首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The enzymatic dearomatization of aromatic ring systems by reduction represents a highly challenging redox reaction in biology and plays a key role in the degradation of aromatic compounds under anoxic conditions. In anaerobic bacteria, most monocyclic aromatic growth substrates are converted to benzoyl‐coenzyme A (CoA), which is then dearomatized to a conjugated dienoyl‐CoA by ATP‐dependent or ‐independent benzoyl‐CoA reductases. It was unresolved whether or not related enzymes are involved in the anaerobic degradation of environmentally relevant polycyclic aromatic hydrocarbons (PAHs). In this work, a previously unknown dearomatizing 2‐naphthoyl‐CoA reductase was purified from extracts of the naphthalene‐degrading, sulphidogenic enrichment culture N47. The oxygen‐tolerant enzyme dearomatized the non‐activated ring of 2‐naphthoyl‐CoA by a four‐electron reduction to 5,6,7,8‐tetrahydro‐2‐naphthoyl‐CoA. The dimeric 150 kDa enzyme complex was composed of a 72 kDa subunit showing sequence similarity to members of the flavin‐containing ‘old yellow enzyme’ family. NCR contained FAD, FMN, and an iron‐sulphur cluster as cofactors. Extracts of Escherichia coli expressing the encoding gene catalysed 2‐naphthoyl‐CoA reduction. The identified NCR is a prototypical enzyme of a previously unknown class of dearomatizing arylcarboxyl‐CoA reductases that are involved in anaerobic PAH degradation; it fundamentally differs from known benzoyl‐CoA reductases.  相似文献   

2.
The degradation of the industrially produced and environmentally relevant phthalate esters by microorganisms is initiated by the hydrolysis to alcohols and phthalate (1,2‐dicarboxybenzene). In the absence of oxygen the further degradation of phthalate proceeds via activation to phthaloyl‐CoA followed by decarboxylation to benzoyl‐CoA. Here, we report on the first purification and characterization of a phthaloyl‐CoA decarboxylase (PCD) from the denitrifying Thauera chlorobenzoica. Hexameric PCD belongs to the UbiD‐family of (de)carboxylases and contains prenylated FMN (prFMN), K+ and, unlike other UbiD‐like enzymes, Fe2+ as cofactors. The latter is suggested to be involved in oxygen‐independent electron‐transfer during oxidative prFMN maturation. Either oxidation to the Fe3+‐state in air or removal of K+ by desalting resulted in >92% loss of both, prFMN and decarboxylation activity suggesting the presence of an active site prFMN/Fe2+/K+‐complex in PCD. The PCD‐catalysed reaction was essentially irreversible: neither carboxylation of benzoyl‐CoA in the presence of 2 M bicarbonate, nor an isotope exchange of phthaloyl‐CoA with 13C‐bicarbonate was observed. PCD differs in many aspects from prFMN‐containing UbiD‐like decarboxylases and serves as a biochemically accessible model for the large number of UbiD‐like (de)carboxylases that play key roles in the anaerobic degradation of environmentally relevant aromatic pollutants.  相似文献   

3.
Quaternary carbon-containing compounds exist in natural and fossil oil-derived products and are used in chemical and pharmaceutical applications up to industrial scale. Due to the inaccessibility of the quaternary carbon atom for a direct oxidative or reductive attack, they are considered as persistent in the environment. Here, we investigated the unknown degradation of the quaternary carbon-containing model compound pivalate (2,2-dimethyl-propionate) in the denitrifying bacterium Thauera humireducens strain PIV-1 (formerly Thauera pivalivorans). We provide multiple evidence for a pathway comprising the activation to pivalyl-CoA and the carbon skeleton rearrangement to isovaleryl-CoA. Subsequent reactions proceed similar to the catabolic leucine degradation pathway such as the carboxylation to 3-methylglutaconyl-CoA and the cleavage of 3-methyl-3-hydroxyglutaryl-CoA to acetyl-CoA and acetoacetate. The completed genome of Thauera humireducens strain PIV-1 together with proteomic data was used to identify pivalate-upregulated gene clusters including genes putatively encoding pivalate CoA ligase and adenosylcobalamin-dependent pivalyl-CoA mutase. A pivalate-induced gene encoding a putative carboxylic acid CoA ligase was heterologously expressed, and its highly enriched product exhibited pivalate CoA ligase activity. The results provide the first experimental insights into the biodegradation pathway of a quaternary carbon-containing model compound that serves as a blueprint for the degradation of related quaternary carbon-containing compounds.  相似文献   

4.
5.
Phthalic acid esters (phthalates) are anthropogenic compounds that are used as plasticizers. Unfortunately, because phthalates are non‐covalently intercalated into plastic polymers they leach into the environment, accumulating in anoxic sediments. This has negative consequences for animal and human health. Denitrifying Betaproteobacteria, such as Aromatoleum aromaticum, can use ortho‐phthalate, derived by ester hydrolysis, as a carbon and energy source. Mergelsberg et al. ( 2018 ) deconstruct the pathway whereby ortho‐phthalic acid is converted, via the highly unstable phthaloyl‐CoA, to the central intermediate of anaerobic aromatic degradation, benzoyl‐CoA. The latter reaction is catalysed by UbiD‐like phthaloyl‐CoA decarboxylase (PCD). Succinyl‐CoA:o‐phthalate CoA‐transferase (SPT) generates phthaloyl‐CoA, which accumulates at only sub‐micromolar concentrations, while the Km of PCD for phthaloyl‐CoA is two‐orders of magnitude higher. This seemingly insurmountable kinetic barrier is overcome because A. aromatoleum massively over‐produces PCD and because the decarboxylation reaction is irreversible. These features of the pathway facilitate capture of phthaloyl‐CoA as it is released from SPT without the need for direct substrate‐channelling. The authors provide strong evidence from both in vivo and in vitro studies to support their conclusions. This work reveals how these anaerobic bacteria have rapidly evolved a stop‐gap measure to allow them to completely degrade an otherwise recalcitrant aromatic xenobiotic.  相似文献   

6.
Aim: To characterize the microbial community structure and bamA gene diversity involved in anaerobic degradation of toluene and benzoate under denitrifying conditions. Methods and Results: Nitrate‐reducing enrichment cultures were established on either toluene, benzoate or without additional substrate. Bacterial community structures were characterized by 16S rRNA gene–based PCR‐DGGE analysis. bamA gene diversity was analysed using DGGE and cloning methods. The results showed that bamA gene related to bamA of Thauera chlorobenzoica was dominant in toluene and benzoate cultures. However, a greater diversity of sequences was obtained in benzoate cultures. Low diversity of bamA gene was observed, and some similarities of bamA were also found between active cultures and backgrounds, suggesting that potential natural attenuation of aromatic compounds might occur. Conclusions: The combined analysis of 16S rRNA and bamA genes suggests that the species related to genera Thauera dominated toluene‐ and benzoate‐degrading cultures. The combination of multiple methods (DGGE and cloning) provides a more complete picture of bamA gene diversity. Significance and Impact of the Study: To our knowledge, this is the first report of bamA gene in denitrifying enrichments using DGGE and cloning analysis.  相似文献   

7.
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   

8.
Denitrifying bacteria capable of degrading halobenzoates were isolated from various geographical and ecological sites. The strains were isolated after initial enrichment on one of the monofluoro-, monochloro-, or monobromo-benzoate isomers with nitrate as an electron acceptor, yielding a total of 33 strains isolated from the different halobenzoate-utilizing enrichment cultures. Each isolate could grow on the selected halobenzoate with nitrate as the terminal electron acceptor. The isolates obtained on 2-fluorobenzoate could use 2-fluorobenzoate under both aerobic and denitrifying conditions, but did not degrade other halobenzoates. In contrast, the 4-fluorobenzoate isolates degraded 4-fluorobenzoate under denitrifying conditions only, but utilized 2-fluorobenzoate under both aerobic and denitrifying conditions. The strains isolated on either 3-chlorobenzoate or 3-bromobenzoate could use 3-chlorobenzoate, 3-bromobenzoate, and 2- and 4-fluorobenzoates under denitrifying conditions. The isolates were identified and classified on the basis of 16S rRNA gene sequence analysis and their cellular fatty acid profiles. They were placed in nine genera belonging to either the alpha-, beta-, or gamma-branch of the Proteobacteria, namely, Acidovorax, Azoarcus, Bradyrhizobium, Ochrobactrum, Paracoccus, Pseudomonas, Mesorhizobium, Ensifer, and Thauera. These results indicate that the ability to utilize different halobenzoates under denitrifying conditions is ubiquitously distributed in the Proteobacteria and that these bacteria are widely distributed in soils and sediments.  相似文献   

9.
In the denitrifying member of the beta-Proteobacteria Thauera aromatica, the anaerobic metabolism of aromatic acids such as benzoate or 2-aminobenzoate is initiated by the formation of the coenzyme A (CoA) thioester, benzoyl-CoA and 2-aminobenzoyl-CoA, respectively. Both aromatic substrates were transformed to the acyl-CoA intermediate by a single CoA ligase (AMP forming) that preferentially acted on benzoate. This benzoate-CoA ligase was purified and characterized as a 57-kDa monomeric protein. Based on V(max)/K(m), the specificity constant for 2-aminobenzoate was 15 times lower than that for benzoate; this may be the reason for the slower growth on 2-aminobenzoate. The benzoate-CoA ligase gene was cloned and sequenced and was found not to be part of the gene cluster encoding the general benzoyl-CoA pathway of anaerobic aromatic metabolism. Rather, it was located in a cluster of genes coding for a novel aerobic benzoate oxidation pathway. In line with this finding, the same CoA ligase was induced during aerobic growth with benzoate. A deletion mutant not only was unable to grow anaerobically on benzoate or 2-aminobenzoate, but also aerobic growth on benzoate was affected. This suggests that benzoate induces a single benzoate-CoA ligase. The product of benzoate activation, benzoyl-CoA, then acts as inducer of separate anaerobic or aerobic pathways of benzoyl-CoA, depending on whether oxygen is lacking or present.  相似文献   

10.
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   

11.
Xenobiotic phthalates are industrially produced on the annual million ton scale. The oxygen‐independent enzymatic reactions involved in anaerobic phthalate degradation have only recently been elucidated. In vitro assays suggested that phthalate is first activated to phthaloyl‐CoA followed by decarboxylation to benzoyl‐CoA. Here, we report the heterologous production and characterization of the enzyme initiating anaerobic phthalate degradation from ‘Aromatoleum aromaticum’: a highly specific succinyl‐CoA:phthalate CoA transferase (SPT, class III CoA transferase). Phthaloyl‐CoA formed by SPT accumulated only to sub‐micromolar concentrations due to the extreme lability of the product towards intramolecular substitution with a half‐life of around 7 min. Upon addition of excess phthaloyl‐CoA decarboxylase (PCD), the combined activity of both enzymes was drastically shifted towards physiologically relevant benzoyl‐CoA formation. In conclusion, a massive overproduction of PCD in phthalate‐grown cells to concentrations >140 μM was observed that allowed for efficient phthaloyl‐CoA conversion at concentrations 250‐fold below the apparent Km‐value of PCD. The results obtained provide insights into an only recently evolved xenobiotic degradation pathway where a massive cellular overproduction of PCD compensates for the formation of the probably most unstable CoA ester intermediate in biology.  相似文献   

12.
The denitrifying betaproteobacterium Sterolibacterium denitrificans serves as model organism for studying the oxygen‐independent degradation of cholesterol. Here, we demonstrate its capability of degrading various globally abundant side chain containing zoo‐, phyto‐ and mycosterols. We provide the complete genome that empowered an integrated genomics/proteomics/metabolomics approach, accompanied by the characterization of a characteristic enzyme of steroid side chain degradation. The results indicate that individual molybdopterin‐containing steroid dehydrogenases are involved in C25‐hydroxylations of steroids with different isoprenoid side chains, followed by the unusual conversion to C26‐oic acids. Side chain degradation to androsta‐1,4‐diene‐3,17‐dione (ADD) via aldolytic C–C bond cleavages involves acyl‐CoA synthetases/dehydrogenases specific for the respective 26‐, 24‐ and 22‐oic acids/‐oyl‐CoAs and promiscuous MaoC‐like enoyl‐CoA hydratases, aldolases and aldehyde dehydrogenases. Degradation of rings A and B depends on gene products uniquely found in anaerobic steroid degraders, which after hydrolytic cleavage of ring A, again involves CoA‐ester intermediates. The degradation of the remaining CD rings via hydrolytic cleavage appears to be highly similar in aerobic and anaerobic bacteria. Anaerobic cholesterol degradation employs a composite repertoire of more than 40 genes partially known from aerobic degradation in gammaproteobacteria/actinobacteria, supplemented by unique genes that are required to circumvent oxygenase‐dependent reactions.  相似文献   

13.
The initial reactions possibly involved in the acrobic and anaerobic metabolism of aromatic acids by a denitrifying Pseudomonas strain were studied. Several acyl CoA synthetases were found supporting the view that activation of several aromatic acids preceeds degradation. A benzoyl CoA synthetase activity (AMP forming) (apparent K m values of the enzyme from nitrate grown cells: 0.01 mM benzoate, 0.2 mM ATP, 0.2 mM coenzyme A) was present in aerobically grown and anaerobically, nitrate grown cells when benzoate or other aromatic acids were present. In addition to benzoate and fluorobenzoates, also 2-amino-benzoate was activated, albeit with unfavorable K m (0.5 mM 2-aminobenzoate). A 2-aminobenzoyl CoA synthetase (AMP forming) was induced both aerobically and anaerobically with 2-aminobenzoate as growth substrate which had a similar substrate spectrum but a low K m for 2-aminobenzoate (<0.02 mM). Anaerobic growth on 4-hydroxybenzoate induced a 4-hydroxybenzoyl CoA synthetase, and cyclohexanecarboxylate induced another synthetase. In contrast, 3-hydroxybenzoate and phenyl-acetate grown anaerobic cells appeared not to activate the respective substrates at sufficient rates. Contrary to an earlier report extracts from aerobic and anaerobic 2-aminobenzoate grown cells catalysed a 2-aminobenzoyl CoA-dependent NADH oxidation. This activity was 10–20 times higher in aerobic cells and appeared to be induced by 2-aminobenzoate and oxygen. In vitro, 2-aminobenzoyl CoA reduction was dependent on 2-aminobenzoyl CoA NAD(P)H, and oxygen. A novel mechanism of aerobic 2-aminobenzoate degradation is suggested, which proceeds via 2-aminobenzoyl CoA.  相似文献   

14.
The microbially mediated reductive dehalogenation of aromatic compounds is potentially important in removal of chlorinated aromatic compounds from the environment. Thermodynamic data are presented which show that the reductive dechlorination of 3-chlorobenzoate to benzoate is exergonic, which led to the hypothesis that reductive elimination of chlorine from 3-chlorobenzoate yields biologically useful energy. In the present paper this hypothesis is tested. Experimental data were obtained with a defined 3-chlorobenzoate degrading methanogenic consortium. These data showed that (i) the molar growth yield of a defined 3-chlorobenzoate degrading consortium increased from 4.9 g protein per mol benzoate metabolized to 6.8 g protein per mol 3-chlorobenzoate when 3-chlorobenzoate replaced benzoate as energy source, and that (ii) the ATP level in starved consortium cells was twice as high when the cells were fed 3-chlorobenzoate than when fed benzoate. These observations show that the electrochemical potential between the redox partners of the H+/H2 (electron-donating) and 3-chlorobenzoate/benzoate (electron-accepting) couples is a potential source of energy and are consistent with the hypothesis that reductive dechlorination of aromatic compounds is coupled to a novel type of microbial chemotrophy.  相似文献   

15.
Methane-producing freshwater lake sediment was found to dehalogenate chloro-, bromo-, and iodobenzoates by a reductive reaction in which the halogen was replaced by a hydrogen atom. The identity of the dehalogenated products was confirmed by mass spectrometry, nuclear magnetic resonance, or cochromatography. Removal of the halogens to produce benzoate was necessary before mineralization to CH4 + CO2 could occur. The dehalogenation occurred after a lag period which lasted from 1 week to more than 6 months, depending on the chemical. Dehalogenation was not observed in the absence of CH4 production, and it was inhibited by the addition of 20% O2. Once sediment was acclimated to halobenzoate dehalogenation, new additions of the halobenzoate were degraded without lag. Acclimation was observed regardless of whether the parent substrates were eventually mineralized to CH4 + CO2. Sediment acclimated to bromo- and chlorobenzoate degradation generally metabolized bromo- and chlorobenzoates, but sediment acclimated to iodobenzoate degradation only metabolized iodobenzoate. Prior acclimation of sediment to benzoate decomposition did not alter the pattern of dehalogenation, and sediment acclimated to dehalogenation was not concurrently acclimated to benzoate degradation. The presence of this apparent specificity, the lag period, and subsequent acclimation, together with our findings of the absence of dehalogenation in sterile sediments and by sediments previously incubated at ≥39°C, suggests that this reaction was biologically catalyzed. Apparently, a pathway for the reductive dehalogenation of aryl halides is present in anaerobic microorganisms of this methanogenic sediment.  相似文献   

16.
Dehalogenation is the key step in the degradation of halogenated aromatics, while reductive dehalogenation is originally thought to rarely occur in aerobes. In this study, an aerobic strain of Comamonas sp. 7D‐2 was shown to degrade the brominated aromatic herbicide bromoxynil completely and release two equivalents of bromides under aerobic conditions. The enzymes involved in the degradation of bromoxynil to 4‐carboxy‐2‐hydroxymuconate‐6‐semialdehyde, including nitrilase, reductive dehalogenase (BhbA), 4‐hydroxybenzoate 3‐monooxygenase and protocatechuate 4,5‐dioxygenase, were molecularly characterized. The novel dehalogenase BhbA was shown to be a complex of a respiration‐linked reductive dehalogenase (RdhA) domain and a NAD(P)H‐dependent oxidoreductase domain and to have key features of anaerobic respiratory RdhAs, including two predicted binding motifs for [4Fe‐4S] clusters and a close association with a hydrophobic membrane protein (BhbB). BhbB was confirmed to anchor BhbA to the membrane. BhbA was partially purified and found to use NAD(P)H as electron donors. Full‐length bhbA homologues were found almost exclusively in marine aerobic proteobacteria, suggesting that reductive dehalogenation occurs extensively in aerobes and that bhbA is horizontally transferred from marine microorganisms. The discovery of a functional reductive dehalogenase and ring‐cleavage oxygenases in an aerobe opens up possibilities for basic research as well as the potential application for bioremediation.  相似文献   

17.
Next to carbohydrates, aromatic compounds are the second most abundant class of natural organic molecules in living organic matter but also make up a significant proportion of fossil carbon sources. Only microorganisms are capable of fully mineralizing aromatic compounds. While aerobic microbes use well‐studied oxygenases for the activation and cleavage of aromatic rings, anaerobic bacteria follow completely different strategies to initiate catabolism. The key enzymes related to aromatic compound degradation in anaerobic bacteria are comprised of metal‐ and/or flavin‐containing cofactors, of which many use unprecedented radical mechanisms for C–H bond cleavage or dearomatization. Over the past decade, the increasing number of completed genomes has helped to reveal a large variety of anaerobic degradation pathways in Proteobacteria, Gram‐positive microbes and in one archaeon. This review aims to update our understanding of the occurrence of aromatic degradation capabilities in anaerobic microorganisms and serves to highlight characteristic enzymatic reactions involved in (i) the anoxic oxidation of alkyl side chains attached to aromatic rings, (ii) the carboxylation of aromatic rings and (iii) the reductive dearomatization of central arylcarboxyl‐coenzyme A intermediates. Depending on the redox potential of the electron acceptors used and the metabolic efficiency of the cell, different strategies may be employed for identical overall reactions.  相似文献   

18.
The anaerobic metabolism of catechol (1,2-dihydroxybenzene) was studied in the betaproteobacterium Thauera aromatica that was grown with CO2 as a cosubstrate and nitrate as an electron acceptor. Based on different lines of evidence and on our knowledge of enzymes and genes involved in the anaerobic metabolism of other aromatic substrates, the following pathway is proposed. Catechol is converted to catechylphosphate by phenylphosphate synthase, which is followed by carboxylation by phenylphosphate carboxylase at the para position to the phosphorylated phenolic hydroxyl group. The product, protocatechuate (3,4-dihydroxybenzoate), is converted to its coenzyme A (CoA) thioester by 3-hydroxybenzoate-CoA ligase. Protocatechuyl-CoA is reductively dehydroxylated to 3-hydroxybenzoyl-CoA, possibly by 4-hydroxybenzoyl-CoA reductase. 3-Hydroxybenzoyl-CoA is further metabolized by reduction of the aromatic ring catalyzed by an ATP-driven benzoyl-CoA reductase. Hence, the promiscuity of several enzymes and regulatory proteins may be sufficient to create the catechol pathway that is made up of elements of phenol, 3-hydroxybenzoate, 4-hydroxybenzoate, and benzoate metabolism.  相似文献   

19.
The enzymes catalyzing the formation of coenzyme A (CoA) thioesters of benzoate and 2-aminobenzoate were studied in a denitrifying Pseudomonas sp. anaerobically grown with these aromatic acids and nitrate as sole carbon and energy sources. Three different rather specific aromatic acyl-CoA ligases, E1, E2, and E3, were found which catalyze the formation of CoA thioesters of benzoate, fluorobenzoates, and 2-aminobenzoate. ATP is cleaved into AMP and pyrophosphate. The enzymes were purified, their N-terminal amino acid sequences were determined, and their catalytic and molecular properties were studied. Cells anaerobically grown on benzoate and nitrate contain one CoA ligase (AMP forming) for benzoic acid (E1). It is a homodimer of Mr 120,000 which prefers benzoate as a substrate but shows some activity also with 2-aminobenzoate and fluorobenzoates, although with lower Km. Cells anaerobically grown on 2-aminobenzoate and nitrate contain three different CoA ligases for aromatic acids. The first one is identical with benzoate-CoA ligase (E1). The second enzyme is a 2-aminobenzoate-CoA ligase (E2). It is a monomer of Mr 60,000 which prefers 2-aminobenzoate but also activates benzoate, fluorobenzoates and, less effectively, 2-methylbenzoate, with lower affinities to the latter substrates. The enzymes E1 and E2 have similar activity levels; a third minor CoA ligase activity is due to a different 2-aminobenzoate-CoA ligase. The enzyme (E3) is a monomer of Mr, 65,000 which 2-aminobenzoate pathway (U. Altenschmidt, C. Eckerskorn, and G. Fuchs, Eur. J. Biochem. 194:647-653, 1990); apparently, it is not completely repressed under anaerobic conditions and therefore also is induced to a small extent by 2-aminobenzoate under anaerobic growth conditions.  相似文献   

20.
The cyclohexane derivative cis‐2‐(carboxymethyl)cyclohexane‐1‐carboxylic acid [(1R,2R)‐/(1S,2S)‐2‐(carboxymethyl)cyclohexane‐1‐carboxylic acid] has previously been identified as metabolite in the pathway of anaerobic degradation of naphthalene by sulfate‐reducing bacteria. We tested the corresponding CoA esters of isomers and analogues of this compound for conversion in cell free extracts of the anaerobic naphthalene degraders Desulfobacterium strain N47 and Deltaproteobacterium strain NaphS2. Conversion was only observed for the cis‐isomer, verifying that this is a true intermediate and not a dead‐end product. Mass‐spectrometric analyses confirmed that conversion is performed by an acyl‐CoA dehydrogenase and a subsequent hydratase yielding an intermediate with a tertiary hydroxyl‐group. We propose that a novel kind of ring‐opening lyase is involved in the further catabolic pathway proceeding via pimeloyl‐CoA. In contrast to degradation pathways of monocyclic aromatic compounds where ring‐cleavage is achieved via hydratases, this lyase might represent a new ring‐opening strategy for the degradation of polycyclic compounds. Conversion of the potential downstream metabolites pimeloyl‐CoA and glutaryl‐CoA was proved in cell free extracts, yielding 2,3‐dehydropimeloyl‐CoA, 3‐hydroxypimeloyl‐CoA, 3‐oxopimeloyl‐CoA, glutaconyl‐CoA, crotonyl‐CoA, 3‐hydroxybutyryl‐CoA and acetyl‐CoA as observable intermediates. This indicates a link to central metabolism via β‐oxidation, a non‐decarboxylating glutaryl‐CoA dehydrogenase and a subsequent glutaconyl‐CoA decarboxylase.  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