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 125 毫秒
1.
Muconate cycloisomerase (EC 5.5.1.1) and chloromuconate cycloisomerase (EC 5.5.1.7) were purified from extracts of Rhodococcus erythropolis 1CP cells grown with benzoate or 4-chlorophenol, respectively. Both enzymes discriminated between the two possible directions of 2-chloro-cis, cis-muconate cycloisomerization and converted this substrate to 5-chloromuconolactone as the only product. In contrast to chloromuconate cycloisomerases of gram-negative bacteria, the corresponding R. erythropolis enzyme is unable to catalyze elimination of chloride from (+)-5-chloromuconolactone. Moreover, in being unable to convert (+)-2-chloromuconolactone, the two cycloisomerases of R. erythropolis 1CP differ significantly from the known muconate and chloromuconate cycloisomerases of gram-negative strains. The catalytic properties indicate that efficient cycloisomerization of 3-chloro- and 2,4-dichloro-cis,cis-muconate might have evolved independently among gram-positive and gram-negative bacteria.  相似文献   

2.
Muconate cycloisomerases are known to catalyze the reversible conversion of 2-chloro-cis,cis-muconate by 1,4- and 3,6-cycloisomerization into (4S)-(+)-2-chloro- and (4R/5S)-(+)-5-chloromuconolactone. 2-Chloromuconolactone is transformed by muconolactone isomerase with concomitant dechlorination and decarboxylation into the antibiotic protoanemonin. The low k(cat) for this compound compared to that for 5-chloromuconolactone suggests that protoanemonin formation is of minor importance. However, since 2-chloromuconolactone is the initially predominant product of 2-chloromuconate cycloisomerization, significant amounts of protoanemonin were formed in reaction mixtures containing large amounts of muconolactone isomerase and small amounts of muconate cycloisomerase. Such enzyme ratios resemble those observed in cell extracts of benzoate-grown cells of Ralstonia eutropha JMP134. In contrast, cis-dienelactone was the predominant product formed by enzyme preparations, in which muconolactone isomerase was in vitro rate limiting. In reaction mixtures containing chloromuconate cycloisomerase and muconolactone isomerase, only minute amounts of protoanemonin were detected, indicating that only small amounts of 2-chloromuconolactone were formed by cycloisomerization and that chloromuconate cycloisomerase actually preferentially catalyzes a 3,6-cycloisomerization.  相似文献   

3.
The tfdC(I)D(I)E(I)F(I,) and tfdD(II)C(II)E(II)F(II) gene modules of plasmid pJP4 of Ralstonia eutropha JMP134 encode complete sets of functional enzymes for the transformation of chlorocatechols into 3-oxoadipate, which are all expressed during growth on 2,4-dichlorophenoxyacetate (2,4-D). However, activity of tfd(I)-encoded enzymes was usually higher than that of tfd(II)-encoded enzymes, both in the wild-type strain grown on 2,4-D and in 3-chlorobenzoate-grown derivatives harboring only one tfd gene module. The tfdD(II)-encoded chloromuconate cycloisomerase exhibited special kinetic properties, with high activity against 3-chloromuconate and poor activity against 2-chloromuconate and unsubstituted muconate, thus explaining the different phenotypic behaviors of R. eutropha strains containing different tfd gene modules. The enzyme catalyzes the formation of an equilibrium between 2-chloromuconate and 5-chloro- and 2-chloromuconolactone and very inefficiently catalyzes dehalogenation to form trans-dienelactone as the major product, thus differing from all (chloro)muconate cycloisomerases described thus far.  相似文献   

4.
The conversion of 2-chloro-cis,cis-muconate by muconate cycloisomerase from Pseudomonas putida PRS2000 yielded two products which by nuclear magnetic resonance spectroscopy were identified as 2-chloro- and 5-chloromuconolactone. High-pressure liquid chromatography analyses showed the same compounds to be formed also by muconate cycloisomerases from Acinetobacter calcoaceticus ADP1 and Pseudomonas sp. strain B13. During 2-chloro-cis,cis-muconate turnover by the enzyme from P. putida, 2-chloromuconolactone initially was the major product. After prolonged incubation, however, 5-chloromuconolactone dominated in the resulting equilibrium. In contrast to previous assumptions, both chloromuconolactones were found to be stable at physiological pH. Since the chloromuconate cycloisomerases of Pseudomonas sp. strain B13 and Alcaligenes eutrophus JMP134 have been shown previously to produce the trans-dienelactone (trans-4-carboxymethylene-but-2-en-4-olide) from 2-chloro-cis,cis-muconate, they must have evolved the capability to cleave the carbon-chlorine bond during their divergence from normal muconate cycloisomerases.  相似文献   

5.
2-Chloromuconate cycloisomerase from the Gram-positive bacterium Rhodococcus opacus 1CP (Rho-2-CMCI) is an enzyme of a modified ortho-pathway, in which 2-chlorophenol is degraded using 3-chlorocatechol as the central intermediate. In general, the chloromuconate cycloisomerases catalyze not only the cycloisomerization, but also the process of dehalogenation of the chloromuconate to dienelactone. However Rho-2-CMCI, unlike the homologous enzymes from the Gram-negative bacteria, is very specific for only one position of the chloride on the substrate chloromuconate. Furthermore, Rho-2-CMCI is not able to dehalogenate the 5-chloromuconolactone and therefore it cannot generate the dienelactone.  相似文献   

6.
Chloromuconate cycloisomerases of bacteria utilizing chloroaromatic compounds are known to convert 3-chloro-cis,cis-muconate to cis-dienelactone (cis-4-carboxymethylenebut-2-en-4-olide), while usual muconate cycloisomerases transform the same substrate to the bacteriotoxic protoanemonin. Formation of protoanemonin requires that the cycloisomerization of 3-chloro-cis,cis-muconate to 4-chloromuconolactone is completed by protonation of the exocyclic carbon of the presumed enol/enolate intermediate before chloride elimination and decarboxylation take place to yield the final product. The formation of cis-dienelactone, in contrast, could occur either by dehydrohalogenation of 4-chloromuconolactone or, more directly, by chloride elimination from the enol/enolate intermediate. To reach a better understanding of the mechanisms of chloride elimination, the proton-donating Lys169 of Pseudomonas putida muconate cycloisomerase was changed to alanine. As expected, substrates requiring protonation, such as cis,cis-muconate as well as 2- and 3-methyl-, 3-fluoro-, and 2-chloro-cis,cis-muconate, were not converted at a significant rate by the K169A variant. However, the variant was still active with 3-chloro- and 2,4-dichloro-cis,cis-muconate. Interestingly, cis-dienelactone and 2-chloro-cis-dienelactone were formed as products, whereas the wild-type enzyme forms protoanemonin and the not previously isolated 2-chloroprotoanemonin, respectively. Thus, the chloromuconate cycloisomerases may avoid (chloro-)protoanemonin formation by increasing the rate of chloride abstraction from the enol/enolate intermediate compared to that of proton addition to it.  相似文献   

7.
Evolution of chlorocatechol catabolic pathways   总被引:15,自引:0,他引:15  
The aerobic bacterial degradation of chloroaromatic compounds often involves chlorosubstituted catechols as central intermediates. They are converted to 3-oxoadipate in a series of reactions similar to that for catechol catabolism and therefore designated as modifiedortho-cleavage pathway. Among the enzymes of this catabolic route, the chlorocatechol 1,2-dioxygenases are known to have a relaxed substrate specificity. In contrast, several chloromuconate cycloisomerases are more specific, and the dienelactone hydrolases of chlorocatechol catabolic pathways do not even convert the corresponding intermediate of catechol degradation, 3-oxoadipate enol-lactone. While the sequences of chlorocatechol 1,2-dioxygenases and chloromuconate cycloisomerases are very similar to those of catechol 1,2-dioxygenases and muconate cycloisomerases, respectively, the relationship between dienelactone hydrolases and 3-oxoadipate enol-lactone hydrolases is more distant. They seem to share an / hydrolase fold, but the sequences comprising the fold are quite dissimilar. Therefore, for chlorocatechol catabolism, dienelactone hydrolases might have been recruited from some other, preexisting pathway. Their relationship to dienelactone (hydrolases identified in 4-fluorobenzoate utilizing strains ofAlcaligenes andBurkholderia (Pseudomonas) cepacia is investigated). Sequence evidence suggests that the chlorocatechol catabolic operons of the plasmids pJP4, pAC27, and pP51 have been derived from a common precursor. The latter seems to have evolved for the purpose of halocatechol catabolism, and may be considerably older than the chemical industry.  相似文献   

8.
To elucidate possible reasons for the recalcitrance of 2-chlorotoluene, the metabolism of chloromethylcatechols, formed after dioxygenation and dehydrogenation by Ralstonia sp. strain PS12 tetrachlorobenzene dioxygenase and chlorobenzene dihydrodiol dehydrogenase, was monitored using chlorocatechol dioxygenases and chloromuconate cycloisomerases partly purified from Ralstonia sp. strain PS12 and Wautersia eutropha JMP134. Two chloromethylcatechols, 3-chloro-4-methylcatechol and 4-chloro-3-methylcatechol, were formed from 2-chlorotoluene. 3-Chloro-4-methylcatechol was transformed into 5-chloro-4-methylmuconolactone and 2-chloro-3-methylmuconolactone. For mechanistic reasons neither of these cycloisomerization products can be dehalogenated by chloromuconate cycloisomerases, with the result that 3-chloro-4-methylcatechol cannot be mineralized by reaction sequences related to catechol ortho-cleavage pathways known thus far. 4-Chloro-3-methylcatechol is only poorly dehalogenated during enzymatic processing due to the kinetic properties of the chloromuconate cycloisomerases. Thus, degradation of 2-chlorotoluene via a dioxygenolytic pathway is evidently problematic. In contrast, 5-chloro-3-methylcatechol, the major dioxygenation product formed from 3-chlorotoluene, is subject to quantitative dehalogenation after successive transformation by chlorocatechol 1,2-dioxygenase and chloromuconate cycloisomerase, resulting in the formation of 2-methyldienelactone. 3-Chloro-5-methylcatechol is transformed to 2-chloro-4-methylmuconolactone.  相似文献   

9.
The chloromuconate cycloisomerase of Pseudomonas sp. B13 was purified from 3-chlorobenzoate-grown wild-type cells while the chloromuconate cycloisomerases of Ralstonia eutropha JMP134 (pJP4) and Pseudomonas sp. P51 (pP51) were purified from Escherichia coli strains expressing the corresponding gene. Kinetic studies were performed with various chloro-, fluoro-, and methylsubstituted cis,cis-muconates. 2,4-Dichloro-cis,cis-muconate proved to be the best substrate for all three chloromuconate cycloisomerases. Of the three enzymes, TfdD of Ralstonia eutropha JMP134 (pJP4) was most specific, since its specificity constant for 2,4-dichloro-cis,cis-muconate was the highest, while the constants for cis,cis-muconate, 2-chloro- and 2,5-dichloro-cis,cis-muconate were especially poor. The sequence of ClcB of the 3-chlorobenzoate-utilizing strain Pseudomonas sp. B13 was determined and turned out to be identical to that of the corresponding enzyme of pAC27 (though slightly different from the published sequences). Corresponding to 2-chloro-cis,cis-muconate being a major metabolite of 3-chlorobenzoate degradation, the k cat/K m with 2-chloro-cis,cis-muconate was relatively high, while that with the still preferred substrate 2,4-dichloro-cis,cis-muconate was relatively low. This enzyme was thus the least specific and the least active among the three compared enzymes. TcbD of Pseudomonas sp. P51 (pP51) took an intermediate position with respect to both the degree of specificity and the activity with the preferred substrate. Received: 7 August 1998 / Received revision: 24 November 1998 / Accepted: 29 November 1998  相似文献   

10.
Muconate cycloisomerases play a crucial role in the bacterial degradation of aromatic compounds by converting cis,cis-muconate, the product of catechol ring cleavage, to (4S)-muconolactone. Chloromuconate cycloisomerases catalyze both the corresponding reaction and a dehalogenation reaction in the transformation of chloroaromatic compounds. This study reports the first thorough examination of the substrate specificity of the muconate cycloisomerases from Pseudomonas putida PRS2000 and Acinetobacter calcoaceticus” ADP1. We show that they transform, in addition to cis,cis-muconate, 3-fluoro-, 2-methyl-, and 3-methyl-cis,cis-muconate with high specificity constants but not 2-fluoro-, 2-chloro-, 3-chloro-, or 2,4-dichloro-cis,cis-muconate. Based on known three-dimensional structures, variants of P. putida muconate cycloisomerase were constructed by site-directed mutagenesis to contain amino acids found in equivalent positions in chloromuconate cycloisomerases. Some of the variants had significantly increased specificity constants for 3-chloro- or 2,4-dichloromuconate (e.g., A271S and I54V showed 27- and 22-fold increases, respectively, for the former substrate). These kinetic improvements were not accompanied by a change from protoanemonin to cis,cis-dienelactone as the product of 3-chloro-cis,cis-muconate conversion. The rate of 2-chloro-cis,cis-muconate turnover was not significantly improved, nor was this compound dehalogenated to any significant extent. However, the direction of 2-chloro-cis,cis-muconate cycloisomerization could be influenced by amino acid exchange. While the wild-type enzyme discriminated only slightly between the two possible cycloisomerization directions, some of the enzyme variants showed a strong preference for either (+)-2-chloro- or (+)-5-chloromuconolactone formation. These results show that the different catalytic characteristics of muconate and chloromuconate cycloisomerases are due to a number of features that can be changed independently of each other.  相似文献   

11.
D Ghosal  I S You 《Gene》1989,83(2):225-232
Alcaligenes eutrophus harboring plasmid pJP4 (strain JMP134) is capable of growing on both 2,4-dichlorophenoxyacetate (2,4-D) and 3-chlorobenzoate (3-Cba), while Pseudomonas putida carrying plasmid pAC27 (strain AC867) can utilize only 3-Cba as the sole carbon source. The tfdCDEF operon of the pJP4 plasmid and the clcABD operon of plasmid pAC27 each encode enzymes for the degradation of chlorocatechols (Clc), key intermediates in the catabolism of 2,4-D and 3-Cba. Similarities in the nucleotide (nt) sequences of genes tfdC and clcA, encoding pyrocatechases, were reported earlier [Ghosal and You, Mol. Gen. Genet. 211 (1988a) 113-120]. Genes tfdD and clcB, encoding Clc-specific cycloisomerases, have been completely sequenced. The tfdD gene (1107 bp) is slightly smaller than gene clcB (1113 bp). Comparison of the two cycloisomerase-encoding genes reveals that the nt sequences are 63% homologous with 62% homology in the deduced amino acid (aa) sequences of the polypeptides they encode. Genes tfdD and tfdE are contiguous in the tfdCDEF operon, whereas the corresponding genes, clcB and clcD, of the clcABD operon, are known to be separated by a long open reading frame of unknown function. The predicted N-terminal aa sequences of the two hydrolase-encoding genes, tfdE and clcD, also show homology. The structural and nt homologies between the two Clc operons, tfdCDEF and clcABD, suggest their relatedness.  相似文献   

12.
Plasmid pJP4 of Alcaligenes eutrophus JMP134 encodes the degradation of 2,4-dichlorophenoxyacetic acid. A 1.2-kb BamHI-XhoI region of the restriction fragment BamHI-E has been proposed to contain the regulatory gene tfdR (A. R. Harker, R. H. Olsen, and R. J. Seidler, J. Bacteriol. 171:314-320, 1989; B. Kaphammer, J. J. Kukor, and R. H. Olsen, J. Bacteriol. 172:2280-2286, 1990). When sequenced and analyzed, the region is shown to contain two incomplete open reading frames (ORFs) positioned divergently. The complete DNA sequence for one of the two ORFs was obtained by sequencing the adjacent restriction fragment BamHI-F. The DNA sequence reveals 100% identify with the regulatory gene tfdS of pJP4. An XbaI-PstI fragment, containing the complete ORF, encodes a 32,000-Da protein which binds to the promoter regions upstream from tfdA and tfdDII. The deduced amino acid sequence of the complete ORF shows similarity with sequences of activator proteins TcbR, CatM, and CatR of the LysR family. The complete ORF represents the regulatory gene tfdR. The deduced amino acid sequence of the incomplete ORF, situated divergently from tfdR, indicates similarity to chloromuconate cycloisomerases produced by genes tfdD and tcbD of plasmids pJP4 and pP51, respectively. This ORF is identified as part of a putative isofunctional gene, tfdDII.  相似文献   

13.
The broad host range plasmid pJP4, which carries genes for the degradation of 2,4-dichlorophenoxyacetic acid (2,4-D), 2-methyl-4-chlorophenoxyacetic acid, and 3-chlorobenzoic acid, was used in conjugation experiments with mixed cultures enriched from water and sediment samples from an alkaline pond in the area of Szegedi Fehértó, a soda lake in south Hungary. pJP4-encoded mercury resistance was used as a selection marker. One of the transconjugants, the alkaliphilic, moderately halophilic strain EF43, stably maintained the plasmid and was able to degrade 2,4-D and 3-chlorobenzoate under alkaline conditions in the presence of an additional carbon source such as pyruvate, benzoate, or alpha-ketoglutarate, indicating that the degradative genes of pJP4 were expressed in this strain. However, it was unable to grow on these chloroaromatic substrates when the substrate was the sole source of carbon and energy. Chemostat cultivation experiments revealed that the 2,4-D degradation rate during growth on benzoate or pyruvate was limited by the low activity of chlorocatechol-degrading enzymes, particularly chloromuconate cycloisomerase. Strain EF43 was identified as Halomonas sp. on the basis of 16S rRNA sequencing and additional taxonomic studies. 16S rRNA sequence analysis revealed that strain EF43 is closely related to typical soda lake isolates belonging to the genus Halomonas.  相似文献   

14.
Growth of Alcaligenes eutrophus JMP134 on 2,4-dichlorophenoxyacetate requires a 2,4-dichlorphenol hydroxylase encoded by gene tfdB. Catabolism of either 2,4-dichlorophenoxyacetate or 3-chlorobenzoate involves enzymes encoded by the chlorocatechol oxidative operon consisting of tfdCDEF, which converts 3-chloro- and 3,5-dichlorocatechol to maleylacetate and chloromaleylacetate, respectively. Transposon mutagenesis has localized tfdB and tfdCDEF to EcoRI fragment B of plasmid pJP4 (R. H. Don, A. J. Wieghtman, H.-J. Knackmuss, and K. N. Timmis, J. Bacteriol. 161:85-90, 1985). We present the complete nucleotide sequence of tfdB and tfdCDEF contained within a 7,954-base-pair HindIII-SstI fragment from EcoRI fragment B. Sequence and expression analysis of tfdB in Escherichia coli suggested that 2,4-dichlorophenol hydroxylase consists of a single subunit of 65 kilodaltons. The amino acid sequences of proteins encoded by tfdD and tfdE were found to be 63 and 53% identical to those of functionally similar enzymes encoded by clcB and clcD, respectively, from plasmid pAC27 of Pseudomonas putida. P. putida(pAC27) can utilize 3-chlorocatechol but not dichlorinated catechols. A region of DNA adjacent to clcD in pAC27 was found to be 47% identical in amino acid sequence to tfdF, a gene important in catabolizing dichlorocatechols. The region in pAC27 does not appear to encode a protein, suggesting that the absence of a functional trans-chlorodienelactone isomerase may prevent P. putida(pAC27) from utilizing 3,5-dichlorocatechol.  相似文献   

15.
Summary The 2,4-dichlorophenoxyacetate (2,4-D) catabolic plasmid pJP4 of Alcaligenes eutrophus JMP134 contains two sets of nonidentical chlorocatechol oxidation gene sequences physically separated by a 7 kb DNA region. We determined the nucleotide sequence of the 1.6 kb HindIII fragment containing the known genes tfdC and tfdD (Don et al. 1985) which encode pyrocatechase and cycloisomerase, respectively. The 1.3 kb BglII-HindIII segment of recombinant plasmid pDC25 containing at least three chlorocatechol (clc) oxidation genes of the pAC27 plasmid in Pseudomonas putida AC868 (Ghosal et al. 1985a; Frantz and Chakrabarty 1986), was also sequenced. When the tfdC gene of the pJP4 plasmid was compared with gene clcA of plasmid pAC27, which encodes the chlorocatechol specific pyrocatechase (pyrocatechase II), the two genes showed 63% nucleotide sequence homology with 60% homology in their amino acid sequences. In both plasmid pJP4 and pAC27, the two genes encoding the pyrocatechase and the cycloisomerase showed a 4 bp overlap spanning the initiation codon of the cycloisomerase gene and the termination codon of the pyrocatechase gene. The sizes of the polypeptides encoded by the isofunctional genes tfdC and clcA are very similar and thus reflect their functional homology.  相似文献   

16.
The 4-chloro- and 2,4-dichlorophenol-degrading strain Rhodococcus opacus 1CP has previously been shown to acquire, during prolonged adaptation, the ability to mineralize 2-chlorophenol. In addition, homogeneous chlorocatechol 1,2-dioxygenase from 2-chlorophenol-grown biomass has shown relatively high activity towards 3-chlorocatechol. Based on sequences of the N terminus and tryptic peptides of this enzyme, degenerate PCR primers were now designed and used for cloning of the respective gene from genomic DNA of strain 1CP. A 9.5-kb fragment containing nine open reading frames was obtained on pROP1. Besides other genes, a gene cluster consisting of four chlorocatechol catabolic genes was identified. As judged by sequence similarity and correspondence of predicted N termini with those of purified enzymes, the open reading frames correspond to genes for a second chlorocatechol 1,2-dioxygenase (ClcA2), a second chloromuconate cycloisomerase (ClcB2), a second dienelactone hydrolase (ClcD2), and a muconolactone isomerase-related enzyme (ClcF). All enzymes of this new cluster are only distantly related to the known chlorocatechol enzymes and appear to represent new evolutionary lines of these activities. UV overlay spectra as well as high-pressure liquid chromatography analyses confirmed that 2-chloro-cis,cis-muconate is transformed by ClcB2 to 5-chloromuconolactone, which during turnover by ClcF gives cis-dienelactone as the sole product. cis-Dienelactone was further hydrolyzed by ClcD2 to maleylacetate. ClcF, despite its sequence similarity to muconolactone isomerases, no longer showed muconolactone-isomerizing activity and thus represents an enzyme dedicated to its new function as a 5-chloromuconolactone dehalogenase. Thus, during 3-chlorocatechol degradation by R. opacus 1CP, dechlorination is catalyzed by a muconolactone isomerase-related enzyme rather than by a specialized chloromuconate cycloisomerase.  相似文献   

17.
The biochemical characterization of the muconate and the chloromuconate cycloisomerases of the chlorophenol-utilizing Rhodococcus erythropolis strain 1CP previously indicated that efficient chloromuconate conversion among the gram-positive bacteria might have evolved independently of that among gram-negative bacteria. Based on sequences of the N terminus and of tryptic peptides of the muconate cycloisomerase, a fragment of the corresponding gene has now been amplified and used as a probe for the cloning of catechol catabolic genes from R. erythropolis. The clone thus obtained expressed catechol 1,2-dioxygenase, muconate cycloisomerase, and muconolactone isomerase activities. Sequencing of the insert on the recombinant plasmid pRER1 revealed that the genes are transcribed in the order catA catB catC. Open reading frames downstream of catC may have a function in carbohydrate metabolism. The predicted protein sequence of the catechol 1,2-dioxygenase was identical to the one from Arthrobacter sp. strain mA3 in 59% of the positions. The chlorocatechol 1,2-dioxygenases and the chloromuconate cycloisomerases of gram-negative bacteria appear to be more closely related to the catechol 1,2-dioxygenases and muconate cycloisomerases of the gram-positive strains than to the corresponding enzymes of gram-negative bacteria.  相似文献   

18.
Plasmid pJP4 permits its host bacterium, strain JMP134, to degrade and utilize as sole sources of carbon and energy 3-chlorobenzoate and 2,4-dichlorophenoxyacetic acid (R. H. Don and J. M. Pemberton, J. Bacteriol. 145:681-686, 1981). Mutagenesis of pJP4 by transposons Tn5 and Tn1771 enabled localization of five genes for enzymes involved in these catabolic pathways. Four of the genes, tfdB, tfdC, tfdD, and tfdE, encoded 2,4-dichlorophenol hydroxylase, dichlorocatechol 1,2-dioxygenase, chloromuconate cycloisomerase, and chlorodienelactone hydrolase, respectively. No function has been assigned to the fifth gene, tfdF, although it may encode a trans-chlorodiene-lactone isomerase. Inactivation of genes tfdC, tfdD, and tfdE, which encode the transformation of dichlorocatechol to chloromaleylacetic acid, prevented host strain JMP134 from degrading both 3-chlorobenzoate and 2,4-dichlorophenoxyacetic acid, which indicates that the pathways for these two substrates utilize common enzymes for the dissimilation of chlorocatechols. Studies with cloned catabolic genes from pJP4 indicated that whereas all essential steps in the degradation of 2,4-dichlorophenoxyacetic acid are plasmid encoded, the conversion of 3-chlorobenzoate to chlorocatechol is specified by chromosomal genes.  相似文献   

19.
Chloromuconolactone dehalogenase ClcF plays a unique role in 3-chlorocatechol degradation by Rhodococcus opacus 1CP by compensating the inability of its chloromuconate cycloisomerase ClcB2 to dechlorinate the chemically stable cycloisomerization product (4R,5S)-5-chloromuconolactone (5CML). High sequence similarities showed relatedness of ClcF to muconolactone isomerases (MLIs, EC 5.3.3.4) of the 3-oxoadipate pathway. Although both enzyme types share the ability to dechlorinate 5CML, comparison of kcat/Km indicated a significant extent of specialization of ClcF for dechlorination. This assumption was substantiated by an almost complete inability of ClcF to convert (4S)-muconolactone and the exclusive formation of cis-dienelactone from 5CML. Mutational analysis of ClcF by means of variants E27D, E27Q, Y50A, N52A, and A89S indicated relevance of some highly conserved residues for substrate binding and catalysis. Based on the putative isomerization mechanism of MLI, evidence was provided for a role of E27 in initial proton abstraction as well as of Y50 and N52 in substrate binding. In case of N52 substrate binding is likely to occur to the carboxylic group of 5CML as indicated by a significant change of product specificity. Expression in Escherichia coli BL21-CP(DE)-RIL followed by a three-step purification procedure with heat treatment is a convenient strategy to obtain recombinant ClcF and variants thereof.  相似文献   

20.
The enzymes chlorocatechol-1,2-dioxygenase, chloromuconate cycloisomerase, dienelactone hydrolase, and maleylacetate reductase allow Ralstonia eutropha JMP134(pJP4) to degrade chlorocatechols formed during growth in 2,4-dichlorophenoxyacetate or 3-chlorobenzoate (3-CB). There are two gene modules located in plasmid pJP4, tfdC(I)D(I)E(I)F(I) (module I) and tfdD(II)C(II)E(II)F(II) (module II), putatively encoding these enzymes. To assess the role of both tfd modules in the degradation of chloroaromatics, each module was cloned into the medium-copy-number plasmid vector pBBR1MCS-2 under the control of the tfdR regulatory gene. These constructs were introduced into R. eutropha JMP222 (a JMP134 derivative lacking pJP4) and Pseudomonas putida KT2442, two strains able to transform 3-CB into chlorocatechols. Specific activities in cell extracts of chlorocatechol-1,2-dioxygenase (tfdC), chloromuconate cycloisomerase (tfdD), and dienelactone hydrolase (tfdE) were 2 to 50 times higher for microorganisms containing module I compared to those containing module II. In contrast, a significantly (50-fold) higher activity of maleylacetate reductase (tfdF) was observed in cell extracts of microorganisms containing module II compared to module I. The R. eutropha JMP222 derivative containing tfdR-tfdC(I)D(I)E(I)F(I) grew four times faster in liquid cultures with 3-CB as a sole carbon and energy source than in cultures containing tfdR-tfdD(II)C(II)E(II)F(II). In the case of P. putida KT2442, only the derivative containing module I was able to grow in liquid cultures of 3-CB. These results indicate that efficient degradation of 3-CB by R. eutropha JMP134(pJP4) requires the two tfd modules such that TfdCDE is likely supplied primarily by module I, while TfdF is likely supplied by module II.  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