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 15 毫秒
1.
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   

2.
Biochemical investigations of the muconate and chloromuconate cycloisomerases from the chlorophenol-utilizing strain Rhodococcus opacus (erythropolis) 1CP had previously indicated that the chlorocatechol catabolic pathway of this strain may have developed independently from the corresponding pathways of proteobacteria. To test this hypothesis, we cloned the chlorocatechol catabolic gene cluster of strain 1CP by using PCR with primers derived from sequences of N termini and peptides of purified chlorocatechol 1,2-dioxygenase and chloromuconate cycloisomerase. Sequencing of the clones revealed that they comprise different parts of the same gene cluster in which five open reading frames have been identified. The clcB gene for chloromuconate cycloisomerase is transcribed divergently from a gene which codes for a LysR-type regulatory protein, the presumed ClcR. Downstream of clcR but separated from it by 222 bp, we detected the clcA and clcD genes, which could unambiguously be assigned to chlorocatechol 1,2-dioxygenase and dienelactone hydrolase. A gene coding for a maleylacetate reductase could not be detected. Instead, the product encoded by the fifth open reading frame turned out to be homologous to transposition-related proteins of IS1031 and Tn4811. Sequence comparisons of ClcA and ClcB to other 1,2-dioxygenases and cycloisomerases, respectively, clearly showed that the chlorocatechol catabolic enzymes of R. opacus 1CP represent different branches in the dendrograms than their proteobacterial counterparts. Thus, while the sequences diverged, the functional adaptation to efficient chlorocatechol metabolization occurred independently in proteobacteria and gram-positive bacteria, that is, by functionally convergent evolution.  相似文献   

3.
Xanthobacter flavus 14p1 used 1,4-dichlorobenzene as the sole source of carbon and energy but did not grow on other (chloro)aromatic compounds. 1,4-Dichlorobenzene was attacked by a chlorobenzene dioxygenase, and the intermediate chlorocatechol was metabolized by the modified ortho pathway. All enzymes necessary to convert 1,4-dichlorobenzene to 3-oxoadipate showed a low substrate specificity and also accepted the respective intermediates of chlorobenzene or 1,3-dichlorobenzene degradation. Of the three compounds chlorobenzene, 1,4-dichlorobenzene, and 1,3-dichlorobenzene, the latter was the most toxic for X. flavus 14p1. Furthermore, 1,3-dichlorobenzene did not induce chlorocatechol 1,2-dioxygenase activity of the organism. Chlorobenzene, however, induced chlorocatechol 1,2-dioxygenase, dienelactone hydrolase, and maleylacetate reductase activities. As demonstrated by chloride release, also chlorobenzene dioxygenase, chlorobenzene cis-dihydrodiol dehydrogenase, and chloromuconate cycloisomerase activities were present in chlorobenzene-induced cells, but chlorobenzene failed to support growth. Presumably a toxic compound was formed from one of the intermediates. Received: 10 June 1996 / Revision received: 23 December 1996 / Accepted: 18 January 1997  相似文献   

4.
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   

5.
Pseudomonas reinekei MT1 has previously been reported to degrade 4- and 5-chlorosalicylate by a pathway with 4-chlorocatechol, 3-chloromuconate, 4-chloromuconolactone, and maleylacetate as intermediates, and a gene cluster channeling various salicylates into an intradiol cleavage route has been reported. We now report that during growth on 5-chlorosalicylate, besides a novel (chloro)catechol 1,2-dioxygenase, C12OccaA, a novel (chloro)muconate cycloisomerase, MCIccaB, which showed features not yet reported, was induced. This cycloisomerase, which was practically inactive with muconate, evolved for the turnover of 3-substituted muconates and transforms 3-chloromuconate into equal amounts of cis-dienelactone and protoanemonin, suggesting that it is a functional intermediate between chloromuconate cycloisomerases and muconate cycloisomerases. The corresponding genes, ccaA (C12OccaA) and ccaB (MCIccaB), were located in a 5.1-kb genomic region clustered with genes encoding trans-dienelactone hydrolase (ccaC) and maleylacetate reductase (ccaD) and a putative regulatory gene, ccaR, homologous to regulators of the IclR-type family. Thus, this region includes genes sufficient to enable MT1 to transform 4-chlorocatechol to 3-oxoadipate. Phylogenetic analysis showed that C12OccaA and MCIccaB are only distantly related to previously described catechol 1,2-dioxygenases and muconate cycloisomerases. Kinetic analysis indicated that MCIccaB and the previously identified C12OsalD, rather than C12OccaA, are crucial for 5-chlorosalicylate degradation. Thus, MT1 uses enzymes encoded by a completely novel gene cluster for degradation of chlorosalicylates, which, together with a gene cluster encoding enzymes for channeling salicylates into the ortho-cleavage pathway, form an effective pathway for 4- and 5-chlorosalicylate mineralization.The aerobic degradation of chloroaromatic compounds usually proceeds via chlorocatechols as central intermediates (20, 47), which in most of the cases reported thus far, are further degraded by enzymes of the chlorocatechol pathway (44). This pathway involves ortho-cleavage by a chlorocatechol 1,2-dioxygenase with high activity for chlorocatechols (12), a chloromuconate cycloisomerase with high activity for chloromuconates (54), a dienelactone hydrolase active with both cis- and trans-dienelactone (4-carboxymethylenebut-2-en-4-olide) (54), and a maleylacetate reductase (MAR) (28).However, it has become evident in recent years that microorganisms have evolved various alternative strategies to mineralize chlorocatechols. Pseudomonas putida GJ31 was found to degrade chlorobenzene rapidly via 3-chlorocatechol using a catechol meta-cleavage pathway (33). Two alternative pathways for 3- and 4-chlorocatechol degradation that involve reactions known from the chlorocatechol, as well as the 3-oxoadipate, pathway have recently been observed in Rhodococcus opacus 1CP (35) and Pseudomonas reinekei MT1 (39). In R. opacus 1CP, 3-chloro- and 2,4-dichloro-cis,cis-muconate (the ring cleavage products of 4-chlorocatechol and 3,5-dichlorocatechol, respectively) are converted to the respective cis-dienelactones (35, 58), similar to the reaction described for proteobacterial chloromuconate cycloisomerases (54). However, proteobacterial chloromuconate cycloisomerase can dehalogenate 2-chloromuconate (the ring cleavage product of 3-chlorocatechol) and transform this compound via 5-chloromuconolactone into trans-dienelactone (54, 65), whereas none of the described chloromuconate cycloisomerases of R. opacus 1CP can catalyze such a dehalogenation, and 5-chloromuconolactone is the product of the cycloisomerization reaction (35, 58). Dehalogenation is achieved by an enzyme with high sequence similarity to muconolactone isomerases (35), which in proteobacteria have been shown to be capable of dehalogenating 5-chloromuconolactone to cis-dienelactone (46).In P. reinekei MT1, a trans-dienelactone hydrolase (trans-DLH) was identified as the key enzyme involved in the degradation of 4- and 5-chlorosalicylate via 4-chlorocatechol as an intermediate (39). In contrast to all previously described dienelactone hydrolases involved in chlorocatechol degradation, which belong to the α/β hydrolase fold enzymes with a catalytic triad consisting of Cys, His, and Asp (10), trans-DLH was shown to be a zinc-dependent hydrolase (8). The function of this enzyme in the 4-chlorocatechol metabolic pathway was to interact with the muconate cycloisomerase (MCI)-mediated transformation of 3-chloromuconate into protoanemonin. By acting on the reaction intermediate 4-chloromuconolactone, trans-DLH prevents the formation of protoanemonin by catalyzing its hydrolysis to maleylacetate (39). Maleylacetate, in turn, is reduced by MAR to 3-oxoadipate.A more detailed genetic and biochemical analysis of the degradation of differently substituted salicylates (7) had shown the presence of two catabolic gene clusters in MT1. An archetype catRBCA gene cluster was shown to be involved in salicylate degradation. The second gene cluster (sal) had a novel gene arrangement, with salA, encoding a salicylate 1-hydroxylase, clustered with the salCD genes, encoding MCI and catechol 1,2-dioxygenase (C12O), respectively. As these genes were expressed during growth on differently substituted salicylates, it was proposed that the function of the sal gene cluster is to channel both chlorosubstituted and methylsubstituted salicylates into a catechol ortho-cleavage pathway, followed by dismantling of the formed substituted muconolactones through specific pathways. However, previous analyses had indicated the presence of an additional and thus third (chloro)muconate cycloisomerase in MT1 during growth on chlorosalicylate, which is distinct from both previously described MCIs encoded by the cat cluster (MCIcatB) and the sal cluster (MCIsalC), as it transforms 3-chloromuconate into approximately equal amounts of cis-dienelactone and protoanemonin (39). In the present report, this cycloisomerase is biochemically and genetically described and shown to be located in a third gene cluster involved in the degradation of 5-chlorosalicylate by strain MT1. This cluster comprises genes encoding a third C12O, trans-DLH (8), and a MAR. Evidently, P. reinekei MT1 is the first microorganism in which such a complex net of genes involved in chlorocatechol degradation has been described.  相似文献   

6.
Alcaligenes eutrophus CH34 used benzoate as a sole source of carbon and energy, degrading it through the 3-oxoadipate pathway. All the enzymes required for this degradation were shown to be encoded by chromosomal genes. Catechol 1,2-dioxygenase activity was induced by benzoate, catechol, 4-chlorocatechol, and muconate. The enzyme is most likely a homodimer, with an apparent molecular weight of 76,000 ± 500. According to several criteria, its properties are intermediate between those of catechol 1,2-dioxygenases (CatA) and chlorocatechol 1,2-dioxygenases (ClcA). The determined K m for catechol is the lowest among known catechol and chlorocatechol dioxygenases. Similar K m values were found for para-substituted catechols, although the catalytic constants were much lower. The catechol 1,2-dioxygenase from strain CH34 is unique in its property to transform tetrachlorocatechol; however, excess substrate led to a marked reversible inhibition. Some meta- and multi-substituted catechols behaved similarly. The determined K m (or K i) values for para- or meta-substituted catechols suggest that the presence of an electron-withdrawing substituent at one of these positions results in a higher affinity of the enzyme for the ligand. Results of studies of recognition by the enzyme of various nonmetabolised aromatic compounds are also discussed. Received: 20 November 1996 / Accepted: 11 April 1996  相似文献   

7.
The catechol and protocatechuate branches of the 3-oxoadipate pathway, which are important for the bacterial degradation of aromatic compounds, converge at the common intermediate 3-oxoadipate enol-lactone. A 3-oxoadipate enol-lactone-hydrolyzing enzyme, purified from benzoate-grown cells of Rhodococcus opacus (erythropolis) 1CP, was found to have a larger molecular mass under denaturing conditions than the corresponding enzymes previously purified from γ-proteobacteria. Sequencing of the N terminus and of tryptic peptides allowed cloning of the gene coding for the 3-oxoadipate enol-lactone hydrolase by using PCR with degenerate primers. Sequencing showed that the gene belongs to a protocatechuate catabolic gene cluster. Most interestingly, the hydrolase gene, usually termed pcaD, was fused to a second gene, usually termed pcaC, which encodes the enzyme catalyzing the preceding reaction, i.e., 4-carboxymuconolactone decarboxylase. The two enzymatic activities could not be separated chromatographically. At least six genes of protocatechuate catabolism appear to be transcribed in the same direction and in the following order: pcaH and pcaG, coding for the subunits of protocatechuate 3,4-dioxygenase, as shown by N-terminal sequencing of the subunits of the purified protein; a gene termed pcaB due to the homology of its gene product to 3-carboxy-cis,cis-muconate cycloisomerases; pcaL, the fused gene coding for PcaD and PcaC activities; pcaR, presumably coding for a regulator of the IclR-family; and a gene designated pcaF because its product resembles 3-oxoadipyl coenzyme A (3-oxoadipyl-CoA) thiolases. The presumed pcaI, coding for a subunit of succinyl-CoA:3-oxoadipate CoA-transferase, was found to be transcribed divergently from pcaH.  相似文献   

8.
Of various benzoate-utilizing bacteria tested, Alcaligenes eutrophus 335, A. eutrophus H16, A. eutrophus JMP222, A. eutrophus JMP134, Alcaligenes strain A7, and Pseudomonas cepacia were able to grow with 4-fluorobenzoate as the sole source of carbon and energy. P. cepacia also utilizes 3-fluorobenzoate. Except for A. eutrophus JMP134, which is known to grow with 2,4-dichlorophenoxyacetate and 3-chlorobenzoate (R. H. Don and J. M. Pemberton, J. Bacteriol. 145:681-686, 1981), the strains were unable to grow at the expense of these compounds or 4-chlorobenzoate. Assays of cell extracts revealed that all strains express dienelactone hydrolase and maleylacetate reductase activities in addition to enzymes of the catechol branch of the 3-oxoadipate pathway when growing with 4-fluorobenzoate. Induction of dienelactone hydrolase and maleylacetate reductase apparently is not necessarily connected to synthesis of catechol 1,2-dioxygenase type II and chloromuconate cycloisomerase activities, which are indispensable for the degradation of chlorocatechols. Substrate specificities of the dienelactone hydrolases provisionally differentiate among three types of this activity. (i) Extracts of A. eutrophus 335, A. eutrophus H16, A. eutrophus JMP222, and Alcaligenes strain A7 convert trans-4-carboxymethylenebut-2-en-4-olide (trans-dienelactone) much faster than the cis-isomer (type I). (ii) The enzyme present in P. cepacia shows the opposite preference for the isomeric substrates (type II). (iii) Cell extracts of A. eutrophus JMP134, as well as purified dienelactone hydrolase from Pseudomonas strain B13 (E. Schmidt and H.-J. Knackmuss, Biochem. J. 192:339-347, 1980), hydrolyze both dienelactones at rates that are of the same order of magnitude (type III). This classification implies that A. eutrophus JMP134 possesses at least two different dienelactone hydrolases, one of type III encoded by the plasmid pJP4 and one of type I, which is also present in the cured strain JMP222.  相似文献   

9.
The genes responsible for the degradation of 2,4-dichlorophenoxyacetate (2,4-D) by -Proteobacteria have previously been difficult to detect by using gene probes or polymerase chain reaction (PCR) primers. PCR products of the chlorocatechol 1,2-dioxygenase gene, tfdC, now allowed cloning of two chlorocatechol gene clusters from the Sphingomonas sp. strain TFD44. Sequence characterization showed that the first cluster, tfdD,RFCE, comprises all the genes necessary for the conversion of 3,5-dichlorocatechol to 3-oxoadipate, including a presumed regulatory gene, tfdR, of the LysR-type family. The second gene cluster, tfdC2E2F2, is incomplete and appears to lack a chloromuconate cycloisomerase gene and a regulatory gene. Purification and N-terminal sequencing of selected enzymes suggests that at least representatives of both gene clusters (TfdD of cluster 1 and TfdC2 of cluster 2) are induced during the growth of strain TFD44 with 2,4-D. A mutant constructed to contain an insertion in the chloromuconate cycloisomerase gene tfdD still was able to grow with 2,4-D, but more slowly and with a longer lag phase. This, and the detection of additional activity peaks during protein purification suggest that strain TFD44 harbors at least another chloromuconate cycloisomerase gene. The sequence of the tfdCE region was almost identical to that of a partially characterized chlorocatechol catabolic gene cluster of Sphingomonas herbicidovorans MH, whereas the sequence of the tfdC2E2F2 cluster was different. The similarity of the predicted proteins of the tfdD,RFCE and tfdC2E2F2 clusters to known sequences of other Proteobacteria in the database ranged from 42 to 61% identical positions for the first cluster and from 45.5 to 58% identical positions for the second cluster. Between both clusters, the similarities of their predicted proteins ranged from 44.5 to 64% identical positions. Thus, both clusters (together with those of S. herbicidovorans MH) represent deep-branching lines in the respective dendrograms, and the sequence information will help future primer design for the detection of corresponding genes in the environment.  相似文献   

10.
Degradation of para-toluate by Rhodococcus opacus 1cp was investigated. Activities of the key enzymes of this process, catechol 1,2-dioxygenase and muconate cycloisomerase, are detected in this microorganism. Growth on p-toluate was accompanied by induction of two catechol 1,2-dioxygenases. The substrate specificity and physicochemical properties of one enzyme are identical to those of chlorocatechol 1,2-dioxygenase; induction of the latter enzyme was observed during R. opacus 1cp growth on 4-chlorophenol. The other enzyme isolated from the biomass grown on p-toluate exhibited lower rate of chlorinated substrate cleavage compared to the catechol substrate. However, this enzyme is not identical to the catechol 1,2-dioxygenase cloned in this strain within the benzoate catabolism operon. This supports the hypothesis on the existence of multiple forms of dioxygenases as adaptive reactions of microorganisms in response to environmental stress.  相似文献   

11.
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   

12.
Dienelactone hydrolase from Pseudomonas sp. strain B13.   总被引:6,自引:5,他引:1       下载免费PDF全文
Dienelactone hydrolase (EC 3.1.1.45) catalyzes the conversion of cis- or trans-4-carboxymethylenebut-2-en-4-olide (dienelactone) to maleylacetate. An approximately 24-fold purification from extracts of 3-chlorobenzoate-grown Pseudomonas sp. strain B13 yielded a homogeneous preparation of the enzyme. The purified enzyme crystallized readily and proved to be a monomer with a molecular weight of about 30,000. Each dienelactone hydrolase molecule contains two cysteinyl side chains. One of these was readily titrated by stoichiometric amounts of p-chloromercuribenzoate, resulting in inactivation of the enzyme; the inactivation could be reversed by the addition of dithiothreitol. The other cysteinyl side chain appeared to be protected in the native protein against chemical reaction with p-chloromercuribenzoate. The properties of sulfhydryl side chains in dienelactone hydrolase resembled those that have been characterized for bacterial 4-carboxymethylbut-3-en-4-olide (enol-lactone) hydrolases (EC 3.1.1.24), which also are monomers with molecular weights of about 30,000. The amino acid composition of the dienelactone hydrolase resembled the amino acid composition of enol-lactone hydrolase from Pseudomonas putida, and alignment of the NH2-terminal amino acid sequence of the dienelactone hydrolase with the corresponding sequence of an Acinetobacter calcoaceticus enol-lactone hydrolase revealed sequence identity at 8 of the 28 positions. These observations foster the hypothesis that the lactone hydrolases share a common ancestor. The lactone hydrolases differed in one significant property: the kcat of dienelactone hydrolase was 1,800 min-1, an order of magnitude below the kcat observed with enol-lactone hydrolases. The relatively low catalytic activity of dienelactone hydrolase may demand its production at the high levels observed for induced cultures of Pseudomonas sp. strain B13.  相似文献   

13.
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   

14.
The nucleotide sequence of a 10,528-bp region comprising the chlorocatechol pathway gene cluster tetRtetCDEF of the 1,2,3,4-tetrachlorobenzene via the tetrachlorocatechol-mineralizing bacterium Pseudomonas chlororaphis RW71 (T. Potrawfke, K. N. Timmis, and R.-M. Wittich, Appl. Environ. Microbiol. 64:3798-3806, 1998) was analyzed. The chlorocatechol 1,2-dioxygenase gene tetC was cloned and overexpressed in Escherichia coli. The recombinant gene product was purified, and the alpha,alpha-homodimeric TetC was characterized. Electron paramagnetic resonance measurements confirmed the presence of a high-spin-state Fe(III) atom per monomer in the holoprotein. The productive transformation by purified TetC of chlorocatechols bearing chlorine atoms in positions 4 and 5 provided strong evidence for a significantly broadened substrate spectrum of this dioxygenase compared with other chlorocatechol dioxygenases. The conversion of 4,5-dichloro- or tetrachlorocatechol, in the presence of catechol, displayed strong competitive inhibition of catechol turnover. 3-Chlorocatechol, however, was simultaneously transformed, with a rate similar to that of the 4,5-halogenated catechols, indicating similar specificity constants. These novel characteristics of TetC thus differ significantly from results obtained from hitherto analyzed catechol 1,2-dioxygenases and chlorocatechol 1,2-dioxygenases.  相似文献   

15.
Maleylacetate reductases (EC 1.3.1.32) have been shown to contribute not only to the bacterial catabolism of some usual aromatic compounds like quinol or resorcinol but also to the degradation of aromatic compounds carrying unusual substituents, such as halogen atoms or nitro groups. Genes coding for maleylacetate reductases so far have been analyzed mainly in chloroaromatic compound-utilizing proteobacteria, in which they were found to belong to specialized gene clusters for the turnover of chlorocatechols or 5-chlorohydroxyquinol. We have now cloned the gene macA, which codes for one of apparently (at least) two maleylacetate reductases in the gram-positive, chlorophenol-degrading strain Rhodococcus opacus 1CP. Sequencing of macA showed the gene product to be relatively distantly related to its proteobacterial counterparts (ca. 42 to 44% identical positions). Nevertheless, like the known enzymes from proteobacteria, the cloned Rhodococcus maleylacetate reductase was able to convert 2-chloromaleylacetate, an intermediate in the degradation of dichloroaromatic compounds, relatively fast and with reductive dehalogenation to maleylacetate. Among the genes ca. 3 kb up- and downstream of macA, none was found to code for an intradiol dioxygenase, a cycloisomerase, or a dienelactone hydrolase. Instead, the only gene which is likely to be cotranscribed with macA encodes a protein of the short-chain dehydrogenase/reductase family. Thus, the R. opacus maleylacetate reductase gene macA clearly is not part of a specialized chlorocatechol gene cluster.Maleylacetate reductases (EC 1.3.1.32) have long been known to be involved in the degradation of chloroaromatic compounds via chlorocatechols as intermediates (10, 31). By reduction of a carbon-carbon double bond they form 3-oxoadipate, a metabolite also of catechol catabolism, and thus compensate for the different oxidation states of chlorinated and nonchlorinated compounds. 2-Chloromaleylacetate, which is formed during turnover of several dichlorocatechols, is initially reductively dechlorinated and then reduced to 3-oxoadipate in a second reaction (22, 47).Corresponding to the biochemical function in chlorocatechol degradation, the following maleylacetate reductase genes have been shown to be associated with dioxygenase, cycloisomerase, and dienelactone hydrolase genes as components of specialized chlorocatechol catabolic operons: tfdF and tfdFII on pJP4 from the 2,4-dichlorophenoxyacetate-utilizing strain Ralstonia eutropha (Alcaligenes eutrophus) JMP134 (29, 33, 37, 44), tcbF on pP51 from the 1,2,4-trichlorobenzene-degrading strain Pseudomonas sp. strain P51 (45), and clcE from the 3-chlorobenzoate catabolizing strains Pseudomonas sp. strain B13 and Pseudomonas putida AC866(pAC27) (15, 20, 21). Catechol degradation, in contrast, does not require a maleylacetate reductase activity, and corresponding genes do not belong to the known catechol operons. Thus, while at least two of the chlorocatechol catabolic enzymes, i.e., the dioxygenases and cycloisomerases, appear to have been recruited from catechol catabolism, maleylacetate reductase genes must have had a different origin and original function (34).The postulated original function of the maleylacetate reductases is still under discussion. In bacteria, these enzymes have been shown to play a role, for example, in quinol, resorcinol, and 2,4-dihydroxybenzoate degradation (6, 25, 41). Other aromatic growth substrates involving the action of maleylacetate reductase are more exotic, since they carry a fluorine substituent (35), a sulfo group (14), a nitro group (18, 40), or several chlorine substituents (8, 26, 48). Maleylacetate reductase genes have been shown to be part of a specialized gene cluster for 2,4,5-trichlorophenoxyacetate degradation (8, 9) and of a gene cluster for hydroxyquinol conversion which contributes to 4-nitrophenol turnover (4).The chlorocatechol pathway of the chlorophenol-utilizing strain Rhodococcus opacus (erythropolis) 1CP obviously evolved functionally convergent to the corresponding pathway in the proteobacteria mentioned above (13, 39). Thus, it is not surprising that the chlorocatechol gene cluster of strain 1CP is organized differently from the corresponding proteobacterial operons; in fact, its characterization showed that it does not comprise a maleylacetate reductase gene (13). Thus, the nature of the gene cluster(s) encoding a maleylacetate reductase in R. opacus remained to be elucidated. Such gene clusters could complement otherwise incomplete pathways, and they might also have provided the source from which the maleylacetate reductase gene was recruited during evolution of dedicated pathways, such as the proteobacterial chlorocatechol catabolic route.(Some of the results presented here have previously been reported in a preliminary communication [38].)  相似文献   

16.
During cultivation in a liquid medium, the bacterium Rhodococcus opacus 1G was capable of growing on phenol at a concentration of up to 0.75 g/l. Immobilization of Rhodococcus opacus 1G had a positive effect on cell growth in the presence of phenol at high concentrations. The substrate at concentrations of 1.0 and 1.5 g/l was completely utilized over 24 and 48 h, respectively. The key enzymes of phenol degradation (two catechol 1,2-dioxygenases and muconate cycloisomerase) were isolated. One of the dioxygenases was very unstable. By substrate specificity, another enzyme belonged to catechol 1,2-dioxygenases of the classical ortho-pathway. Chlorocatechols and chlorophenols served as competitive inhibitors of catechol 1,2-dioxygenases. The inhibitory effect of other aromatic compounds was less significant. Our results suggest that this strain holds promise for bioremediation of phenol wastewater.  相似文献   

17.
Phenoxyalkanoic acid degradation is well studied in Beta- and Gammaproteobacteria, but the genetic background has not been elucidated so far in Alphaproteobacteria. We report the isolation of several genes involved in dichlor- and mecoprop degradation from the alphaproteobacterium Sphingomonas herbicidovorans MH and propose that the degradation proceeds analogously to that previously reported for 2,4-dichlorophenoxyacetic acid (2,4-D). Two genes for alpha-ketoglutarate-dependent dioxygenases, sdpA(MH) and rdpA(MH), were found, both of which were adjacent to sequences with potential insertion elements. Furthermore, a gene for a dichlorophenol hydroxylase (tfdB), a putative regulatory gene (cadR), two genes for dichlorocatechol 1,2-dioxygenases (dccA(I/II)), two for dienelactone hydrolases (dccD(I/II)), part of a gene for maleylacetate reductase (dccE), and one gene for a potential phenoxyalkanoic acid permease were isolated. In contrast to other 2,4-D degraders, the sdp, rdp, and dcc genes were scattered over the genome and their expression was not tightly regulated. No coherent pattern was derived on the possible origin of the sdp, rdp, and dcc pathway genes. rdpA(MH) was 99% identical to rdpA(MC1), an (R)-dichlorprop/alpha-ketoglutarate dioxygenase from Delftia acidovorans MC1, which is evidence for a recent gene exchange between Alpha- and Betaproteobacteria. Conversely, DccA(I) and DccA(II) did not group within the known chlorocatechol 1,2-dioxygenases, but formed a separate branch in clustering analysis. This suggests a different reservoir and reduced transfer for the genes of the modified ortho-cleavage pathway in Alphaproteobacteria compared with the ones in Beta- and Gammaproteobacteria.  相似文献   

18.
Ralstonia sp. strain PS12 is able to use 2,4-, 2,5-, and 3,4-dichlorotoluene as growth substrates. Dichloromethylcatechols are central intermediates that are formed by TecA tetrachlorobenzene dioxygenase-mediated activation at two adjacent unsubstituted carbon atoms followed by TecB chlorobenzene dihydrodiol dehydrogenase-catalyzed rearomatization and then are channeled into a chlorocatechol ortho cleavage pathway involving a chlorocatechol 1,2-dioxygenase, chloromuconate cycloisomerase, and dienelactone hydrolase. However, completely different metabolic routes were observed for the three dichloromethylcatechols analyzed. Whereas 3,4-dichloro-6-methylcatechol is quantitatively transformed into one dienelactone (5-chloro-2-methyldienelactone) and thus is degraded via a linear pathway, 3,5-dichloro-2-methylmuconate formed from 4,6-dichloro-3-methylcatechol is subject to both 1,4- and 3,6-cycloisomerization and thus is degraded via a branched metabolic route. 3,6-Dichloro-4-methylcatechol, on the first view, is transformed predominantly into one (2-chloro-3-methyl-trans-) dienelactone. In situ (1)H nuclear magnetic resonance analysis revealed the intermediate formation of 2,5-dichloro-4-methylmuconolactone, showing that both 1,4- and 3,6-cycloisomerization occur with this muconate and indicating a degradation of the muconolactone via a reversible cycloisomerization reaction and the dienelactone-forming branch of the pathway. Diastereomeric mixtures of two dichloromethylmuconolactones were prepared chemically to proof such a hypothesis. Chloromuconate cycloisomerase transformed 3,5-dichloro-2-methylmuconolactone into a mixture of 2-chloro-5-methyl-cis- and 3-chloro-2-methyldienelactone, affording evidence for a metabolic route of 3,5-dichloro-2-methylmuconolactone via 3,5-dichloro-2-methylmuconate into 2-chloro-5-methyl-cis-dienelactone. 2,5-Dichloro-3-methylmuconolactone was transformed nearly exclusively into 2-chloro-3-methyl-trans-dienelactone.  相似文献   

19.
In various bacterial strains belonging to the β-subdivision of proteobacteria which are capable of degrading chlorinated monoaromatic compounds, chlorocatechol 1,2-dioxygenase genes were detected by PCR and Southern hybridization. Using PCR primers derived from the conserved sequence motifs of chlorocatechol 1,2-dioxygenase genes tfdC, clcA and tcbC, PCR products of the expected size were obtained with the test strains, but not with negative control strains. The specificity of the PCR products was verified by hybridization using an oligonucleotide probe for an internal sequence motif which is evolutionarily conserved among chlorocatechol 1,2-dioxygenases and some other dioxygenases that catalyze the intradiol aromatic-ring-cleavage. Hybridization with the tfdC PCR product from the 2,4-D degradative plasmid pJP4 under stringent conditions revealed different extents of homology of the chlorocatechol 1,2-dioxygenase genes to the canonical tfdC sequence in the various strains. These findings were confirmed by the nucleotide sequence analysis of the tfdC-specific PCR products. From our results, we conclude that the PCR primer set is more suitable than the hybridization with pJP4-derived gene probes for the detection of diverse chlorocatechol 1,2-dioxygenase genes in proteobacteria.  相似文献   

20.
Enzymatic conversion of 4-fluorocatechol in the simultaneous presence of partially purified preparations of catechol 1,2-dioxygenase from Pseudomonas cepacia and muconate cycloisomerase from Alcaligenes eutrophus 335 yielded a product that was unambiguously identified as (+)-4-fluoromuconolactone [(+)-4-carboxymethyl-4-fluoro-but-2-en-4-olide]. This compound was shown to be the only major product formed from 3-fluoro-cis,cis-muconate by the action of muconate cycloisomerases from A. eutrophus 335, A. eutrophus JMP134, and P. cepacia as well as by the action of dichloromuconate cycloisomerase from A. eutrophus JMP134. This finding implies that dichloromuconate cycloisomerase, like the muconate cycloisomerases, catalyzes primarily a cycloisomerization reaction, which only in the case of chloro- and bromo-substituted substrates is connected to a dehalogenation. 4-Fluoromuconolactone at pH 7 decomposes by spontaneous reactions mainly to maleylacetate, which then decarboxylates to give cis-acetylacrylate. Although significant amounts of an unidentified compound are also formed from the fluorolactone, HF elimination to the two isomeric dienelactones (4-carboxymethylenebut-2-en-4-olides) is negligible. However, all spontaneous reactions proceed so slowly that an enzymatic conversion of 4-fluoromuconolactone must be assumed. Participation of dienelactone hydrolases in this reaction is indicated by their induction during growth of various strains with 4-fluorobenzoate. However, experiments with cell extracts of P. putida A3.12 suggest that at least one other hydrolytic enzyme is able to contribute to 4-fluoromuconolactone conversion. In light of these observations, earlier proposals for a 4-fluorobenzoate degradative pathway are discussed.  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