J. S., Hay D. A., Rieck J. A., Zmijewski M. J. 1998; Bioreduction of (R)-carvone and regioselective Baeyer–Villiger oxidations: application to the asymmetric synthesis of cryptophycin fragment A. Tetrahedron Lett 39:8405–8408 [CrossRef]
    [Google Scholar]
  30. Verstegen-Kaaksma A. A., Swarts H. J., Jansen B. J. M., de Groot A., Bottema-MacGillavry N., Witholt B. 1995; Application of S-(+)-carvone in the synthesis of biologically active natural products using chemical transformations and bioconversions. Ind Crops Prod 4:15–21 [CrossRef]
    [Google Scholar]
  31. Welsh F. W., Murray W. D., Williams R. E. 1989; Microbiological and enzymatic production of flavor and fragrance chemicals. Crit Rev Biotechnol 9:105–169 [CrossRef]
    [Google Scholar]
  32. van der Werf M. J. 2000; Purification and characterization of a Baeyer–Villiger monooxygenase from Rhodococcus erythropolis DCL14 involved in three different monocyclic monoterpene degradation pathways. Biochem J (in press)
    [Google Scholar]
  33. van der Werf M. J., Guettler M. V., Jain M. K., Zeikus J. G. 1997; Environmental and physiological factors affecting the succinate product ratio during carbohydrate fermentation by Actinobacillus sp. 130Z. Arch Microbiol 167:332–342 [CrossRef]
    [Google Scholar]
  34. van der Werf M. J., Overkamp K. M., de Bont J. A. M. 1998; Limonene-1,2-epoxide hydrolase from Rhodococcus erythropolis DCL14 belongs to a novel class of epoxide hydrolases. J Bacteriol 180:5052–5057
    [Google Scholar]
  35. van der Werf M. J., Swarts H. J., de Bont J. A. M. 1999a; Rhodococcus erythropolis DCL14 contains a novel degradation pathway for limonene. Appl Environ Microbiol 65:2092–2102
    [Google Scholar]
  36. van der Werf M. J., van der Ven C., Barbirato F., Eppink M. H. M., de Bont J. A. M., van Berkel W. J. 1999b; Stereoselective carveol dehydrogenase from Rhodococcus erythropolis DCL14. A novel nicotinoprotein belonging to the short-chain dehydrogenase/reductase superfamily. J Biol Chem 274:26296–26304 [CrossRef]
    [Google Scholar]
  37. Williams D. R., Trudgill P. W. 1994; Ring cleavage reactions in the metabolism of (−)-menthol and (−)-menthone by a Corynebacterium sp. Microbiology 140:611–616 [CrossRef]
    [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/micro/10.1099/00221287-146-5-1129
Loading
/content/journal/micro/10.1099/00221287-146-5-1129
Loading

Data & Media loading...

This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error