1887

Abstract

The gene of sp. strain LB400 has previously been shown to be located within the locus, which specifies the degradation of biphenyl (BP) and chlorobiphenyls, and to encode a glutathione -transferase (GST) which accepts 1-chloro-2,4-dinitrobenzene (CDNB) as substrate. The specific physiological role of this gene is not known. It is now shown that the gene is expressed in the parental organism and that GST activity is induced more than 20-fold by growth of the strain on BP relative to succinate when these compounds serve as sole carbon source. Approximately the same induction factor was observed for 2,3-dihydroxybiphenyl 1,2-dioxygenase activity, which is encoded by the 5′-adjacent gene. This suggests that the expression of is coregulated with the expression of genes responsible for the catabolism of BP. A probe detected only a single copy of the gene in strain LB400. A spontaneous BP mutant of the organism neither gave a signal with the probe nor showed CDNB-accepting GST activity, suggesting that this activity is solely encoded by . Complementation of the mutant with a gene cluster devoid of restored the ability to grow on BP, indicating that is not essential for utilization of this carbon source. BphK activity proved to be almost unaffected by up to 100-fold differences in proton concentration or ionic strength. The enzyme showed a narrow range with respect to a variety of widely used electrophilic GST substrates, accepting only CDNB. A number of established laboratory strains as well as novel isolates able to grow on BP as sole carbon and energy source were examined for BphK activity and the presence of a analogue. CDNB assays, probe hybridizations and PCR showed that several, but not all, BP degraders possess this type of GST activity and/or a closely related gene. In all bacteria showing BphK activity, this was induced by growth on BP as sole carbon source, although activity levels differed by up to 10-fold after growth on BP and by up to 60-fold after growth on succinate. This resulted in a variation of induction factors between 2 and 30. In the majority of bacteria examined, the gene appeared to be part of LB400-like gene clusters. DNA sequencing revealed almost complete identity of genes from five different gene clusters. These results suggest that genes, although not essential, fulfil a strain-specific function related to the utilization of BPs by their host organisms. The usefulness of BphK as a reporter enzyme for monitoring the expression of catabolic pathways is discussed.

Loading

Article metrics loading...

/content/journal/micro/10.1099/00221287-145-10-2821
1999-10-01
2024-04-20
Loading full text...

Full text loading...

/deliver/fulltext/micro/145/10/1452821a.html?itemId=/content/journal/micro/10.1099/00221287-145-10-2821&mimeType=html&fmt=ahah

References

  1. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. (1990). Basic local alignment search tool. J Mol Biol 215, 403-410.[CrossRef] [Google Scholar]
  2. Altschul, S. F., Madden, T. L., Schäffer, A. A., Zhang, J., Zhang, Z., Miller, W. & Lipman, D. J. (1997). Gapped blast and psi-blast: a new generation of protein database search programs. Nucleic Acids Res 25, 3389-3402.[CrossRef] [Google Scholar]
  3. Aoki, Y., Satoh, K., Sato, K. & Suzuki, K. T. (1992). Induction of glutathione S-transferase P-form in primary cultured rat liver parenchymal cells by co-planar polychlorinated biphenyl congeners. Biochem J 281, 539-543. [Google Scholar]
  4. Arca, P., Rico, M., Brana, A. F., Villar, C. J., Hardisson, C. & Suarez, J. E. (1988). Formation of an adduct between fosfomycin and glutathione: a new mechanism of antibiotic resistance in bacteria. Antimicrob Agents Chemother 32, 1552-1556.[CrossRef] [Google Scholar]
  5. Armengaud, J. & Timmis, K. N. (1997). Molecular characterization of Fdx1, a putidaredoxin-type [2Fe–2S] ferredoxin able to transfer electrons to the dioxin dioxygenase of Sphingomonas sp. RW1. Eur J Biochem 247, 833-842.[CrossRef] [Google Scholar]
  6. Asturias, J. A., Eltis, L. D., Prucha, M. & Timmis, K. N. (1994a). Analysis of three 2,3-dihydroxybiphenyl-1,2-dioxygenases found in Rhodococcus globerulus P6. J Biol Chem 269, 7807-7815. [Google Scholar]
  7. Asturias, J. A., Moore, E., Yakimov, M. M., Klatte, S. & Timmis, K. N. (1994b). Reclassification of the polychlorinated biphenyl-degraders Acinetobacter sp. strain P6 and Corynebacterium sp. strain MB1 as Rhodococcus globerulus. Syst Appl Microbiol 17, 226-231.[CrossRef] [Google Scholar]
  8. Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. D., Smith, J. A. & Struhl, K. (1994).Current Protocols in Molecular Biology. New York: Wiley.
  9. Bauchop, T. & Elsden, S. R. (1960). The growth of micro-organisms in relation to their energy supply. J Gen Microbiol 23, 457-469.[CrossRef] [Google Scholar]
  10. Bedard, D. L., Haberl, M. L., May, R. J. & Brennan, M. J. (1987). Evidence for novel mechanisms of polychlorinated biphenyl metabolism in Alcaligenes eutrophus H850. Appl Environ Microbiol 53, 1103-1112. [Google Scholar]
  11. Blocki, F. A., Ellis, L. B. M. & Wackett, L. P. (1993). MIF proteins are theta-class glutathione S-transferase homologs. Protein Sci 2, 2095-2102.[CrossRef] [Google Scholar]
  12. Blumenroth, P. (1997).Konstruktion gentechnisch modifizierter Mikroorganismen für den in situ-Abbau polychlorierter Biphenyle (PCBs) in Sediment. PhD thesis, Technical University of Braunschweig, Germany.
  13. Board, P. G., Baker, R. T., Chelvanayagam, G. & Jermiin, L. S. (1997). Zeta, a novel class of glutathione transferases in a range of species from plants to humans. Biochem J 328, 929-935. [Google Scholar]
  14. Bopp, L. H. (1986). Degradation of highly chlorinated PCBs by Pseudomonas strain LB400. J Ind Microbiol 1, 23-29.[CrossRef] [Google Scholar]
  15. Bopp, L. H., Chakrabarty, A. M. & Ehrlich, H. L. (1983). Chromate resistance plasmid in Pseudomonas fluorescens. J Bacteriol 155, 1105-1109. [Google Scholar]
  16. 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]
  17. Burlingame, R. & Chapman, P. J. (1983). Catabolism of phenylpropionic acid and its 3-hydroxy derivative by Escherichia coli. J Bacteriol 155, 113-121. [Google Scholar]
  18. Catelani, D., Colombi, A., Sorlini, C. & Treccani, V. (1973). 2-Hydroxy-6-oxo-6-phenylhexa-2,4-dienoate: the meta-cleavage product from 2,3-dihydroxybiphenyl by Pseudomonas putida. Biochem J 134, 1063-1066. [Google Scholar]
  19. Crawford, R. L. & Frick, T. D. (1977). Rapid spectrophotometric differentiation between glutathione-dependent and glutathione-independent gentisate and homogentisate pathways. Appl Environ Microbiol 34, 170-174. [Google Scholar]
  20. Daubaras, D. L., Saido, K. & Chakrabarty, A. M. (1996). Purification of 1,2-dioxygenase and maleylacetate reductase: the lower pathway of 2,4,5-trichlorophenoxyacetic acid metabolism by Burkholderia capacia AC1100. Appl Environ Microbiol 62, 4276-4279. [Google Scholar]
  21. Di Ilio, C., Aceto, A., Piccolomini, R., Allocati, N., Faraone, A., Cellini, L., Ravagnan, G. & Federici, G. (1988). Purification and characterization of three forms of glutathione transferase from Proteus mirabilis. Biochem J 255, 971-975. [Google Scholar]
  22. Di Ilio, C., Aceto, A., Allocati, N. & 7 other authors (1993). Characterization of glutathione transferase from Xanthomonas campestris. Arch Biochem Biophys 305, 110–114.[CrossRef] [Google Scholar]
  23. Eaton, R. W. (1994). Organization and evolution of naphthalene catabolic pathways: sequence of the DNA encoding 2-hydroxychromene-2-carboxylate isomerase and trans-o-hydroxybenzylidenepyruvate hydratase-aldolase from the NAH7 plasmid. J Bacteriol 176, 7757-7762. [Google Scholar]
  24. Eltis, L. D., Hofmann, B., Hecht, H.-J., Lünsdorf, H. & Timmis, K. N. (1993). Purification and crystallization of 2,3-dihydroxybiphenyl-1,2-dioxygenase. J Biol Chem 268, 2727-2732. [Google Scholar]
  25. Erb, R. W. & Wagner-Döbler, I. (1993). Detection of polychlorinated biphenyl degradation genes in polluted sediments by direct DNA extraction and polymerase chain reaction. Appl Environ Microbiol 59, 4065-4073. [Google Scholar]
  26. Erickson, B. & Mondello, F. (1992). Nucleotide sequencing and transcriptional mapping of the genes encoding biphenyl dioxygenase, a multicomponent polychlorinated-biphenyl-degrading enzyme in Pseudomonas strain LB400. J Bacteriol 174, 2903-2912. [Google Scholar]
  27. Fahey, R. C. & Sundquist, A. R. (1991). Evolution of glutathione metabolism. Adv Enzymol Relat Areas Mol Biol 64, 1-53. [Google Scholar]
  28. Favaloro, B., Melino, S., Petruzzelli, R., Di Ilio, C. & Rotilio, D. (1998a). Purification and characterization of a novel glutathione transferase from Ochrobactrum anthropi. FEMS Microbiol Lett 160, 81-86.[CrossRef] [Google Scholar]
  29. Favaloro, B., Tamburro, A., Angelucci, S., De Luca, A., Di Ilio, C. & Rotilio, D. (1998b). Molecular cloning, expression and site-directed mutagenesis of glutathione S-transferase from Ochrobactrum anthropi. Biochem J 335, 573-579. [Google Scholar]
  30. Ferrandez, A., Garcia, J. L. & Diaz, E. (1997). Genetic characterization and expression in heterologous hosts of the 3-(3-hydroxyphenyl)propionate catabolic pathway of Escherichia coli K12. J Bacteriol 179, 2573-2581. [Google Scholar]
  31. Fleischmann, R. D., Adams, M. D., White, O. & 37 other authors (1995). Whole-genome random sequencing and assembly of Haemophilus influenzae Rd. Science 269, 496–512.[CrossRef] [Google Scholar]
  32. Fujita, M., Hanada, Y. & Adachi, Y. (1994). Glutathione S-transferases in sarcocarp tissue of pumpkin fruit exposed to 2,4-dichlorophenoxyacetic acid. Biosci Biotechnol Biochem 58, 1349-1350.[CrossRef] [Google Scholar]
  33. Furukawa, K., Matsumura, F. & Tonomura, K. (1978).Alcaligenes and Acinetobacter strains capable of degrading polychlorinated biphenyls. Agric Biol Chem 42, 543-548.[CrossRef] [Google Scholar]
  34. Furukawa, K., Simon, J. R. & Chakrabarty, A. M. (1983). Common induction and regulation of biphenyl, xylene/toluene and salicylate catabolism in Pseudomonas paucimobilis Q1. J Bacteriol 154, 1356-1362. [Google Scholar]
  35. Grant, S. G. N., Jessee, J., Bloom, F. R. & Hanahan, D. (1990). Differential plasmid rescue from transgenic mouse DNAs into Escherichia coli methylation-restriction mutants. Proc Natl Acad Sci USA 87, 4645-4649.[CrossRef] [Google Scholar]
  36. Gutell, R. R., Weiser, B., Woese, C. R. & Noller, H. F. (1985). Comparative anatomy of 16S-like ribosomal RNA. Prog Nucleic Acid Res Mol Biol 32, 155-216. [Google Scholar]
  37. Habig, W. H., Pabst, M. J. & Jakoby, B. J. (1974). Glutathione S-transferases. The first enzymatic step in mercapturic acid formation. J Biol Chem 249, 7130-7139. [Google Scholar]
  38. Hayes, J. D. & Pulford, D. J. (1995). The glutathione S-transferase supergene family: regulation of GST and the contribution of the isoenzymes to cancer chemoprotection and drug resistance. Crit Rev Biochem Mol Biol 30, 445-600.[CrossRef] [Google Scholar]
  39. Henikoff, S. & Henikoff, J. G. (1992). Amino acid substitution matrices from protein blocks. Proc Natl Acad Sci USA 89, 10915-10919.[CrossRef] [Google Scholar]
  40. Hofer, B., Eltis, D., Dowling, D. & Timmis, K. N. (1993). Genetic analysis of a Pseudomonas locus encoding a pathway for biphenyl/polychlorinated biphenyl degradation. Gene 130, 47-55.[CrossRef] [Google Scholar]
  41. Hofer, B., Backhaus, S. & Timmis, K. N. (1994). The biphenyl/polychlorinated biphenyl-degradation locus (bph) of Pseudomonas sp. LB400 encodes four additional metabolic enzymes. Gene 144, 9-16.[CrossRef] [Google Scholar]
  42. van Hylckama Vlieg, J. E. T., Kingma, J., van den Wijngaard, A. J. & Janssen, D. B. (1998). A glutathione S-transferase with activity towards cis-1,2-dichloroepoxyethane is involved in isoprene utilization by Rhodococcus sp. strain AD45. Appl Environ Microbiol 64, 2800-2805. [Google Scholar]
  43. Iizuka, M., Inoue, Y., Murata, K. & Kimura, A. (1989). Purification and some properties of glutathione S-transferase from Escherichia coli B. J Bacteriol 171, 6039-6042. [Google Scholar]
  44. Karlson, U., Dwyer, D. F., Hooper, S. W., Moore, E. R. B., Timmis, K. N. & Eltis, L. D. (1993). Two independently regulated cytochromes P-450 in a Rhodococcus rhodochrous strain that degrades 2-ethoxyphenol and 4-methoxybenzoate. J Bacteriol 175, 1467-1474. [Google Scholar]
  45. Lane, D. J. (1991). 16S/23S sequencing. In Nucleic Acid Techniques in Bacterial Systematics, pp. 115-175. Edited by E. Stackebrandt & M. Goodfellow. Chichester: Wiley.
  46. La Roche, S. D. & Leisinger, T. (1990). Sequence analysis and expression of the bacterial dichloromethane dehalogenase structural gene, a member of the glutathione S-transferase supergene family. J Bacteriol 172, 164-171. [Google Scholar]
  47. Lloyd-Jones, G. & Lau, P. C. K. (1997). Glutathione S-transferase-encoding gene as a potential probe for environmental bacterial isolates capable of degrading polycyclic aromatic hydrocarbons. Appl Environ Microbiol 63, 3286-3290. [Google Scholar]
  48. Maidak, B. L., Olsen, G. J., Larsen, N., Overbeek, R., McCaughey, M. J. & Woese, C. R. (1997). The RDP (ribosomal database project). Nucleic Acids Res 25, 109-110.[CrossRef] [Google Scholar]
  49. Mannervik, B. (1985). The isoenzymes of glutathione transferase. Adv Enzymol Relat Areas Mol Biol 57, 357-417. [Google Scholar]
  50. Masai, E., Katayama, Y., Kubota, S., Kawai, S., Yamasaki, M. & Morohoshi, N. (1993). A bacterial enzyme degrading the model lignin compound β-etherase is a member of the glutathione S-transferase superfamily. FEBS Lett 323, 135-140.[CrossRef] [Google Scholar]
  51. Moore, E. R. B., Wittich, R.-M., Fortnagel, P. & Timmis, K. N. (1993). 16S ribosomal RNA gene sequence characterization and phylogenetic analysis of a dibenzo-p-dioxin degrading isolate within the new genus Sphingomonas. Lett Appl Microbiol 17, 115-118.[CrossRef] [Google Scholar]
  52. Morales, V. M., Bäckman, A. & Bagdasarian, M. (1991). A series of wide-host-range low-copy-number vectors that allow direct screening for recombinants. Gene 97, 39-47.[CrossRef] [Google Scholar]
  53. Mullis, K. B. & Faloona, F. (1987). Specific synthesis of DNA in vitro via a polymerase-catalyzed chain reaction. Methods Enzymol 155, 335-350. [Google Scholar]
  54. Nishida, M., Kong, K.-H., Inoue, H. & Takahashi, K. (1994). Molecular cloning and site-directed mutagenesis of glutathione S-transferase from Escherichia coli. The conserved tyrosyl residue near the N terminus is not essential for catalysis. J Biol Chem 269, 32536-32541. [Google Scholar]
  55. Nohynek, L. J., Suhonen, E. L., Nurmiaho-Lassila, E. L., Hantula, J. & Salkinoja-Salonen, M. (1996). Description of four pentachlorophenol-degrading bacterial strains as Sphingomonas chlorophenolica sp. nov. Syst Appl Microbiol 18, 527-538. [Google Scholar]
  56. Perito, B., Allocati, N., Casalone, E., Masulli, M., Dragani, B., Polsinelli, M., Aceto, A. & Di Ilio, C. (1996). Molecular cloning and overexpression of a glutathione transferase gene from Proteus mirabilis. Biochem J 318, 157-162. [Google Scholar]
  57. Pickett, C. B. & Lu, A. Y. H. (1989). Glutathione S-transferases: gene structure, regulation, and biological function. Annu Rev Biochem 58, 743-764.[CrossRef] [Google Scholar]
  58. Prohaska, J. R. (1980). The glutathione peroxidase activity of glutathione S-transferases. Biochim Biophys Acta 611, 87-98.[CrossRef] [Google Scholar]
  59. Sambrook, J., Fritsch, E. F. & Maniatis, T. (1989).Molecular Cloning: a Laboratory Manual, 2nd edn. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory.
  60. Seeger, M., Timmis, K. N. & Hofer, B. (1995). Conversion of chlorobiphenyls into phenylhexadienoates and benzoates by the enzymes of the upper pathway for polychlorobiphenyl degradation encoded by the bph locus of Pseudomonas sp. strain LB400. Appl Environ Microbiol 61, 2654-2658. [Google Scholar]
  61. Shin, H.-J., Kim, S.-J. & Kim, Y.-C. (1997). Sequence analysis of the phnD gene encoding 2-hydroxymuconic semialdehyde hydrolase in Pseudomonas sp. strain DJ77. Biochem Biophys Res Commun 232, 288-291.[CrossRef] [Google Scholar]
  62. Simon, R., Priefer, U. & Pühler, A. (1983). A broad host range mobilization system for in vivo genetic engineering: transposon mutagenesis in Gram negative bacteria. Biotechnology 1, 784-790.[CrossRef] [Google Scholar]
  63. Stoesser, G., Sterk, P., Tuli, M. A., Stoehr, P. J. & Cameron, G. N. (1997). The EMBL nucleotide sequence database. Nucleic Acids Res 25, 7-13.[CrossRef] [Google Scholar]
  64. Studier, F. W. (1991). Use of bacteriophage T7 lysozyme to improve an inducible T7 expression system. J Mol Biol 219, 37-44.[CrossRef] [Google Scholar]
  65. Studier, F. W. & Moffat, B. A. (1986). Use of bacteriophage T7 RNA polymerase to direct selective high-level expression of cloned genes. J Mol Biol 189, 113-130.[CrossRef] [Google Scholar]
  66. Taira, K., Hayase, N., Arimura, N., Yamashita, S., Miyazaki, T. & Furukawa, K. (1988). Cloning and nucleotide sequence of the 2,3-dihydroxybiphenyl dioxygenase gene from the PCB-degrading strain of Pseudomonas paucimobilis Q1. Biochemistry 27, 3990-3996.[CrossRef] [Google Scholar]
  67. Tanno, K. & Aoki, Y. (1996). Phosphorylation of c-Jun stimulated in primary cultured rat liver parenchymal cells by a coplanar polychlorinated biphenyl. Biochem J 313, 863-866. [Google Scholar]
  68. Toung, Y.-P. S., Hsieh, T.-s. & Tu, C.-P. D. (1993). The glutathione S-transferase D genes. A divergently organized, intronless gene family in Drosophila melanogaster. J Biol Chem 268, 9737-9746. [Google Scholar]
  69. Vuilleumier, S. (1997). Bacterial glutathione S-transferases: what are they good for? J Bacteriol 179, 1431-1441. [Google Scholar]
  70. Wang, Y. & Lau, P. C. (1996). Sequence and expression of an isocitrate dehydrogenase-encoding gene from a polycyclic aromatic hydrocarbon oxidizer, Sphingomonas yanoikuyae B1. Gene 168, 15-21.[CrossRef] [Google Scholar]
  71. Wang, Y., Lau, P. C. & Button, D. K. (1996). A marine oligobacterium harboring genes known to be part of aromatic hydrocarbon degradation pathways of soil pseudomonads. Appl Environ Microbiol 62, 2169-2173. [Google Scholar]
  72. Werwath, J., Arfmann, H.-A., Pieper, D. H., Timmis, K. N. & Wittich, R.-M. (1998). Biochemical and genetic characterization of a gentisate 1, 2-dioxygenase from Sphingomonas sp. strain RW5. J Bacteriol 180, 4171-4176. [Google Scholar]
  73. Williams, W. A., Lobos, J. H. & Cheetham, W. E. (1997). A phylogenetic analysis of aerobic polychlorinated biphenyl-degrading bacteria. Int J Syst Bacteriol 47, 207-210.[CrossRef] [Google Scholar]
  74. Wittich, R.-M., Wilkes, H., Sinnwell, V., Francke, W. & Fortnagel, P. (1992). Metabolism of dibenzo-p-dioxin by Sphingomonas sp. strain RW1. Appl Environ Microbiol 58, 1005-1010. [Google Scholar]
  75. Woese, C. R., Gutel, R. R., Gupta, R. & Noller, H. F. (1983). Detailed higher-order structure of 16S-like ribosomal ribonucleic acids. Microbiol Rev 47, 621-669. [Google Scholar]
  76. Xinhua, J., Zhang, P., Armstrong, R. N. & Gilliland, G. L. (1992). The three-dimensional structure of a glutathione S-transferase from the Mu gene class. Structural analysis of the binary complex of isoenzyme 3–3 and glutathione at 2·2-Å resolution. Biochemistry 31, 10169-10184.[CrossRef] [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/micro/10.1099/00221287-145-10-2821
Loading
/content/journal/micro/10.1099/00221287-145-10-2821
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