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 93 毫秒
1.
The degradation of 4-chlorobenzoate (4-CBA) by Pseudomonas sp. strain CBS3 is thought to proceed first by the dehalogenation of 4-CBA to 4-hydroxybenzoate (4-HBA), which is then metabolized following the protocatechuate branch of the beta-ketoadipate pathway. The cloning of the 4-CBA dehalogenation system was carried out by constructing a gene bank of Pseudomonas sp. strain CBS3 in Pseudomonas putida. Hybrid plasmid pPSA843 contains a 9.5-kilobase-pair fragment derived from the chromosome of Pseudomonas sp. strain CBS3. This plasmid confers on P. putida the ability to dehalogenate 4-CBA and grow on 4-CBA as the only source of carbon. However, pPSA843 did not complement mutants of P. putida unable to grow on 4-HBA (POB-), showing that the genes involved in the metabolism of 4-HBA were not cloned. Subcloning of Pseudomonas sp. strain CBS3 genes revealed that most of the insert is required for the dehalogenation of 4-CBA, suggesting that more than one gene product is involved in this dehalogenation.  相似文献   

2.
Abstract Anaerobic bacteria are known to catalyze the removal of a variety of aromatic substituents, including -COOH, -OH, -OCH3, -CH3, and halogens. We investigated whether reductive dehalogenation was related to other types of aryl substituent removal reactions. A dehalogenating bacterial consortium was tested for its ability to use benzoic acids substituted in the 3 position with the functional groups listed above. In addition to dehalogenation, the enrichment (as well as the dehalogenating pure culture) was able to transform 3-methoxybenzoic acid to 3-hydroxybenzoic acid without a lag. This reaction exhibited Michaelis-Menten kinetics with an apparent K m of 5 μM. To test the hypothesis that the two reactions were related, we developed a mathematical model incorporating a competitive inhibition term to account for the influence of one substrate on the degradation of the other. However, experimental evidence showed no significant difference in the rates of 3-chlorobenzoic acid or 3-methoxybenzoic acid degradation in either the presence or absence of the other substrate. In addition, an anaerobe known to degrade methoxylated aromatic substrates was not able to transform chlorinated analogues. Finally, the isolated dechlorinating organism, strain DCB-1 was able to transform 3-methoxybenzoic acid in the presence of 1 mM thiosulfate, but the dehalogenation of 3-chlorobenzoic acid under these conditions was completely inhibited. Therefore, it is unlikely that a relationship exists between dehalogenation and other anaerobic aromatic substituent removal mechanisms.  相似文献   

3.
Summary Resting cells of the coryneform strain NTB-1, previously incorrectly classified as Alcaligenes denitrificans NTB-1, quantitatively converted 4-chlorobenzoate (4-CBA) to 4-hydroxybenzoate (4-HBA) under strict anaerobic conditions in the presence of ferricyanide or nitrate. 4-HBA formation was enhanced by supplying anaerobic cells with glucose as an energy source. Using permeabilized cells it was shown that energy is not needed to drive the energy-dependent uptake of 4-CBA but also to convert 4-CBA into 4-HBA. In extracts it was subsequently demonstrated that a coenzymeA-thioester of 4-CBA is involved in the metabolism of 4-CBA. Offprint requests to: P. E. J. Groenewegen  相似文献   

4.
O-methyl substituents of aromatic compounds can provide C1 growth substrates for facultative and strict anaerobic bacteria isolated from diverse environments. The mechanism of the bioconversion of methoxylated benzoic acids to the hydroxylated derivatives was investigated with a model substrate and cultures of one anaerobic consortium, eight strict anaerobic bacteria, and one facultative anaerobic microorganism. Using high-pressure liquid chromatography and gas chromatography-mass spectral analysis, we found that a haloaromatic dehalogenating consortium, a dehalogenating isolate from that consortium, Eubacterium limosum, and a strain of Acetobacterium woodii metabolized 3-[methoxy-18O]methoxybenzoic acid (3-anisic acid) to 3-[hydroxy-18O]hydroxybenzoic acid stoichiometrically at rates of 1.5, 3.2, 52.4, and 36.7 nmol/min per mg of protein, respectively. A different strain of Acetobacterium and strains of Syntrophococcus, Clostridium, Desulfotomaculum, Enterobacter, and an anaerobic bacterium, strain TH-001, were unable to transform this compound. The O-demethylating ability of E. limosum was induced only with appropriate methoxylated benzoates but not with D-glucose, lactate, isoleucine, or methanol. Cross-acclimation and growth experiments with E. limosum showed a rate of metabolism that was an order of magnitude slower and showed no growth with either 4-methoxysalicylic acid (2-hydroxy-4-methoxybenzoic acid) or 4-anisic acid (4-methoxybenzoic acid) when adapted to 3-anisic acid. However, A. woodii NZva-16 showed slower rates and no growth with 3- or 4-methoxysalicylic acid when adapted to 3-anisic acid in similar experiments. The results clearly indicate a methyl rather than methoxy group removal mechanism for such reactions.  相似文献   

5.
Microbial dehalogenation of chlorinated compounds in anaerobic environments is well known, but the degradation of fluorinated compounds under similar conditions has rarely been described. Here, we report on the isolation of a bovine rumen bacterium that metabolizes fluoroacetate under anaerobic conditions, the mode of degradation and its presence in gut ecosystems. The bacterium was identified using 16S rRNA gene sequence analysis as belonging to the phylum Synergistetes and was designated strain MFA1. Growth was stimulated by amino acids with greater quantities of amino acids metabolized in the presence of fluoroacetate, but sugars were not fermented. Acetate, formate, propionate, isobutryate, isovalerate, ornithine and H(2) were end products of amino acid metabolism. Acetate was the primary end product of fluoroacetate dehalogenation, and the amount produced correlated with the stoichiometric release of fluoride which was confirmed using fluorine nuclear magnetic resonance ((19) F NMR) spectroscopy. Hydrogen and formate produced in situ were consumed during dehalogenation. The growth characteristics of strain MFA1 indicated that the bacterium may gain energy via reductive dehalogenation. This is the first study to identify a bacterium that can anaerobically dehalogenate fluoroacetate. Nested 16S rRNA gene-specific PCR assays detected the bacterium at low numbers in the gut of several herbivore species.  相似文献   

6.
O-methyl substituents of aromatic compounds can provide C1 growth substrates for facultative and strict anaerobic bacteria isolated from diverse environments. The mechanism of the bioconversion of methoxylated benzoic acids to the hydroxylated derivatives was investigated with a model substrate and cultures of one anaerobic consortium, eight strict anaerobic bacteria, and one facultative anaerobic microorganism. Using high-pressure liquid chromatography and gas chromatography-mass spectral analysis, we found that a haloaromatic dehalogenating consortium, a dehalogenating isolate from that consortium, Eubacterium limosum, and a strain of Acetobacterium woodii metabolized 3-[methoxy-18O]methoxybenzoic acid (3-anisic acid) to 3-[hydroxy-18O]hydroxybenzoic acid stoichiometrically at rates of 1.5, 3.2, 52.4, and 36.7 nmol/min per mg of protein, respectively. A different strain of Acetobacterium and strains of Syntrophococcus, Clostridium, Desulfotomaculum, Enterobacter, and an anaerobic bacterium, strain TH-001, were unable to transform this compound. The O-demethylating ability of E. limosum was induced only with appropriate methoxylated benzoates but not with D-glucose, lactate, isoleucine, or methanol. Cross-acclimation and growth experiments with E. limosum showed a rate of metabolism that was an order of magnitude slower and showed no growth with either 4-methoxysalicylic acid (2-hydroxy-4-methoxybenzoic acid) or 4-anisic acid (4-methoxybenzoic acid) when adapted to 3-anisic acid. However, A. woodii NZva-16 showed slower rates and no growth with 3- or 4-methoxysalicylic acid when adapted to 3-anisic acid in similar experiments. The results clearly indicate a methyl rather than methoxy group removal mechanism for such reactions.  相似文献   

7.
The Arthrobacter sp. strain SU 4-chlorobenzoate (4-CBA) dehalogenation pathway converts 4-CBA to 4-hydroxybenzoate (4-HBA). The pathway operon contains the genes fcbA, fcbB, and fcbC (A. Schmitz, K. H. Gartemann, J. Fiedler, E. Grund, and R. Eichenlaub, Appl. Environ. Microbiol. 58:4068-4071, 1992). Genes fcbA and fcbB encode 4-CBA-coenzyme A (CoA) ligase and 4-CBA-CoA dehalogenase, respectively, whereas the function of fcbC is not known. We subcloned fcbC and expressed it in Escherichia coli, and we purified and characterized the FcbC protein. A substrate activity screen identified benzoyl-CoA thioesters as the most active substrates. Catalysis of 4-HBA-CoA hydrolysis to 4-HBA and CoA occurred with a k(cat) of 6.7 s(-1) and a K(m) of 1.2 micro M. The k(cat) pH rate profile for 4-HBA-CoA hydrolysis indicated optimal activity over a pH range of 6 to 10. The amino acid sequence of the FcbC protein was compared to other sequences contained in the protein sequence data banks. A large number of sequence homologues of unknown function were identified. On the other hand, the 4-HBA-CoA thioesterases isolated from 4-CBA-degrading Pseudomonas strains did not share significant sequence identity with the FcbC protein, indicating early divergence of the thioesterase-encoding genes.  相似文献   

8.
DL-2-Haloacid dehalogenase from Pseudomonas sp. 113 (DL-DEX 113) catalyzes the hydrolytic dehalogenation of D- and L-2-haloalkanoic acids, producing the corresponding L- and D-2-hydroxyalkanoic acids, respectively. Every halidohydrolase studied so far (L-2-haloacid dehalogenase, haloalkane dehalogenase, and 4-chlorobenzoyl-CoA dehalogenase) has an active site carboxylate group that attacks the substrate carbon atom bound to the halogen atom, leading to the formation of an ester intermediate. This is subsequently hydrolyzed, resulting in the incorporation of an oxygen atom of the solvent water molecule into the carboxylate group of the enzyme. In the present study, we analyzed the reaction mechanism of DL-DEX 113. When a single turnover reaction of DL-DEX 113 was carried out with a large excess of the enzyme in H(2)(18)O with a 10 times smaller amount of the substrate, either D- or L-2-chloropropionate, the major product was found to be (18)O-labeled lactate by ionspray mass spectrometry. After a multiple turnover reaction in H(2)(18)O, the enzyme was digested with trypsin or lysyl endopeptidase, and the molecular masses of the peptide fragments were measured with an ionspray mass spectrometer. No peptide fragments contained (18)O. These results indicate that the H(2)(18)O of the solvent directly attacks the alpha-carbon of 2-haloalkanoic acid to displace the halogen atom. This is the first example of an enzymatic hydrolytic dehalogenation that proceeds without producing an ester intermediate.  相似文献   

9.
Reductive anaerobic dehalogenation is a useful method for remediation of sites contaminated by chlorinated ethylenes, where hydrogen concentration plays the key role. Under anaerobic conditions, dehalogenating bacteria compete best against methanogenic consortia when the hydrogen level is low; and methanogenic consortia outplay dehalogenating bacteria when the hydrogen level is high. Thus, in an anaerobic mixed culture, efficient use of hydrogen for dehalogenation can be achieved by strategies that maintain hydrogen at a certain low concentration. However, due to the role of acetate, expected dehalogenating results cannot be obtained and unexpected methane formation can be encountered in practice.  相似文献   

10.
Abstract Desulfomonile tiedjei is a Gram-negative sulfate-reducing bacterium capable of catalyzing aryl reductive dehalogenation reactions. Since many toxic and persistent contaminants in the subsurface are halogenated aromatic compounds, the detection and enumeration of dehalogenating microorganisms in the environment may be a useful tool for planning and evaluating bioremediation efforts. In this study, we show that D. tiedjei contains unique lipopolysaccharide branched 3-hydroxy fatty acids, unknown as yet in other bacteria, and that it is possible to detect the bacterium in inoculated aquifer sediments based on these signature lipid biomarkers. The detection of D. tiedjei and other dehalogenating microorganisms possessing similar cellular properties in environmental matrices may be possible by this technique. Additionally, the effect of such inoculation on dehalogenation activity is examined.  相似文献   

11.
We isolated a strain of Rhodopseudomonas palustris (RCB100) by selective enrichment in light on 3-chlorobenzoate to investigate the steps that it uses to accomplish anaerobic dechlorination. Analyses of metabolite pools as well as enzyme assays suggest that R. palustris grows on 3-chlorobenzoate by (i) converting it to 3-chlorobenzoyl coenzyme A (3-chlorobenzoyl-CoA), (ii) reductively dehalogenating 3-chlorobenzoyl-CoA to benzoyl-CoA, and (iii) degrading benzoyl-CoA to acetyl-CoA and carbon dioxide. R. palustris uses 3-chlorobenzoate only as a carbon source and thus incorporates the acetyl-CoA that is produced into cell material. The reductive dechlorination route used by R. palustris for 3-chlorobenzoate degradation differs from those previously described in that a CoA thioester, rather than an unmodified aromatic acid, is the substrate for complete dehalogenation.  相似文献   

12.
An Arthrobacter sp. has been shown to dehalogenate 4-chlorobenzoate yielding 4-hydroxybenzoate. Experiments with 18O indicate that, in the presence of cell-free extracts, the hydroxyl group which is substituted onto the aromatic nucleus during dehalogenation is derived from water and not from molecular oxygen. Dehalogenation therefore is not catalysed by a mixed-function oxidase; instead a novel aromatic hydroxylase is implicated in the reaction.  相似文献   

13.
Rhodococcus wratislaviensis strain KT112-7 (VKM As-2623D) degraded monosubstituted 2-/4-chlorobiphenyls (94.25 mg/L) and 2,4'-dichlorobiphenyl (22.3 mg/L) when the sodium chloride content in the culture medium did not exceed 50 g/L. We have calculated the kinetic parameters of chlorobiphenyl (CB) decomposition by the КТ112-7 strain, which demonstrated decreased decomposition efficiency with increasing NaCl concentration. A linear correlation has been found between the content of saturated, unsaturated, and branched fatty acids in the cell wall of the KT112-7 strain and the NaCl content in the culture medium. The change in the content and composition of fatty acids in the cell wall apparently led to a change in the permeability of the CB cell wall. R. wratislaviensis strain KT112-7 was able to oxidize the unsubstituted ring in 2-CB and 4-CB at all studied NaCl concentrations in the medium and efficiently decomposed the chlorobenzoic acids (CBAs) that formed as a result of the oxidation. Decomposition of 2,4'-CB at low NaCl concentrations in the medium (not higher than 10 g/L) involved the oxidation of the ring substituted by chlorine in the position 4, while the presence of 20–50 g/L of NaCl led to the oxidation of the ring substituted by chlorine in the position 2. The main detected metabolite was 2-chlorobenzoic acid in the former case and 4-chlorobenzoic acid in the latter. The 4-CBA transformation proceeded via the formation of protocatechuic acid (3,4-hydroxybenzoic acid) and catechol at all of the studied sodium chloride concentrations in the medium and through another pathway involving the formation of chlorocatechol at NaCl concentrations of 30, 40, and 50 g/L.  相似文献   

14.
The Arthrobacter sp. strain SU 4-chlorobenzoate (4-CBA) dehalogenation pathway converts 4-CBA to 4-hydroxybenzoate (4-HBA). The pathway operon contains the genes fcbA, fcbB, and fcbC (A. Schmitz, K. H. Gartemann, J. Fiedler, E. Grund, and R. Eichenlaub, Appl. Environ. Microbiol. 58:4068-4071, 1992). Genes fcbA and fcbB encode 4-CBA-coenzyme A (CoA) ligase and 4-CBA-CoA dehalogenase, respectively, whereas the function of fcbC is not known. We subcloned fcbC and expressed it in Escherichia coli, and we purified and characterized the FcbC protein. A substrate activity screen identified benzoyl-CoA thioesters as the most active substrates. Catalysis of 4-HBA-CoA hydrolysis to 4-HBA and CoA occurred with a kcat of 6.7 s−1 and a Km of 1.2 μM. The kcat pH rate profile for 4-HBA-CoA hydrolysis indicated optimal activity over a pH range of 6 to 10. The amino acid sequence of the FcbC protein was compared to other sequences contained in the protein sequence data banks. A large number of sequence homologues of unknown function were identified. On the other hand, the 4-HBA-CoA thioesterases isolated from 4-CBA-degrading Pseudomonas strains did not share significant sequence identity with the FcbC protein, indicating early divergence of the thioesterase-encoding genes.  相似文献   

15.
An enrichment culture, derived from the anaerobic stage of a two-step sequential anaerobic-aerobic reactor system which mineralized 2,4,6-trichlorophenol, stoichiometrically converted 2,4,6-trichlorophenol to 4-chlorophenol. Dehalogenation occurred only in alkaline media (pH 8–9) at concentrations of substrate up to 1 mmol 11. Formate plus acetate or trypticase could serve as electron donors. Neither vitamins nor trace elements were required in a chloride-free defined medium. The dehalogenating organism was oxygen-resistant, but was not active in media which were oxidized with respect to resazurin indicator dye. Most probable number counts of the dehalogenating cultures showed that the dehalogenating organisms were present in very small numbers, yet catalysed dehalogenation at rates considerably faster than other dehalogenating organisms described in the literature.  相似文献   

16.
The strains of Arthrobacter globiformis KZT1, Corynebacterium sepedonicum KZ4 and Pseudomonas cepacia KZ2 capable of early dehalogenation and complete oxidation of 4-chloro-, 2,4-dichloro-and 2-chlorobenzoic acids, respectively, have been analyzed for the origin of the genetic control of degradation. The occurrence and molecular sizes of plasmids in all the strains have been established. Plasmid pBS1501 was shown to control 4-chlorobenzoate dehalogenation in the case of KZT1 strain. The same possibility is proposed for plasmid pBS1502 for dehalogenation of 2,4DCBA by KZ4 strain. The chromosome localization of the genes controlling oxidation of 4-hydroxybenzoate in strain KZT1 is shown. Localization of the whole set of genes responsible for 2CBA degradation in the strain KZ2 chromosome is suggested.  相似文献   

17.
An Arthrobacter strain, able to utilize 4-chlorobenzoic acid as the sole carbon and energy source, was isolated and characterized. The first step of the catabolic pathway was found to proceed via a hydrolytic dehalogenation that leads to the formation of 4-hydroxybenzoic acid. The dehalogenase encoding genes (fcb) were sequenced and found highly homologous to and organized as those of other 4-chlorobenzoic acid degrading Arthrobacter strains. The fcb genes were cloned and successfully expressed in the heterologous host Pseudomonas putida PaW340 and P. putida KT2442 upper TOL, which acquired the ability to grow on 4-chlorobenzoic acid and 4-chlorotoluene, respectively. The cloned dehalogenase displayed a high specificity for para-substituted haloaromatics with affinity Cl > Br > I > F, in the order.  相似文献   

18.
The enzyme which catalyzes the dehalogenation of 2,4,6-trichlorophenol (TCP) was purified to apparent homogeneity from an extract of TCP-induced cells of Azotobacter sp. strain GP1. The initial step of TCP degradation in this bacterium is inducible by TCP; no activity was found in succinate-grown cells or in phenol-induced cells. NADH, flavin adenine dinucleotide, and O2 are required as cofactors. As reaction products, 2,6-dichlorohydroquinone and Cl- ions were identified. Studies of the stoichiometry revealed the consumption of 2 mol of NADH plus 1 mol of O2 per mol of TCP and the formation of 1 mol of Cl- ions. No evidence for membrane association or for a multicomponent system was obtained. Molecular masses of 240 kDa for the native enzyme and 60 kDa for the subunit were determined, indicating a homotetrameric structure. Cross-linking studies with dimethylsuberimidate were consistent with this finding. TCP was the best substrate for 2,4,6-trichlorophenol-4-monooxygenase (TCP-4-monooxygenase). The majority of other chlorophenols converted by the enzyme bear a chloro substituent in the 4-position. 2,6-Dichlorophenol, also accepted as a substrate, was hydroxylated in the 4-position to 2,6-dichlorohydroquinone in a nondehalogenating reaction. NADH and O2 were consumed by the pure enzyme also in the absence of TCP with simultaneous production of H2O2. The NH2-terminal amino acid sequence of TCP-4-monooxygenase from Azotobacter sp. strain GP1 revealed complete identity with the nucleotide-derived sequence from the analogous enzyme from Pseudomonas pickettii and a high degree of homology with two nondehalogenating monooxygenases. The similarity in enzyme properties and the possible evolutionary relatedness of dehalogenating and nondehalogenating monooxygenases are discussed.  相似文献   

19.
Cupriavidus sp. strain SK-3, previously isolated on polychlorinated biphenyl mixtures, was found to aerobically utilize a wide spectrum of substituted aromatic compounds including 4-fluoro-, 4-chloro- and 4-bromobenzoic acids as a sole carbon and energy source. Other chlorobenzoic acid (CBA) congeners such as 2-, 3-, 2,3-, 2,5-, 3,4- and 3,5-CBA were all rapidly transformed to respective chlorocatechols (CCs). Under aerobic conditions, strain SK-3 grew readily on 4-CBA to a maximum concentration of 5 mM above which growth became impaired and yielded no biomass. Growth lagged significantly at concentrations above 3 mM, however chloride elimination was stoichiometric and generally mirrored growth and substrate consumption in all incubations. Experiments with resting cells, cell-free extracts and analysis of metabolite pools suggest that 4-CBA was metabolized in a reaction exclusively involving an initial hydrolytic dehalogenation yielding 4-hydroxybenzoic acid, which was then hydroxylated to protocatechuic acid (PCA) and subsequently metabolized via the β-ketoadipate pathway. When strain SK-3 was grown on 4-CBA, there was gratuitous induction of the catechol-1,2-dioxygenase and gentisate-1,2-dioxygenase pathways, even if both were not involved in the metabolism of the acid. While activities of the modified ortho- and meta-cleavage pathways were not detectable in all extracts, activity of PCA-3,4-dioxygenase was over ten-times higher than those of catechol-1,2- and gentisate-1,2-dioxygenases. Therefore, the only reason other congeners were not utilized for growth was the accumulation of CCs, suggesting a narrow spectrum of the activity of enzymes downstream of benzoate-1,2-dioxygenase, which exhibited affinity for a number of substituted analogs, and that the metabolic bottlenecks are either CCs or catabolites of the modified ortho-cleavage metabolic route.  相似文献   

20.
Rhodopseudomonas palustris strain RCB100 degrades 3-chlorobenzoate (3-CBA) anaerobically. We purified from this strain a coenzyme A ligase that is active with 3-CBA and determined its N-terminal amino acid sequence to be identical to that of a cyclohexanecarboxylate-CoA ligase encoded by aliA from the R. palustris strain (CGA009) that has been sequenced. Strain CGA009 differs from strain RCB100 in that it does not use 3-CBA as a sole carbon source. The aliA gene from the 3-CBA degrading strain differed by a single nucleotide from the aliA gene from strain CGA009, causing the substitution of a serine for a threonine at position 208. Both AliA enzymes, purified as His-tagged fusion proteins, had comparable activities with cyclohexanecarboxylate. However, AliA from the 3-CBA degrading strain was 10-fold more active with 3-CBA (kcat/Km of 4.3 x 10(4) M(-1) s(-1)) than the enzyme from the sequenced strain (kcat/Km 0.32 x 10(4) M(-1) s(-1)). The CGA009 enzyme was not sufficiently active with 3-CBA to complement an RCB100 aliA mutant for growth on this compound. Here, whole genome sequence information enabled us to identify a single nucleotide among 5.4 million nucleotides that contributes to the substrate preference of a coenzyme A ligase.  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