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.
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   

2.
Photoheterotrophic metabolism of two meta-hydroxy-aromatic acids, meta-, para-dihydroxybenzoate (protocatechuate) and meta-hydroxybenzoate, was investigated in Rhodopseudomonas palustris. When protocatechuate was the sole organic carbon source, photoheterotrophic growth in R. palustris was slow relative to cells using compounds known to be metabolized by the benzoyl coenzyme A (benzoyl-CoA) pathway. R. palustris was unable to grow when meta-hydroxybenzoate was provided as a sole source of organic carbon under photoheterotrophic growth conditions. However, in cultures supplemented with known benzoyl-CoA pathway inducers (para-hydroxybenzoate, benzoate, or cyclohexanoate), protocatechuate and meta-hydroxybenzoate were taken up from the culture medium. Further, protocatechuate and meta-hydroxybenzoate were each removed from cultures containing both meta-hydroxy-aromatic acids at equimolar concentrations in the absence of other organic compounds. Analysis of changes in culture optical density and in the concentration of soluble organic compounds indicated that the loss of these meta-hydroxy-aromatic acids was accompanied by biomass production. Additional experiments with defined mutants demonstrated that enzymes known to participate in the dehydroxylation of para-hydroxybenzoyl-CoA (HbaBCD) and reductive dearomatization of benzoyl-CoA (BadDEFG) were required for metabolism of protocatechuate and meta-hydroxybenzoate. These findings indicate that, under photoheterotrophic growth conditions, R. palustris can degrade meta-hydroxy-aromatic acids via the benzoyl-CoA pathway, apparently due to the promiscuity of the enzymes involved.  相似文献   

3.
The presence or absence of molecular oxygen has been shown to play a crucial role in the degradability of haloaromatic compounds. In the present study, it was shown that anaerobic phototrophic 3-chlorobenzoate (3CBA) metabolism by Rhodopseudomonas palustris DCP3 is oxygen tolerant up to a concentration of 3 μM O2. Simultaneous oxidation of an additional carbon source permitted light-dependent anaerobic 3CBA degradation at oxygen input levels which, in the absence of such an additional compound, would result in inhibition of light-dependent dehalogenation. Experiments under the same experimental conditions with strain DCP3 in coculture with an aerobic 3CBA-utilizing heterotroph, Alcaligenes sp. strain L6, revealed that light-dependent dehalogenation of 3CBA did not occur. Under both oxygen limitation (O2 < 0.1 μM) and low oxygen concentrations (3 μM O2), all the 3CBA was metabolized by the aerobic heterotroph. These data suggest that biodegradation of (halo)aromatics by photoheterotrophic bacteria such as R. palustris DCP3 may be restricted to anoxic photic environments.  相似文献   

4.
The biochemical effects of aryl substituents on the reductive dechlorination of 3-chlorobenzoate analogs were quantified with (i) a stable 3-chlorobenzoate-grown methanogenic sludge enrichment, (ii) Desulfomonile tiedjei DCB-1, isolated from this enrichment and able to catalyze the reductive dechlorination of 3-chlorobenzoate, and (iii) a defined 3-chlorobenzoate-degrading methanogenic consortium with D. tiedjei as the key dechlorinating organism. The addition of hydrogen stimulated the dechlorination rate in the consortium. The extent of this stimulation depended on the substituent. The data were evaluated with various sets of substituent constants compiled for the Hammett equation. None of the sets yielded a satisfactory correlation between experimental values and theoretical constants. This suggests that the microbially catalyzed reductive dechlorination of 3-chlorobenzoate cannot be described simply as either a nucleophilic or an electrophilic substitution reaction. Nevertheless, observations that the presence of a para-amino or -hydroxy group inhibited the rate of dechlorination suggest that the rate-limiting step in the reductive dechlorination of 3-chlorobenzoate is a nucleophilic attack on the negatively charged π electron cloud around the benzene nucleus.  相似文献   

5.
6.
Benzoyl coenzyme A (benzoyl-CoA) reductase is a central enzyme in the anaerobic degradation of organic carbon, which utilizes a common intermediate (benzoyl-CoA) in the metabolism of many aromatic compounds. The diversity of benzoyl-CoA reductase genes in denitrifying bacterial isolates capable of degrading aromatic compounds and in river and estuarine sediment samples from the Arthur Kill in New Jersey and the Chesapeake Bay in Maryland was investigated. Degenerate primers were developed from the known benzoyl-CoA reductase genes from Thauera aromatica, Rhodopseudomonas palustris, and Azoarcus evansii. PCR amplification detected benzoyl-CoA reductase genes in the denitrifying isolates belonging to α-, β-, or γ-Proteobacteria as well as in the sediment samples. Phylogenetic analysis, sequence similarity comparison, and conserved indel determination grouped the new sequences into either the bcr type (found in T. aromatica and R. palustris) or the bzd type (found in A. evansii). All the Thauera strains and the isolates from the genera Acidovorax, Bradyrhizobium, Paracoccus, Ensifer, and Pseudomonas had bcr-type benzoyl-CoA reductases with amino acid sequence similarities of more than 97%. The genes detected from Azarocus strains were assigned to the bzd type. A total of 50 environmental clones were detected from denitrifying consortium and sediment samples, and 28 clones were assigned to either the bcr or the bzd type of benzoyl-CoA reductase genes. Thus, we could determine the genetic capabilities for anaerobic degradation of aromatic compounds in sediment communities of the Chesapeake Bay and the Arthur Kill on the basis of the detection of two types of benzoyl-CoA reductase genes. The detected genes have future applications as genetic markers to monitor aromatic compound degradation in natural and engineered ecosystems.  相似文献   

7.
8.
In the aerobic metabolism of aromatic substrates, oxygenases use molecular oxygen to hydroxylate and finally cleave the aromatic ring. In the case of the common intermediate benzoate, the ring cleavage substrates are either catechol (in bacteria) or 3,4-dihydroxybenzoate (protocatechuate, mainly in fungi). We have shown before that many bacteria, e.g. Azoarcus evansii, the organism studied here, use a completely different mechanism. This elaborate pathway requires formation of benzoyl-CoA, followed by an oxygenase reaction and a nonoxygenolytic ring cleavage. Benzoyl-CoA transformation is catalyzed by the iron-containing benzoyl-CoA oxygenase (BoxB) in conjunction with an FAD and iron-sulfur centers containing reductase (BoxA), which donates electrons from NADPH. Here we show that benzoyl-CoA oxygenase actually does not form the 2,3-dihydrodiol of benzoyl-CoA, as formerly postulated, but the 2,3-epoxide. An enoyl-CoA hydratase (BoxC) uses two molecules of water to first hydrolytically open the ring of 2,3-epoxybenzoyl-CoA, which may proceed via its tautomeric seven-membered oxepin ring form. Then ring C2 is hydrolyzed off as formic acid, yielding 3,4-dehydroadipyl-CoA semialdehyde. The semialdehyde is oxidized by a NADP+-dependent aldehyde dehydrogenase (BoxD) to 3,4-dehydroadipyl-CoA. Final products of the pathway are formic acid, acetyl-CoA, and succinyl-CoA. This overlooked pathway occurs in 4–5% of all bacteria whose genomes have been sequenced and represents an elegant strategy to cope with the high resonance energy of aromatic substrates by forming a nonaromatic epoxide.  相似文献   

9.
10.
The biodegradation of lignin, one of the most abundant carbon compounds on Earth, has important biotechnological applications in the derivation of useful products from lignocellulosic wastes. The purple photosynthetic bacterium Rhodopseudomonas palustris is able to grow photoheterotrophically under anaerobic conditions on a range of phenylpropeneoid lignin monomers, including coumarate, ferulate, caffeate, and cinnamate. RPA1789 (CouP) is the periplasmic binding-protein component of an ABC system (CouPSTU; RPA1789, RPA1791–1793), which has previously been implicated in the active transport of this class of aromatic substrate. Here, we show using both intrinsic tryptophan fluorescence and isothermal titration calorimetry that CouP binds a range of phenylpropeneoid ligands with K d values in the nanomolar range. The crystal structure of CouP with ferulate as the bound ligand shows H-bond interactions between the 4-OH group of the aromatic ring with His309 and Gln305. H-bonds are also made between the carboxyl group on the ferulate side chain and Arg197, Ser222, and Thr102. An additional transport system (TarPQM; RPA1782–1784), a member of the tripartite ATP-independent periplasmic (TRAP) transporter family, is encoded at the same locus as rpa1789 and several other genes involved in coumarate metabolism. We show that the periplasmic binding-protein of this system (TarP; RPA1782) also binds coumarate, ferulate, caffeate, and cinnamate with nanomolar K d values. Thus, we conclude that R. palustris uses two redundant but energetically distinct primary and secondary transporters that both employ high-affinity periplasmic binding-proteins to maximise the uptake of lignin-derived aromatic substrates from the environment. Our data provide a detailed thermodynamic and structural basis for understanding the interaction of lignin-derived aromatic substrates with proteins and will be of use in the further exploitation of the flexible metabolism of R. palustris for anaerobic aromatic biotransformations.  相似文献   

11.
12.
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   

13.
Desulfomonile tiedjei is the key dechlorinating organism in a three-tiered bacterial consortium that grows on the methanogenic degradation of 3-chlorobenzoate. 2,5-Dichlorobenzoate, however, is only converted to 2-chlorobenzoate and is not a methanogenic substrate for the consortium. The dechlorinator uses hydrogen produced from benzoate by the benzoate degrading member of consortium as its source of reducing equivalents for the dechlorination reaction. Incubation of 3-chlorobenzoate grown consortium cells with 2,5-dichlorobenzoate resulted in the consumption of acetate concurrent with the formation of 2-chlorobenzoate indicating that acetate can serve as an alternative source of reducing equivalents for reductive dechlorination. This interpretation was confirmed by the finding that the formation of 14CO2 from 2-14C-labeled acetate was stoichiometric. The addition of hydrogen to 2,5-dichlorobenzoate metabolizing cells resulted in (i) an 2.7-fold increase in the rate of dechlorination, and (ii) a drop in the amount of label recovered as CO2+CH4 from methyl 14C-labeled acetate, indicating that hydrogen was the preferred source of reducing equivalents for reductive dechlorination. Benzoate, an indirect source of H2 in the consortium, also inhibited the oxidation of acetate, while glucose, methanol, and butyrate did not affect labeled gas production and therefore were not suitable electron donors. Concomittant to dechlorination of 2,5-dichlorobenzoate 3- and 4-methoxybenzoate were converted to 3- and 4-hydroxybenzoate respectively. These conversions stimulated the rate of dechlorination 2-fold. Demethylation of 4-methoxybenzoate stimulated, but demethylation of 3-methoxybenzoate inhibited the oxidation of benzoate during the dechlorination of 2,5-dichlorobenzoate, suggesting that these isomers are metabolized through different pathways. Experiments with benzoate, 3-chlorobenzoate and 2,5-dichlorobenzoate metabolizing cells amended with 14CO2 showed that actively dechlorinating cells catalyzed an exchange reaction between CO2 and acetate.  相似文献   

14.
Desulfomonile tiedjei, a strict anaerobe capable of reductively dechlorinating 3-chlorobenzoate, also dechlorinates tetrachloroethene and trichloroethene. It is not known, however, if the aryl and aliphatic dechlorination activities are catalyzed by the same enzymatic system. Cultures induced for 3-chlorobenzoate activity dechlorinated tetrachloroethene and trichloroethene to lower chlorinated products while uninduced parallel cultures did not dechlorinate either substrate. The observed rate of PCE dechlorination in induced cultures was 22 µmol h–1 g protein–1, which is considerably faster than previous rates obtained with defined cultures of this organism. These results show that both dechlorination activities are co-induced and therefore, that the dechlorination mechanisms may share at least some components.Abbreviations PCE tetrachloroethene - TCE trichloroethene - cis-DCE cis-dichloroethene - trans-DCE trans-dichloroethene - 3FBz 3-fluorobenzoate - 3ClBz 3-chlorobenzoate  相似文献   

15.
Summary The benzoyl-CoA ligase from an anaerobic syntrophic culture was purified to homogeneity. It had a molecular mass of around 420 kDa and consisted of seven or eight subunits of 58 kDa. The temperature optimum was 37–40° C, the optimum pH around 8.0 and optimal activity required 50–100 mM TRIS-HCI buffer, pH 8.0 and 3–7 mM MgCl2; MgCl2 in excess of 10 mM was inhibitory. The activation energy for benzoate was 11.3 kcal/mol. Although growth occured only with benzoate as a carbon source, the benzoyl-coenzyme A (CoA) ligase formed benzoyl-CoA esters with benzoate, 2-, 3- and 4-fluorobenzoate, picolinate, nicotinate and isonicotinate. Acetate was activated to acetyl-CoA by an acetyl-CoA synthetase. The K m values for benzoate, 2-, 3- and 4-fluorobenzoate were 0.04, 0.28, 1.48 and 0.32 mM, the V max values 1.05, 1.0, 0.7 and 0.98 units (U)/mg, respectively. For reduced CoA (CoA-SH) a K m of 0.17 mM and a V max of 1.05 U/mg and for ATP a K m of 0.16 mM and a V max of 1.08 U/mg was determined. Benzoate activation was inhibited by more than 6 mM ATP, presumably by pyrophosphate generation from ATP. The inhibition constant (K i) for pyrophosphate was 5.7 mM. No homology of the N-terminal amino acid sequence with that of a 2-aminobenzoyl-CoA ligase of a denitrifying Pseudomonas sp. was found. Correspondence to: J. Winter  相似文献   

16.
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   

17.
Strain SF3, a gram-negative, anaerobic, motile, short curved rod that grows by coupling the reductive dechlorination of 2-chlorophenol (2-CP) to the oxidation of acetate, was isolated from San Francisco Bay sediment. Strain SF3 grew at concentrations of NaCl ranging from 0.16 to 2.5%, but concentrations of KCl above 0.32% inhibited growth. The isolate used acetate, fumarate, lactate, propionate, pyruvate, alanine, and ethanol as electron donors for growth coupled to reductive dechlorination. Among the halogenated aromatic compounds tested, only the ortho position of chlorophenols was reductively dechlorinated, and additional chlorines at other positions blocked ortho dechlorination. Sulfate, sulfite, thiosulfate, and nitrate were also used as electron acceptors for growth. The optimal temperature for growth was 30°C, and no growth or dechlorination activity was observed at 37°C. Growth by reductive dechlorination was revealed by a growth yield of about 1 g of protein per mol of 2-CP dechlorinated, and about 2.7 g of protein per mole of 2,6-dichlorophenol dechlorinated. The physiological features and 16S ribosomal DNA sequence suggest that the organism is a novel species of the genus Desulfovibrio and which we have designated Desulfovibrio dechloracetivorans. The unusual physiological feature of this strain is that it uses acetate as an electron donor and carbon source for growth with 2-CP but not with sulfate.  相似文献   

18.

Background and Aims

Plant species from various taxa ‘escape’ from low oxygen conditions associated with submergence by a suite of traits collectively called the low oxygen escape syndrome (LOES). The expression of these traits is associated with costs and benefits. Thus far, remarkably few studies have dealt with the expected benefits of the LOES.

Methods

Young plants were fully submerged at initial depths of 450 mm (deep) or 150–240 mm (shallow). Rumex palustris leaf tips emerged from the shallow flooding within a few days, whereas a slight lowering of shallow flooding was required to expose R. acetosa leaf tips to the atmosphere. Shoot biomass and petiole porosity were measured for all species, and treatments and data from the deep and shallow submergence treatments were compared with non-flooded controls.

Key Results

R. palustris is characterized by submergence-induced enhanced petiole elongation. R. acetosa lacked this growth response. Upon leaf tip emergence, R. palustris increased its biomass, whereas R. acetosa did not. Furthermore, petiole porosity in R. palustris was twice as high as in R. acetosa.

Conclusions

Leaf emergence restores gas exchange between roots and the atmosphere in R. palustris. This occurs to a much lesser extent in R. acetosa and is attributable to its lower petiole porosity and therefore limited internal gas transport. Leaf emergence resulting from fast petiole elongation appears to benefit biomass accumulation if these plants contain sufficient aerenchyma in petioles and roots to facilitate internal gas exchange.Key words: Submergence, emergence, enhanced shoot elongation, porosity, aerenchyma, Rumex, cost–benefit analysis, phenotypic plasticity  相似文献   

19.
20.
The anaerobic acetogenic bacterium Acetobacterium woodii couples reduction of caffeate with electrons derived from molecular hydrogen to the synthesis of ATP by a chemiosmotic mechanism with sodium ions as coupling ions. Caffeate is activated to caffeyl coenzyme A (caffeyl-CoA) prior to its reduction, and the caffeate reduction operon encodes an ATP-dependent caffeyl-CoA synthetase that is thought to catalyze the initial caffeate activation. The operon also encodes a potential CoA transferase, the product of carA, which was thought to be involved in subsequent ATP-independent caffeate activation. To prove the proposed function of carA, we overproduced its protein in Escherichia coli and then purified it. Purified CarA drives the formation of caffeyl-CoA from caffeate with hydrocaffeyl-CoA as the CoA donor. The dependence of the reaction on caffeate and hydrocaffeyl-CoA followed Michaelis-Menten kinetics, with apparent Km values of 75 ± 5 μM for caffeate and 8 ± 2 μM for hydrocaffeyl-CoA. The enzyme activity had broad ranges of pH and temperature optima. In addition to being able to use caffeate, CarA could use p-coumarate and ferulate but not cinnamate, sinapate, or p-hydroxybenzoate as a CoA acceptor. Neither acetyl-CoA nor butyryl-CoA served as the CoA donor for CarA. The enzyme uses a ping-pong mechanism for CoA transfer and is the first classified member of a new subclass of family I CoA transferases that has two catalytic domains on one polypeptide chain. Apparently, CarA catalyzes an energy-saving CoA loop for caffeate activation in the steady state of caffeate respiration.  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