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 46 毫秒
1.
The metabolism of benzoate, cyclohex-1-ene carboxylate, and cyclohexane carboxylate by "Syntrophus aciditrophicus" in cocultures with hydrogen-using microorganisms was studied. Cyclohexane carboxylate, cyclohex-1-ene carboxylate, pimelate, and glutarate (or their coenzyme A [CoA] derivatives) transiently accumulated during growth with benzoate. Identification was based on comparison of retention times and mass spectra of trimethylsilyl derivatives to the retention times and mass spectra of authentic chemical standards. (13)C nuclear magnetic resonance spectroscopy confirmed that cyclohexane carboxylate and cyclohex-1-ene carboxylate were produced from [ring-(13)C(6)]benzoate. None of the metabolites mentioned above was detected in non-substrate-amended or heat-killed controls. Cyclohexane carboxylic acid accumulated to a concentration of 260 microM, accounting for about 18% of the initial benzoate added. This compound was not detected in culture extracts of Rhodopseudomonas palustris grown phototrophically or Thauera aromatica grown under nitrate-reducing conditions. Cocultures of "S. aciditrophicus" and Methanospirillum hungatei readily metabolized cyclohexane carboxylate and cyclohex-1-ene carboxylate at a rate slightly faster than the rate of benzoate metabolism. In addition to cyclohexane carboxylate, pimelate, and glutarate, 2-hydroxycyclohexane carboxylate was detected in trace amounts in cocultures grown with cyclohex-1-ene carboxylate. Cyclohex-1-ene carboxylate, pimelate, and glutarate were detected in cocultures grown with cyclohexane carboxylate at levels similar to those found in benzoate-grown cocultures. Cell extracts of "S. aciditrophicus" grown in a coculture with Desulfovibrio sp. strain G11 with benzoate or in a pure culture with crotonate contained the following enzyme activities: an ATP-dependent benzoyl-CoA ligase, cyclohex-1-ene carboxyl-CoA hydratase, and 2-hydroxycyclohexane carboxyl-CoA dehydrogenase, as well as pimelyl-CoA dehydrogenase, glutaryl-CoA dehydrogenase, and the enzymes required for conversion of crotonyl-CoA to acetate. 2-Ketocyclohexane carboxyl-CoA hydrolase activity was detected in cell extracts of "S. aciditrophicus"-Desulfovibrio sp. strain G11 benzoate-grown cocultures but not in crotonate-grown pure cultures of "S. aciditrophicus". These results are consistent with the hypothesis that ring reduction during syntrophic benzoate metabolism involves a four- or six-electron reduction step and that once cyclohex-1-ene carboxyl-CoA is made, it is metabolized in a manner similar to that in R. palustris.  相似文献   

2.
The metabolism of benzoate, cyclohex-1-ene carboxylate, and cyclohexane carboxylate by “Syntrophus aciditrophicus” in cocultures with hydrogen-using microorganisms was studied. Cyclohexane carboxylate, cyclohex-1-ene carboxylate, pimelate, and glutarate (or their coenzyme A [CoA] derivatives) transiently accumulated during growth with benzoate. Identification was based on comparison of retention times and mass spectra of trimethylsilyl derivatives to the retention times and mass spectra of authentic chemical standards. 13C nuclear magnetic resonance spectroscopy confirmed that cyclohexane carboxylate and cyclohex-1-ene carboxylate were produced from [ring-13C6]benzoate. None of the metabolites mentioned above was detected in non-substrate-amended or heat-killed controls. Cyclohexane carboxylic acid accumulated to a concentration of 260 μM, accounting for about 18% of the initial benzoate added. This compound was not detected in culture extracts of Rhodopseudomonas palustris grown phototrophically or Thauera aromatica grown under nitrate-reducing conditions. Cocultures of “S. aciditrophicus” and Methanospirillum hungatei readily metabolized cyclohexane carboxylate and cyclohex-1-ene carboxylate at a rate slightly faster than the rate of benzoate metabolism. In addition to cyclohexane carboxylate, pimelate, and glutarate, 2-hydroxycyclohexane carboxylate was detected in trace amounts in cocultures grown with cyclohex-1-ene carboxylate. Cyclohex-1-ene carboxylate, pimelate, and glutarate were detected in cocultures grown with cyclohexane carboxylate at levels similar to those found in benzoate-grown cocultures. Cell extracts of “S. aciditrophicus” grown in a coculture with Desulfovibrio sp. strain G11 with benzoate or in a pure culture with crotonate contained the following enzyme activities: an ATP-dependent benzoyl-CoA ligase, cyclohex-1-ene carboxyl-CoA hydratase, and 2-hydroxycyclohexane carboxyl-CoA dehydrogenase, as well as pimelyl-CoA dehydrogenase, glutaryl-CoA dehydrogenase, and the enzymes required for conversion of crotonyl-CoA to acetate. 2-Ketocyclohexane carboxyl-CoA hydrolase activity was detected in cell extracts of “S. aciditrophicus”-Desulfovibrio sp. strain G11 benzoate-grown cocultures but not in crotonate-grown pure cultures of “S. aciditrophicus”. These results are consistent with the hypothesis that ring reduction during syntrophic benzoate metabolism involves a four- or six-electron reduction step and that once cyclohex-1-ene carboxyl-CoA is made, it is metabolized in a manner similar to that in R. palustris.  相似文献   

3.
The aerobic benzoate metabolism in Azoarcus evansii follows an unusual route. The intermediates of the pathway are processed as coenzyme A (CoA) thioesters and the cleavage of the aromatic ring is non-oxygenolytic. The enzymes of this pathway are encoded by the box gene cluster which harbors a gene, orf1, coding for a putative thioesterase. Benzoyl-CoA thioesterase activity (20 nmol min−1 mg−1 protein) was present in cells grown aerobically on benzoate, but was lacking in cells grown on other aromatic or aliphatic substrates under oxic or anoxic conditions. The gene was cloned and overexpressed in Escherichia coli to produce a C-terminal His-tag fusion protein. The recombinant enzyme was a homotetramer of 16 kDa subunits. It catalyzed not only the hydrolysis of benzoyl-CoA, but also of 2,3-dihydro-2,3-dihydroxybenzoyl-CoA, the second intermediate in the pathway. The enzyme exhibited higher activity with mono-substituted derivatives of benzoyl-CoA, showing highest activity with 4-hydroxybenzoyl-CoA. Di-substituted derivatives of benzoyl-CoA, phenylacetyl-CoA, and aliphatic CoA thioesters were not hydrolyzed but some acted as inhibitors. The thioesterase appears to protect the cell from CoA pool depletion. It may constitute the prototype of a new subfamily within the hotdog fold enzyme superfamily.  相似文献   

4.
The sulfate-reducing bacterium Desulfococcus multivorans uses various aromatic compounds as sources of cell carbon and energy. In this work, we studied the initial steps in the aromatic metabolism of this strictly anaerobic model organism. An ATP-dependent benzoate coenzyme A (CoA) ligase (AMP plus PPi forming) composed of a single 59-kDa subunit was purified from extracts of cells grown on benzoate. Specific activity was highest with benzoate and some benzoate derivatives, whereas aliphatic carboxylic acids were virtually unconverted. The N-terminal amino acid sequence showed high similarities with benzoate CoA ligases from Thauera aromatica and Azoarcus evansii. When cultivated on benzoate, cells strictly required selenium and molybdenum, whereas growth on nonaromatic compounds, such as cyclohexanecarboxylate or lactate, did not depend on the presence of the two trace elements. The growth rate on benzoate was half maximal with 1 nM selenite present in the growth medium. In molybdenum- and/or selenium-depleted cultures, growth on benzoate could be induced by addition of the missing trace elements. In extracts of cells grown on benzoate in the presence of [75Se]selenite, three radioactively labeled proteins with molecular masses of approximately 100, 30, and 27 kDa were detected by sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis. The 100- and 30-kDa selenoproteins were 5- to 10-fold induced in cells grown on benzoate compared to cells grown on lactate. These results suggest that the dearomatization process in D. multivorans is not catalyzed by the ATP-dependent Fe-S enzyme benzoyl-CoA reductase as in facultative anaerobes but rather involves unknown molybdenum- and selenocysteine-containing proteins.  相似文献   

5.
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   

6.
Thirty-three microorganisms capable of growth with cyclohexane carboxylate as the sole source of carbon were isolated from mud, water, and soil samples from the Aberystwyth area. Preliminary screening and whole-cell oxidation studies suggested that, with one exception, all of the strains metabolized the growth substrate by beta-oxidation of the coenzyme A ester. This single distinctive strain, able to oxidize rapidly trans-4-hydroxycyclohexane carboxylate, 4-ketocyclohexane carboxylate, p-hydroxybenzoate, and protocatechuate when grown with cyclohexane carboxylate, was classified as a strain of Alcaligenes and given the number W1. Enzymes capable of converting cyclohexane carboxylate to p-hydroxybenzoate were induced by growth with the alicyclic acid and included the first unambiguous specimen of a cyclohexane carboxylate hydroxylase. Because it is a very fragile protein, attempts to stabilize the cyclohexane carboxylate hydroxylase so that a purification procedure could be developed have consistently failed. In limited studies with crude cell extracts, we found that hydroxylation occurred at the 4 position, probably yielding the trans isomer of 4-hydroxycyclohexane carboxylate. Simultaneous measurement of oxygen consumption and reduced nicotinamide adenine dinucleotide oxidation, coupled with an assessment of reactant stoichiometry, showed the enzyme to be a mixed-function oxygenase. Mass spectral analysis enabled the conversion of cyclohexane carboxylate to p-hydroxybenzoate by cell extracts to be established unequivocally, and all of our data were consistent with the pathway: cyclohexane carboxylate --> trans-4-hydroxycyclohexane carboxylate --> 4-ketocyclohexane carboxylate --> p-hydroxybenzoate. The further metabolism of p-hydroxybenzoate proceeded by meta fission and by the oxidative branch of the 2-hydroxy-4-carboxymuconic semialde-hyde-cleaving pathway.  相似文献   

7.
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   

8.
Anaerobic benzoate degradation by the phototrophic bacterium Rhodopseudomonas palustris has been proposed to proceed via aromatic ring reduction reactions leading to cyclohex-1-ene-1-carboxyl-coenzyme A (CoA) formation. The alicyclic product is then proposed to undergo three β-oxidation-like modifications resulting in ring cleavage. Illuminated suspensions of benzoate-grown cells converted [7-14C]cyclohex-1-ene-1-carboxylate to intermediates that comigrated with cyclohex-1-ene-1-carboxyl-CoA, 2-hydroxycyclohexanecar-boxyl-CoA, 2-ketocyclohexanecarboxyl-CoA, and pimelyl-CoA by thin-layer chromatography. This set of intermediates was also formed by cells grown anaerobically or aerobically on cyclohex-1-ene-1-carboxylate, indicating that benzoate-grown and cyclohex-1-ene-1-carboxylate-grown cells degrade this alicyclic acid by the same catabolic route. Four enzymatic activities proposed to be required for conversion of cyclohex-1-ene-1-carboxylate to pimelyl-CoA were detected at 3- to 10-fold-higher levels in benzoate-grown cells than in succinate-grown cells. These were cyclohex-1-ene-1-carboxylate-CoA ligase, cyclohex-1-ene-1-carboxyl-CoA hydratase, 2-hydroxycyclohexanecarboxyl-CoA dehydrogenase, and 2-ketocyclohexanecarboxyl-CoA hydrolase (ring cleaving). Pimelyl-CoA was identified in hydrolase reaction mixtures as the product of alicyclic ring cleavage. The results provide a first demonstration of an alicyclic ring cleavage activity.  相似文献   

9.
Syntrophy is essential for the efficient conversion of organic carbon to methane in natural and constructed environments, but little is known about the enzymes involved in syntrophic carbon and electron flow. Syntrophus aciditrophicus strain SB syntrophically degrades benzoate and cyclohexane-1-carboxylate and catalyses the novel synthesis of benzoate and cyclohexane-1-carboxylate from crotonate. We used proteomic, biochemical and metabolomic approaches to determine what enzymes are used for fatty, aromatic and alicyclic acid degradation versus for benzoate and cyclohexane-1-carboxylate synthesis. Enzymes involved in the metabolism of cyclohex-1,5-diene carboxyl-CoA to acetyl-CoA were in high abundance in S. aciditrophicus cells grown in pure culture on crotonate and in coculture with Methanospirillum hungatei on crotonate, benzoate or cyclohexane-1-carboxylate. Incorporation of 13C-atoms from 1-[13C]-acetate into crotonate, benzoate and cyclohexane-1-carboxylate during growth on these different substrates showed that the pathways are reversible. A protein conduit for syntrophic reverse electron transfer from acyl-CoA intermediates to formate was detected. Ligases and membrane-bound pyrophosphatases make pyrophosphate needed for the synthesis of ATP by an acetyl-CoA synthetase. Syntrophus aciditrophicus, thus, uses a core set of enzymes that operates close to thermodynamic equilibrium to conserve energy in a novel and highly efficient manner.  相似文献   

10.
Cell-free extracts of Syntrophomonas wolfei subsp. wolfei grown with crotonate in pure culture or in coculture with Methanospirillum hungatei contained crotonyl-coenzyme A (CoA): acetate CoA-transferase activity. This activity was not detected in cell-free extracts from the butyrate-grown coculture which suggests that the long lag times observed before S. wolfei grew with crotonate were initially due to the inability to activate crotonate. Cell-free extracts of S. wolfei grown in pure culture contained high specific activities of hydrogenase and very low levels of formate dehydrogenase. The low levels suggest a biosynthethic rather than a catabolic role for the latter enzyme when S. wolfei is grown in pure culture. CO dehydrogenase activity was not detected. S. wolfei can form butyrate using a CoA transferase activity, but not by a phosphotransbutyrylase or enoate reductase activity. A c-type cytochrome was detected in S. wolfei grown in pure culture or in coculture indicating the presence of an electron transport system. This is a characteristic which separates S. wolfei from other known crotonate-using bacteria.  相似文献   

11.
Cell-free extracts of Pseudomonas sp. strains KB 740 and K 172 both contained high levels of glutaryl-CoA dehydrogenase when grown anaerobically on benzoate or other aromatic compounds and with nitrate as electron acceptor. These aromatic compounds have in common benzoyl-CoA as the central aromatic intermediate of anerobic metabolism. The enzymatic activity was almost absent in cells grown aerobically on benzoate regardless whether nitrate was present. Glutaryl-CoA dehydrogenase activity was also detected in cell-free extracts of Rhodopseudomonas, Rhodomicrobium and Rhodocyclus after phototrophic growth on benzoate. Parallel to the induction of glutaryl-CoA dehydrogenase as measured with ferricenium ion as electron acceptor, an about equally high glutaconyl-CoA decarboxylase activity was detected in cell-free extracts. The latter activity was measured with the NAD-dependent assay, as described for the biotin-containing sodium ion pump glutaconyl-CoA decarboxylase from glutamate fermenting bacteria. Glutaryl-CoA dehydrogenase was purified to homogeneity from both Pseudomonas strains. The enzymes catalyse the decarboxylation of glutaconyl-CoA at about the same rate as the oxidative decarboxylation of glutaryl-CoA. The green enzymes are homotetramers (m=170 kDa) and contain 1 mol FAD per subunit. No inhibition was observed with avidin indicating the absence of biotin. The N-terminal sequences of the enzymes from both strains are similar (65%).  相似文献   

12.
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   

13.
The strictly anaerobic Syntrophus aciditrophicus is a fermenting deltaproteobacterium that is able to degrade benzoate or crotonate in the presence and in the absence of a hydrogen-consuming partner. During growth in pure culture, both substrates are dismutated to acetate and cyclohexane carboxylate. In this work, the unknown enzymes involved in the late steps of cyclohexane carboxylate formation were studied. Using enzyme assays monitoring the oxidative direction, a cyclohex-1-ene-1-carboxyl-CoA (Ch1CoA)-forming cyclohexanecarboxyl-CoA (ChCoA) dehydrogenase was purified and characterized from S. aciditrophicus and after heterologous expression of its gene in Escherichia coli. In addition, a cyclohexa-1,5-diene-1-carboxyl-CoA (Ch1,5CoA)-forming Ch1CoA dehydrogenase was characterized after purification of the heterologously expressed gene. Both enzymes had a native molecular mass of 150 kDa and were composed of a single, 40- to 45-kDa subunit; both contained flavin adenine dinucleotide (FAD) as a cofactor. While the ChCoA dehydrogenase was competitively inhibited by Ch1CoA in the oxidative direction, Ch1CoA dehydrogenase further converted the product Ch1,5CoA to benzoyl-CoA. The results obtained suggest that Ch1,5CoA is a common intermediate in benzoate and crotonate fermentation that serves as an electron-accepting substrate for the two consecutively operating acyl-CoA dehydrogenases characterized in this work. In the case of benzoate fermentation, Ch1,5CoA is formed by a class II benzoyl-CoA reductase; in the case of crotonate fermentation, Ch1,5CoA is formed by reversing the reactions of the benzoyl-CoA degradation pathway that are also employed during the oxidative (degradative) branch of benzoate fermentation.  相似文献   

14.
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   

15.
3-Hydroxybenzoate:coenzyme A ligase, an enzyme involved in xanthone biosynthesis, was detected in cell-free extracts from cultured cells of Centaurium erythraea Rafn. The enzyme was separated from 4-coumarate:coenzyme A ligase by fractionated ammonium sulphate precipitation and hydrophobic interaction chromatography. The CoA ligases exhibited different substrate specificities. 3-Hydroxybenzoate:coenzyme A ligase activated 3-hydroxybenzoic acid most efficiently and lacked affinity for cinnamic acids. In contrast, 4-coumarate:CoA ligase mainly catalyzed the activation of 4-coumaric acid but did not act on benzoic acids. The two enzymes were similar with respect to their relative molecular weight, their pH and temperature optima, their specific activity and the changes in their activity during cell culture growth. Received: 23 September 1996 / Accepted: 28 November 1996  相似文献   

16.
The anaerobic metabolism of 2-hydroxybenzoic acid (salicylic acid) was studied in a denitrifying bacterium. Cells grown with 2-hydroxybenzoate were simultaneously adapted to degrade benzoate. Extract of these cells formed benzoate or benzoyl-CoA when incubated under reducing conditions with salicylate, MgATP, and coenzyme A, suggesting a degradation of 2-hydroxybenzoate via benzoate or benzoyl-CoA. This suggestion was supported by enzyme activity measurements. In extracts of 2-hydroxybenzoate-grown cells, the following enzyme activities were detected: two CoA ligases, one specific for 2-hydroxybenzoate, the other for benzoate, and two different enzyme activities catalyzing the reductive transformation of 2-hydroxybenzoyl-CoA. These findings suggest a degradation of salicylic acid by two new enzymes, 2-hydroxybenzoate-CoA ligase (AMP-forming) and 2-hydroxybenzoyl-CoA reductase (dehydroxylating), catalyzing (1) 2-hydroxybenzoate + MgATP + CoASH → 2-hydroxybenzoyl-CoA + MgAMP + PPi (2) 2-hydroxybenzoyl-CoA + 2[H] → benzoyl-CoA + H2O Benzoyl-CoA was dearomatized by reduction of the ring. This represents another case in which benzoyl-CoA is a central intermediate in anaerobic aromatic metabolism. Received: 1 February 1996 / Accepted: 24 February 1996  相似文献   

17.
Peroxisomal β‐oxidative degradation of compounds is a common metabolic process in eukaryotes. Reported benzoyl‐coenzyme A (BA‐CoA) thioesterase activity in peroxisomes from petunia flowers suggests that, like mammals and fungi, plants contain auxiliary enzymes mediating β‐oxidation. Here we report the identification of Petunia hybrida thioesterase 1 (PhTE1), which catalyzes the hydrolysis of aromatic acyl‐CoAs to their corresponding acids in peroxisomes. PhTE1 expression is spatially, developmentally and temporally regulated and exhibits a similar pattern to known benzenoid metabolic genes. PhTE1 activity is inhibited by free coenzyme A (CoA), indicating that PhTE1 is regulated by the peroxisomal CoA pool. PhTE1 downregulation in petunia flowers led to accumulation of BA‐CoA with increased production of benzylbenzoate and phenylethylbenzoate, two compounds which rely on the presence of BA‐CoA precursor in the cytoplasm, suggesting that acyl‐CoAs can be exported from peroxisomes. Furthermore, PhTE1 downregulation resulted in increased pools of cytoplasmic phenylpropanoid pathway intermediates, volatile phenylpropenes, lignin and anthocyanins. These results indicate that PhTE1 influences (i) intraperoxisomal acyl‐CoA/CoA levels needed to carry out β‐oxidation, (ii) efflux of β‐oxidative products, acyl‐CoAs and free acids, from peroxisomes, and (iii) flux distribution within the benzenoid/phenylpropanoid metabolic network. Thus, this demonstrates that plant thioesterases play multiple auxiliary roles in peroxisomal β‐oxidative metabolism.  相似文献   

18.
Glutathione thiol esterase activity in cell extracts of a yeast: Saccharomyces cerevisiae was separated into three peaks when filtered on a Sephadex G-150 gel column. One of the enzymes in these peaks was purified. The enzyme was a single polypeptide chain with a molecular weight of 28,000 and catalyzed the complete hydrolysis of S-acetylglutathione and S-lactoylglutathione. S- Methyl-, S-hexyl-, S-glyceryl-, S-succinylglutathiones, and acetyl CoA were not hydrolyzed. In addition to the hydrolytic activity, the purified enzyme showed a group transfer activity and catalyzed the formation of acetyl CoA from S-acetylglutathione and CoA. The purified enzyme was not identical with glyoxalase II in molecular weight, substrate specificity, or behaviors toward inhibitors.  相似文献   

19.
Organohalides are environmentally relevant compounds that can be degraded by aerobic and anaerobic microorganisms. The denitrifying Thauera chlorobenzoica is capable of degrading halobenzoates as sole carbon and energy source under anaerobic conditions. LC‐MS/MS‐based coenzyme A (CoA) thioester analysis revealed that 3‐chloro‐ or 3‐bromobenzoate were preferentially metabolized via non‐halogenated CoA‐ester intermediates of the benzoyl‐CoA degradation pathway. In contrast, 3‐fluorobenzoate, which does not support growth, was converted to dearomatized fluorinated CoA ester dead‐end products. Extracts from cells grown on 3‐chloro‐/3‐bromobenzoate catalysed the Ti(III)‐citrate‐ and ATP‐dependent reductive dehalogenation of 3‐chloro/3‐bromobenzoyl‐CoA to benzoyl‐CoA, whereas 3‐fluorobenzoyl‐CoA was converted to a fluorinated cyclic dienoyl‐CoA compound. The reductive dehalogenation reactions were identified as previously unknown activities of ATP‐dependent class I benzoyl‐CoA reductases (BCR) present in all facultatively anaerobic, aromatic compound degrading bacteria. A two‐step dearomatization/H‐halide elimination mechanism is proposed. A halobenzoate‐specific carboxylic acid CoA ligase was characterized in T. chlorobenzoica; however, no such enzyme is present in Thauera aromatica, which cannot grow on halobenzoates. In conclusion, it appears that the presence of a halobenzoate‐specific carboxylic acid CoA ligase rather than a specific reductive dehalogenase governs whether an aromatic compound degrading anaerobe is capable of metabolizing halobenzoates.  相似文献   

20.
Ruegeria pomeroyi DSS‐3 possesses two general pathways for metabolism of dimethylsulphoniopropionate (DMSP), an osmolyte of algae and abundant carbon source for marine bacteria. In the DMSP cleavage pathway, acrylate is transformed into acryloyl‐CoA by propionate‐CoA ligase (SPO2934) and other unidentified acyl‐CoA ligases. Acryloyl‐CoA is then reduced to propionyl‐CoA by AcuI or SPO1914. Acryloyl‐CoA is also rapidly hydrated to 3‐hydroxypropionyl‐CoA by acryloyl‐CoA hydratase (SPO0147). A SPO1914 mutant was unable to grow on acrylate as the sole carbon source, supporting its role in this pathway. Similarly, growth on methylmercaptopropionate, the first intermediate of the DMSP demethylation pathway, was severely inhibited by a mutation in the gene encoding crotonyl‐CoA carboxylase/reductase, demonstrating that acetate produced by this pathway was metabolized by the ethylmalonyl‐CoA pathway. Amino acids and nucleosides from cells grown on 13C‐enriched DMSP possessed labelling patterns that were consistent with carbon from DMSP being metabolized by both the ethylmalonyl‐CoA and acrylate pathways as well as a role for pyruvate dehydrogenase. This latter conclusion was supported by the phenotype of a pdh mutant, which grew poorly on electron‐rich substrates. Additionally, label from [13C‐methyl] DMSP only appeared in carbons derived from methyl‐tetrahydrofolate, and there was no evidence for a serine cycle of C‐1 assimilation.  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