首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The dibenzofuran-degrading bacterial strain DPO360 represents a new species of the genus Terrabacter together with the previously described dibenzofuran-mineralizing bacterial strain DPO1361 (K.-H. Engesser, V. Strubel, K. Christoglou, P. Fischer, and H. G. Rast, FEMS Microbiol. Lett. 65:205-210, 1989; V. Strubel, Ph.D. thesis, University of Stuttgart, Stuttgart, Germany, 1991; V. Strubel, H. G. Rast, W. Fietz, H.-J. Knackmuss, and K.-H. Engesser, FEMS Microbiol. Lett. 58:233-238, 1989). Two 2,3-dihydroxybiphenyl-1,2-dioxygenases (BphC1 and BphC2) and one catechol-2,3-dioxygenase (C23O) were shown to be expressed in Terrabacter sp. strain DPO360 growing with dibenzofuran as a sole source of carbon and energy. These enzymes exhibited strong sensitivity to oxygen. They were purified to apparent homogeneity as homodimers (BphC and BphC2) and as a homotetrameric catechol-2,3-dioxygenase (C23O). According to their specificity constants kcat/Km, both BphC1 and BphC2 were shown to be responsible for the cleavage of 2,2',3-trihydroxybiphenyl, the first metabolite in dibenzofuran mineralization along the angular dioxygenation pathway. With this substrate, BphC2 exhibited a considerably higher kcat/Km, value (183 microM/min) than BphC1 (29 microM/min). Catechol-2,3-dioxygenase was recognized to be not involved in the ring cleavage of 2,2',3-trihydroxybiphenyl (kcat/Km, 1 microM/min). Analysis of deduced amino acid sequence data of bphC1 revealed 36% sequence identity to nahC from Pseudomonas putida PpG7 (S. Harayama and M. Rekik, J. Biol. Chem. 264:15328-15333, 1989) and about 40% sequence identity to various bphC genes from different Pseudomonas and Rhodococcus strains. In addition, another 2,3-dihydroxybiphenyl-1,2-dioxygenase gene (bphC3) was cloned from the genome of Terrabacter sp. strain DPO360. Expression of this gene, however, could not be detected in Terrabacter sp. strain DPO360 after growth with dibenzofuran.  相似文献   

2.
The cloned bphA gene of the polychlorinated biphenyl (PCB) degrader Rhodococcus sp. strain RHA1 was expressed in Rhodococcus erythropolis IAM1399 cells, resulting in the transformation of di-, tri-, and tetrachlorobiphenyls. Disruption of the bphA1 gene in RHA1 resulted in a lack of growth on biphenyl and a loss of PCB transformation activity. However, the bphA1 insertion mutant of RHA1, designated RDA1, retained the ability to transform PCB congeners when grown on ethylbenzene as its carbon source. It also transformed 4-chlorobiphenyl to 4-chlorobenzoate, although it was suspected to be deficient in bphB and bphC gene activities as well as bphA. This suggested that an alternative PCB degradation system distinct from the one encoded by the cloned bph genes was present.  相似文献   

3.
Yang X  Xie F  Zhang G  Shi Y  Qian S 《Biochimie》2008,90(10):1530-1538
The genes of two 2,3-dihydroxybiphenyl 1,2-dioxygenases (BphC1 and BphC2) were obtained from the gene library of Rhodococcus sp. R04. The enzymes have been purified to apparent electrophoretic homogeneity from the cell extracts of the recombinant harboring bphC1 and bphC2. Both BphC1 and BphC2 were hexamers, consisting of six subunits of 35 and 33kDa as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, respectively. The enzymes had similar optimal pH (pH 9.0), but different temperatures for their maximum activity (30 degrees C for BphC1, 80 degrees C for BphC2). In addition, they exhibited distinct stability at various temperatures. The enzymes could cleave a wide range of catechols, with 2,3-dihydroxybiphenyl being the optimum substrate for BphC1 and BphC2. BphC1 was inhibited by 2,3-dihydroxybiphenyl, catechol and 3-chlorocatechol, whereas BphC2 showed strong substrate inhibition for all the given substrates. BphC2 exhibited a half-life of 15min at 80 degrees C and 50min at 70 degrees C, making it the most thermostable extradiol dioxygenase studied in mesophilic bacteria. After disruption of bphC1 and bphC2 genes, R04DeltaC1 (bphC1 mutant) delayed the time of their completely eliminating biphenyl another 15h compared with its parent strain R04, but R04DeltaC2 (bphC2 mutant) lost the ability to grow on biphenyl, suggesting that BphC1 plays an assistant role in the degrading of biphenyl by strain R04, while BphC2 is essential for the growth of strain R04 on biphenyl.  相似文献   

4.
Rhodococcus rhodochrous K37, a Gram-positive bacterium grown under alkaline conditions, was isolated for its ability to metabolize PCBs. Analysis revealed that it has eight genes encoding extradiol dioxygenase, which has 2,3-dihydroxybiphenyl 1,2-dioxygenase activity, and these genes were designated bphC1 to bphC8. According to the classification of extradiol dioxygenases [Eltis, L. D., and Bolin, J. T., J. Bacteriol., 178, 5930-5937 (1996)], BphC3 and BphC6 belong to the type II enzyme group. The other six BphCs were classified as members of the type I extradiol dioxygenase group. BphC4 and BphC8 were classified into a new subfamily of type I, family 3. Two linear plasmids, 200 kb and 270 kb in size, were found in K37, and the bphC6 and bphC8 genes were located in the 200 kb linear plasmid. Northern hybridization analysis revealed that the bphC1, bphC2, and bphC7 genes were induced in the presence of testosterone, the bphC6 gene was induced by fluorene, and the bphC8 gene was induced by biphenyl. All eight BphC products exhibited much higher substrate activity for 2,3-dihydroxybiphenyl than for catechol, 3-methylcatechol, or 4-methylcatechol.  相似文献   

5.
Rhodococcus sp. strain DK17 is known to metabolize o-xylene and toluene through the intermediates 3,4-dimethylcatechol and 3- and 4-methylcatechol, respectively, which are further cleaved by a common catechol 2,3-dioxygenase. A putative gene encoding this enzyme (akbC) was amplified by PCR, cloned, and expressed in Escherichia coli. Assessment of the enzyme activity expressed in E. coli combined with sequence analysis of a mutant gene demonstrated that the akbC gene encodes the bona fide catechol 2,3-dioxygenase (AkbC) for metabolism of o-xylene and alkylbenzenes such as toluene and ethylbenzene. Analysis of the deduced amino acid sequence indicates that AkbC consists of a new catechol 2,3-dioxygenase class specific for methyl-substituted catechols. A computer-aided molecular modeling studies suggest that amino acid residues (particularly Phe177) in the beta10-beta11 loop play an essential role in characterizing the substrate specificity of AkbC.  相似文献   

6.
1,2-Dihydroxynaphthalene dioxygenase was purified to homogeneity from a bacterium that degrades naphthalenesulfonic acids (strain BN6). The enzyme requires Fe2+ for maximal activity and consists of eight identical subunits with a molecular weight of about 33,000. Analysis of the NH2-terminal amino acid sequence revealed a high degree of homology (22 of 29 amino acids) with the NH2-terminal amino acid sequence of 2,3-dihydroxybiphenyl dioxygenase from strain Pseudomonas paucimobilis Q1. 1,2-Dihydroxynaphthalene dioxygenase from strain BN6 shows a wide substrate specificity and also cleaves 5-, 6-, and 7-hydroxy-1,2-dihydroxynaphthalene, 2,3- and 3,4-dihydroxybiphenyl, catechol, and 3-methyl- and 4-methylcatechol. Similar activities against the hydroxy-1,2-dihydroxynaphthalenes were also found in cell extracts from naphthalene-degrading bacteria.  相似文献   

7.
Mycobacterium bovis BCG and Mycobacterium tuberculosis possess a single arylamine N-acetyltransferase whose gene is predicted to occur within a six-gene operon. Deletion of the nat gene caused an extended lag phase in M. bovis BCG and a cell morphology associated with an altered pattern of cell wall mycolates. Analysis of cDNA from M. bovis BCG shows that during in vitro growth all the genes in the putative nat operon are expressed and the open reading frames are contiguous, supporting the existence of an operon. Two genes in the operon, Mb3599c and Mb3600c, are predicted to encode homologues of enzymes annotated as a 2,3-dihydroxybiphenyl 1,2-dioxygenase (bphC5) and a 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate hydrolase (bphD2), respectively, in Rhodococcus RHA1. As predicted, M. bovis BCG cell lysates metabolized the BphC substrate 2,3-dihydroxybiphenyl (2,3-DHB) to 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA), a BphD substrate, which was subsequently hydrolysed. Immunoprecipitation of the BphD homologue from these lysates led to an accumulation of HOPDA. M. bovis BCG growth on both solid and liquid media was inhibited with either 2,3-DHB or an inhibitor of BphC, 3-chlorocatechol (3-CC). In addition, incubation with 2,3-DHB affects the lipid composition of the cell wall resulting in a diminished level of mycolates and an altered cell morphology similar to the Deltanat strain. We propose the enzymes encoded by the putative operon have a similar endogenous role to that of the NAT enzyme and are part of a pathway important for cell wall synthesis.  相似文献   

8.
Rhodococcus globerulus strain P6 contains at least three genes, bphC1, bphC2, and bphC3, coding for 2,3-dihydroxybiphenyl 1,2-dioxygenases; the latter two specify enzymes of the family of one-domain extradiol dioxygenases. In order to assess the importance of these different isoenzymes for the broad catabolic activity of this organism towards the degradation of polychlorinated biphenyls (PCBs), the capacities of recombinant enzymes expressed in Escherichia coli to transform different chlorosubstituted dihydroxybiphenyls formed by the action of R. globerulus P6 biphenyl dioxygenase and biphenyl 2,3-dihydrodiol dehydrogenase were determined. Whereas both BphC2 and BphC3 showed similar activities for 2,3-dihydroxybiphenyl and all monochlorinated 2,3-dihydroxybiphenyls, BphC1 exhibited only weak activity for 2'-chloro-2,3-dihydroxybiphenyl. More highly chlorinated 2'-chlorosubstituted 2,3-dihydroxybiphenyls were also transformed at high rates by BphC2 and BphC3 but not BphC1. In R. globerulus P6, BphC2 was constitutively expressed, BphC1 expression was induced during growth on biphenyl, and BphC3 was not expressed at significant levels under the experimental conditions. Although we cannot rule out the expression of BphC3 under certain environmental conditions, it seems that the contrasting substrate specificities of BphC1 and BphC2 contribute significantly to the versatile PCB-degrading phenotype of R. globerulus P6.  相似文献   

9.
10.
Four kinds of polychlorinated biphenyl (PCB)-degrading Rhodococcus sp. (TA421, TA431, HA99, and K37) have been isolated from termite ecosystem and under alkaline condition. The bph gene cluster involved in the degradation of PCB/biphenyl has been analyzed in strain TA421. This gene cluster was highly homologous to bph gene clusters in R. globerulus P6 and Rhodococcus sp. RHA1. In this study, we cloned and analyzed the bph gene cluster essential to PCB/biphenyl degradation from R. rhodochrous K37. The order of the genes and the sequence were different in K37 than in P6, RHA1, and TA421. The bphC8(K37) gene was more homologous to the meta-cleavage enzyme involved in phenanthrene metabolism than bphC genes involved in biphenyl metabolism. Two other Rhodococcus strains (HA99 and TA431) had PCB/biphenyl degradation gene clusters similar to that in K37. These findings suggest that these bph gene clusters evolved separately from the well-known bph gene clusters of PCB/biphenyl degraders.  相似文献   

11.
The 2,3-dihydroxybiphenyl dioxygenase from Sphingomonas sp. strain BN6 (BphC1-BN6) differs from most other extradiol dioxygenases by its ability to oxidize 3-chlorocatechol to 3-chloro-2-hydroxymuconic semialdehyde by a distal cleavage mechanism. The turnover of different substrates and the effects of various inhibitors on BphC1-BN6 were compared with those of another 2,3-dihydroxybiphenyl dioxygenase from the same strain (BphC2-BN6) as well as with those of the archetypical catechol 2,3-dioxygenase (C23O-mt2) encoded by the TOL plasmid. Cell extracts containing C23O-mt2 or BphC2-BN6 converted the relevant substrates with an almost constant rate for at least 10 min, whereas BphC1-BN6 was inactivated significantly within the first minutes during the turnover of all substrates tested. Furthermore, BphC1-BN6 was much more sensitive than the other two enzymes to inactivation by the Fe(II) ion-chelating compound o-phenanthroline. The reason for inactivation of BphC1-BN6 appeared to be the loss of the weakly bound ferrous ion, which is the cofactor in the catalytic center. A mutant enzyme of BphC1-BN6 constructed by site-directed mutagenesis showed a higher stability to inactivation by o-phenanthroline and an increased catalytic efficiency for the conversion of 2,3-dihydroxybiphenyl and 3-methylcatechol but was still inactivated during substrate oxidation.  相似文献   

12.
Polybrominated diphenyl ethers (PBDEs) are a class of flame retardants that have been widely used in consumer products, but that are problematic because of their environmental persistence and endocrine‐disrupting properties. To date, very little is known about PBDE degradation by aerobic microorganisms and the enzymes involved in PBDE transformation. Resting cells of the polychlorinated biphenyl‐degrading actinomycete, Rhodococcus jostii RHA1, depleted nine mono‐ through penta‐BDEs in separate assays. Extensive depletion of PBDEs occurred with cells grown on biphenyl, ethylbenzene, propane, or styrene, whereas very limited depletion occurred with cells grown on pyruvate or benzoate. In RHA1, expression of bphAa encoding biphenyl dioxygenase (BPDO) and etbAa1 and etbAc encoding ethylbenzene dioxygenase (EBDO) was induced 30‐ to 3,000‐fold during growth on the substrates that supported PBDE depletion. The BPDO and EBDO enzymes had gene expression profiles that matched the PBDE‐depletion profiles exhibited by RHA1 grown on different substrates. Using the non‐PBDE‐degrading bacterium Rhodococcus erythropolis as a host, two recombinant strains were developed by inserting the eth and bph genes of RHA1, respectively. The resultant EBDO extensively depleted mono‐ through penta‐BDEs, while the BPDO depleted only mono‐, di‐, and one tetra‐BDE. A dihydroxylated‐BDE was detected as the primary metabolite of 4‐bromodiphenyl ether in both recombinant strains. These results indicate that although both dioxygenases are capable of transforming PBDEs, EBDO more potently transforms the highly brominated congeners. The availability of substrates or inducing compounds can markedly affect total PBDE removal as well as patterns of removal of individual congeners. Biotechnol. Bioeng. 2011;108: 313–321. © 2010 Wiley Periodicals, Inc.  相似文献   

13.
Benzoate catabolism is thought to play a key role in aerobic bacterial degradation of biphenyl and polychlorinated biphenyls (PCBs). Benzoate catabolic genes were cloned from a PCB degrader, Rhodococcus sp. strain RHA1, by using PCR amplification and temporal temperature gradient electrophoresis separation. A nucleotide sequence determination revealed that the deduced amino acid sequences encoded by the RHA1 benzoate catabolic genes, benABCDK, exhibit 33 to 65% identity with those of Acinetobacter sp. strain ADP1. The gene organization of the RHA1 benABCDK genes differs from that of ADP1. The RHA1 benABCDK region was localized on the chromosome, in contrast to the biphenyl catabolic genes, which are located on linear plasmids. Escherichia coli cells containing RHA1 benABCD transformed benzoate to catechol via 2-hydro-1,2-dihydroxybenzoate. They transformed neither 2- nor 4-chlorobenzoates but did transform 3-chlorobenzoate. The RHA1 benA gene was inactivated by insertion of a thiostrepton resistance gene. The resultant mutant strain, RBD169, neither grew on benzoate nor transformed benzoate, and it did not transform 3-chlorobenzoate. It did, however, exhibit diminished growth on biphenyl and growth repression in the presence of a high concentration of biphenyl (13 mM). These results indicate that the cloned benABCD genes could play an essential role not only in benzoate catabolism but also in biphenyl catabolism in RHA1. Six rhodococcal benzoate degraders were found to have homologs of RHA1 benABC. In contrast, two rhodococcal strains that cannot transform benzoate were found not to have RHA1 benABC homologs, suggesting that many Rhodococcus strains contain benzoate catabolic genes similar to RHA1 benABC.  相似文献   

14.
AIMS: Isolation of the genes relative to PCB biodegradation and identification of the bph gene function in Rhodococcus sp. R04. METHODS AND RESULTS: A 8.7-kb fragment carrying the biphenyl catabolic genes bphABCD was isolated from the gene library in Rhodococcus sp. R04. Based on the deduced amino acid sequence homology, seven bph genes, bphA1A2A3A4, bphB, bphC and bphD, were thought to be responsible for the initial four steps of biphenyl degradation. In Escherichia coli, BphA exhibited poor activity for biphenyl transformation, and BphB, BphC and BphD were found to be catalytically active towards 2,3-dihydro-2,3-dihydroxybiphenyl, 2,3-dihydroxybiphenyl and 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate, respectively (activities of 50, 8.1 and 2.4 micromol l(-1) min(-1) mg(-1)). SDS-PAGE analysis indicated that the sizes of bphA1A2A3A4, bphB, bphC and bphD gene products were 49, 19, 14, 47, 32, 30 and 31 kDa, respectively. After disruption of bph genes, the bphA1 mutants lost the ability to grow on biphenyl, the bphB and bphD mutants were able to transform a little of biphenyl, but hardly grew on biphenyl. CONCLUSION: The cloned bph genes indeed play an important role in the biphenyl catabolism in this strain. SIGNIFICANCE AND IMPACT OF THE STUDY: This bph gene organization in Rhodococcus sp. R04 differs from that of other biphenyl degraders reported previously, indicating it is a novel type of bph gene cluster. Analysis of the phylogenetic tree suggested that BphA1 and BphA2 in Rhodococcus sp. R04 had a different evolutionary relationship with those in the other PCB degraders.  相似文献   

15.
16.
A bphC gene (915 bp) encoding 2,3-dihydroxybiphenyl 1,2-dioxygenase (BphC) was amplified by PCR from Dyella ginsengisoli LA-4, which was heterologously expressed in Escherichia coli . The purified His-Tag BphC was able to catalyze the meta -cleavage reaction of the dihydroxylated aromatic rings. According to the specificity constant ( K cat/ K m) of BphC_LA-4, the specificity of BphC_LA-4 was determined in the following order: 2,3-dihydroxybiphenyl>3-methylcatechol>catechol>4-chlorocatechol>4-methylcatechol. The experimental data were consistent with the prediction of enzyme–substrate complexes. The highest specific activity of BphC_LA-4 was 118.3 U mg−1 for 2,3-dihydroxybiphenyl.  相似文献   

17.
Proteomics and targeted gene disruption were used to investigate the catabolism of benzene, styrene, biphenyl, and ethylbenzene in Rhodococcus jostii RHA1, a well-studied soil bacterium whose potent polychlorinated biphenyl (PCB)-transforming properties are partly due to the presence of the related Bph and Etb pathways. Of 151 identified proteins, 22 Bph/Etb proteins were among the most abundant in biphenyl-, ethylbenzene-, benzene-, and styrene-grown cells. Cells grown on biphenyl, ethylbenzene, or benzene contained both Bph and Etb enzymes and at least two sets of lower Bph pathway enzymes. By contrast, styrene-grown cells contained no Etb enzymes and only one set of lower Bph pathway enzymes. Gene disruption established that biphenyl dioxygenase (BPDO) was essential for growth of RHA1 on benzene or styrene but that ethylbenzene dioxygenase (EBDO) was not required for growth on any of the tested substrates. Moreover, whole-cell assays of the ΔbphAa and etbAa1::cmrA etbAa2::aphII mutants demonstrated that while both dioxygenases preferentially transformed biphenyl, only BPDO transformed styrene. Deletion of pcaL of the β-ketoadipate pathway disrupted growth on benzene but not other substrates. Thus, styrene and benzene are degraded via meta- and ortho-cleavage, respectively. Finally, catalases were more abundant during growth on nonpolar aromatic compounds than on aromatic acids. This suggests that the relaxed specificities of BPDO and EBDO that enable RHA1 to grow on a range of compounds come at the cost of increased uncoupling during the latter's initial transformation. The stress response may augment RHA1's ability to degrade PCBs and other pollutants that induce similar uncoupling.  相似文献   

18.
The green fluorescent protein gene (gfp) was inserted into the chromosome of Pseudomonas spp. Cam-1 and Sag-50G, two psychrotolerant polychlorinated biphenyl (PCB)-degrading bacteria. The gfp-transformed microorganisms, designated Cam-1-gfp1, Cam-1-gfp2, Sag-50G-gfp1, and Sag-50G-gfp2, exhibited green fluorescence under an epifluorescent microscope. The gfp was inserted into the chromosome of each psychrotolerant strain and was stable with no apparent adverse affects on the metabolism and growth of each organism. Activity of gfp-transformed microorganisms against biphenyl and 2,3-dichlorobiphenyl was determined by assaying for BphC activity and by resting cell assays. The patterns of BphC activity at two different growth temperatures in batch cultures were similar for each of the gfp-transformed microorganisms. Resting cell assays of both the parent strains (Cam-1, Sag-50G) and the gfp-transformed strains (Cam-1-gfp1, Cam-1-gfp2, Sag-50G-gfp1, Sag-50G-gfp2), grown on glycerol or glucose, exhibited BphC activity to a lesser extent and at a slower rate than those observed for biphenyl grown cells. In addition, all gfp-transformed microorganisms degraded 2,3-dichlorobiphenyl (2,3-DCB) in broth to the same extent as the parent strains. When Cam-1-gfp1 and Sag-50G-gfp1 were used as a bioremediation amendment in soil microcosms spiked with 2,3-DCB, both strains survived in high numbers (5.6 to 7.9 log cfu g?1 and 5.6 to 8.0 log cfu g?1) when inoculated into nonsterilized soil over 16 weeks at 22 degrees C and 18 weeks at 4 degrees C, respectively. Biodegradation of 2,3-DCB was enhanced with the microbial amendment; however, the addition of sunflower oil did not help the PCB degrading bacteria and may have enhanced the growth of the indigenous population, thereby decreasing the amended PCB-degrading population.  相似文献   

19.
The actinobacterium Rhodococcus jostii RHA1 grows on a remarkable variety of aromatic compounds and has been studied for applications ranging from the degradation of polychlorinated biphenyls to the valorization of lignin, an underutilized component of biomass. In RHA1, the catabolism of two classes of lignin-derived compounds, alkylphenols and alkylguaiacols, involves a phylogenetically distinct extradiol dioxygenase, AphC, previously misannotated as BphC, an enzyme involved in biphenyl catabolism. To better understand the role of AphC in RHA1 catabolism, we first showed that purified AphC had highest apparent specificity for 4-propylcatechol (kcat/KM ∼106 M−1 s−1), and its apparent specificity for 4-alkylated substrates followed the trend for alkylguaiacols: propyl > ethyl > methyl > phenyl > unsubstituted. We also show AphC only poorly cleaved 3-phenylcatechol, the preferred substrate of BphC. Moreover, AphC and BphC cleaved 3-phenylcatechol and 4-phenylcatechol with different regiospecificities, likely due to the substrates’ binding mode. A crystallographic structure of the AphC·4-ethylcatechol binary complex to 1.59 Å resolution revealed that the catechol is bound to the active site iron in a bidentate manner and that the substrate’s alkyl side chain is accommodated by a hydrophobic pocket. Finally, we show RHA1 grows on a mixture of 4-ethylguaiacol and guaiacol, simultaneously catabolizing these substrates through meta-cleavage and ortho-cleavage pathways, respectively, suggesting that the specificity of AphC helps to prevent the routing of catechol through the Aph pathway. Overall, this study contributes to our understanding of the bacterial catabolism of aromatic compounds derived from lignin, and the determinants of specificity in extradiol dioxygenases.  相似文献   

20.
Bacterial strain M213 was isolated from a fuel oil-contaminated soil in Idaho, USA, by growth on naphthalene as a sole source of carbon, and was identified as Rhodococcus opacus M213 by 16S rDNA sequence analysis and growth on substrates characteristic of this species. M213 was screened for growth on a variety of aromatic hydrocarbons, and growth was observed only on simple 1 and 2 ring compounds. No growth or poor growth was observed with chlorinated aromatic compounds such as 2,4-dichlorophenol and chlorobenzoates. No growth was observed by M213 on salicylate, and M213 resting cells grown on naphthalene did not attack salicylate. In addition, no salicylate hydroxylase activity was detected in cell free lysates, suggesting a pathway for naphthalene catabolism that does not pass through salicylate. Enzyme assays indicated induction of catechol 1,2-dioxygenase and catechol 2,3-dioxygenase on different substrates. Total DNA from M213 was screened for hybridization with a variety of genes encoding catechol dioxygenases, but hybridization was observed only with catA (encoding catechol 1,2-dioxygenase) from R. opacus 1CP and edoD (encoding catechol 2,3-dioxygenase) from Rhodococcus sp. I1. Plasmid analysis indicated the presence of two plasmids (pNUO1 and pNUO2). edoD hybridized to pNUO1, a very large (approximately 750 kb) linear plasmid.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号