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.
A denitrifying bacterium was isolated from a river sediment after enrichment on 3-chlorobenzoate under anoxic, denitrifying conditions. The bacterium, designated strain 3CB-1, degraded 3-chlorobenzoate, 3-bromobenzoate, and 3-iodobenzoate with stoichiometric release of halide under conditions supporting anaerobic growth by denitrification. The 3-halobenzoates and 3-hydroxybenzoate were used as growth substrates with nitrate as the terminal electron acceptor. The doubling time when growing on 3-halobenzoates ranged from 18 to 25 h. On agar plates with 1 mM 3-chlorobenzoate as the sole carbon source and 30 mM nitrate as the electron acceptor, strain 3CB-1 formed small colonies (1–2 mm in diameter) in 2 to 3 weeks. Anaerobic degradation of both 3-chlorobenzoate and 3-hydroxybenzoate was dependent on nitrate as an electron acceptor and resulted in nitrate reduction corresponding to the stoichiometric values for complete oxidation of the substrate to CO2. 3-Chlorobenzoate was not degraded in the presence of oxygen. 3-Bromobenzoate and 3-iodobenzoate were also degraded under denitrifying conditions with stoichiometric release of halide, but 3-fluorobenzoate was not utilized by the bacterium. Utilization of 3-chlorobenzoate was inducible, while synthesis of enzymes for 3-hydroxybenzoate degradation was constitutively low, but inducible. Degradation was specific to the position of the halogen substituent, and strain 3CB-1 did not utilize 2- or 4-chlorobenzoate. Received: 6 November 1998 / Accepted: 19 January 1999  相似文献   

2.
Nitrate, sulfate, and carbonate were used as electron acceptors to examine the anaerobic biodegradability of chlorinated aromatic compounds in estuarine and freshwater sediments. The respective denitrifying, sulfidogenic, and methanogenic enrichment cultures were established on each of the monochlorinated phenol and monochlorinated benzoic acid isomers, using sediment from the upper (freshwater) and lower (estuarine) Hudson River and the East River (estuarine) as source materials. Utilization of each chlorophenol and chlorobenzoate isomer was observed under at least one reducing condition; however, no single reducing condition permitted the metabolism of all six compounds tested. The anaerobic biodegradation of the chlorophenols and chlorobenzoates depended on the electron acceptor available and on the position of the chlorine substituent. In general, similar activities were observed under the different reducing conditions in both the freshwater and estuarine sediments. Under denitrifying conditions, degradation of 3- and 4-chlorobenzoate was accompanied by nitrate loss corresponding reasonably to the stoichiometric values expected for complete oxidation of the chlorobenzoate to CO2. Under sulfidogenic conditions, 3- and 4-chlorobenzoate, but not 2-chlorobenzoate, and all three monochlorophenol isomers were utilized, while under methanogenic conditions all compounds except 4-chlorobenzoate were metabolized. Given that the pattern of activity appears different for these chlorinated compounds under each reducing condition, their biodegradability appears to be more a function of the presence of competent microbial populations than one of inherent molecular structure.  相似文献   

3.
Summary Among four chlorobenzoates tested, only 3-chlorobenzoate and 4-chlorobenzoate were capable of inducing benzoate oxidizing cell activities in Acinetobacter calcoaceticus strain Bs 5, whereas 2-chlorobenzoate and 2,6-dichlorobenzoate were not. With the monochlorobenzoates, this inducing capability decreased with increasing proximity of the chlorine atom to the carboxyl group, i.e. in the order: 4-chlorobenzoate > 3-chlorobenzoate > 2-chlorobenzoate. It is therefore supposed that the induction of benzoate oxidizing cell activities is inhibited primarily be sterical influences of the chlorine substituents of the various chlorobenzoates.With decreasing concentration of 3-chlorobenzoate and 4-chlorobenzoate, the induction of benzoate oxidizing cell activities decreased. Below a critical concentration of 1 M, these activities were no longer detectable in the cells of Acinetobacter calcoaceticus, with the consequence that below this concentration limit, the degradation of 3-chlorobenzoate and 4-chlorobenzoate was no longer possible.  相似文献   

4.
Anaerobic degradation of monochlorophenols and monochlorobenzoates in a variety of aquatic sediments was compared under four enrichment conditions. A broader range of compounds was degraded in enrichments inoculated with sediment exposed to industrial effluents. Degradation of chloroaromatic compounds was observed most often in methanogenic enrichments and in enrichments amended with 1 mM bromoethane sulfonic acid. Degradation was observed least often in enrichments with added nitrate or sulfate. The presence of 10 mM bromoethane sulfonic acid prevented or inhibited degradation of most compounds tested. Primary enrichments in which KNO(3) was periodically replenished to maintain enrichment characteristics degraded chlorobenzoates, but not chlorophenols. In contrast, primary enrichments in which Na(2)SO(4) was periodically replenished failed to degrade any chloroaromatic compounds. Upon transfer to fresh medium, none of the sulfate enrichments required the presence of Na(2)SO(4) for degradation, while only two nitrate enrichments required the presence of KNO(3) for degradation. As a class of compounds, chlorophenols were degraded more readily than chlorobenzoates. However, as individual compounds 3-chlorobenzoate, 2-chlorophenol, and 3-chlorophenol degradation was observed most often and with an equal frequency. Within the chlorophenol class, the relative order of degradability was ortho > meta > para, while that of chlorobenzoates was meta > ortho > para, In laboratory transfers, 2-chlorobenzoate, 3-chlorobenzoate, and 2-chlorophenol degradation was most easily maintained, while degradation of para-chlorinated compounds was very difficult to maintain.  相似文献   

5.
T1, a dentrifying bacterium originally isolated for its ability to grow on toluene, can also metabolize 4-hydroxybenzoate and other aromatic compounds under denitrifying conditions. A cosmid clone carrying the three genes that code for the 4-chlorobenzoate dehalogenase enzyme complex isolated from the aerobic bacterium Pseudomonas sp. strain CBS3 was successfully conjugated into strain T1. The cloned enzyme complex catalyzes the hydrolytic dechlorination of 4-chlorobenzoate to 4-hydroxybenzoate. Since molecular oxygen is not required for the dehalogenation reaction, the transconjugate strain of T1 (T1-pUK45-10C) was able to grow on 4-chlorobenzoate in the absence of O2 under denitrifying conditions. 4-Chlorobenzoate was dehalogenated to 4-hydroxybenzoate, which was then further metabolized by strain T1. The dehalogenation and metabolism of 4-chlorobenzoate were nitrate dependent and were coupled to the production of nitrite and nitrogen gas. 4-Bromobenzoate was also degraded by this strain, while 4-iodobenzoate was not. Additionally, when T1-pUK45-10C was presented with a mixture of 4-chlorobenzoate and toluene, simultaneous degradation of the compounds was observed. These results illustrate that dechlorination and degradation of aromatic xenobiotics can be mediated by a pure culture in the absence of oxygen. Furthermore, it is possible to expand the range of xenobiotic substrates degradable by an organism, and it is possible that concurrent metabolism of these substrates can occur.  相似文献   

6.
Anaerobic degradation of fluorinated aromatic compounds   总被引:1,自引:0,他引:1  
Anaerobic enrichment cultures with sediment from an intertidal strait as inoculum were established under denitrifying, sulfate-reducing, iron-reducing and methanogenic conditions to examine the biodegradation of mono-fluorophenol and mono-fluorobenzoate isomers. Both phenol and benzoate were utilized within 2–6 weeks under all electron-accepting conditions. However, no degradation of the fluorophenols was observed within 1 year under any of the anaerobic conditions tested. Under denitrifying conditions, 2-fluorobenzoate and 4-fluorobenzoate were depleted within 84 days and 28 days, respectively. No loss of 3-fluorobenzoate was observed. All three fluorobenzoate isomers were recalcitrant under sulfate-reducing, iron-reducing, and methanogenic conditions. The degradation of the fluorobenzoate isomers under denitrifying conditions was examined in more detail using soils and sediments from different geographic regions around the world. Stable enrichment cultures were obtained on 2-fluorobenzoate or 4-fluorobenzoate with inoculum from most sites. Fluoride was released stoichiometrically, and nitrate reduction corresponded to the values predicted for oxidation of fluorobenzoate to CO2 coupled to denitrification. The 2-fluorobenzoate-utilizing and 4-fluorobenzoate-utilizing cultures were specific for fluorobenzoates and did not utilize other halogenated (chloro-, bromo-, iodo-) benzoic acids. Two denitrifying strains were isolated that utilized 2-fluorobenzoate and 4-fluorobenzoate as growth substrates. Preliminary characterization indicated that the strains were closely related to Pseudomonas stutzeri. Received: 1 September 1999 / Accepted in revised form: 30 September 1999  相似文献   

7.
Strain DCB-1 is a strict anaerobe capable of the reductive dechlorination of chlorobenzoates. The effect of dechlorination on the yield of pure cultures of DCB-1 was tested. Cultures were incubated with formate or H2 as electron donors and CO2 as a putative carbon source. Relative to control cultures with benzoate, cultures which dechlorinated 3-chlorobenzoate and 3,5-dichlorobenzoate had higher yields measured both as protein and cell density. On the media tested the apparent growth yield was 1.7 to 3.4 g cell protein per mole Cl- removed. Dechlorination also stimulated formate oxidation by growing cultures. Resuspended cells required an electron donor for dechlorination activity, with either formate or elemental iron serving this function. Resuspended cells did not require an electron acceptor for formate consumption, but reductive dechlorination of 3CB to benzoate stoichiometrically stimulated oxidation of formate to CO2. These results indicate that DCB-1 conserves energy for growth by coupling formate, and probably, H2 oxidation to reductive dechlorination.Non-standard abbreviations 3CB 3-chlorobenzoate - 35DCB 3,5-dichlorobenzoate - PCF Propionibacterium sp. culture fluid  相似文献   

8.
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   

9.
A bacterial isolate, Pseudomonas aeruginosa 3mT, exhibited the ability to degrade high concentrations of 3-chlorobenzoate (3-CBA, 8 g l-1) and 4-chlorobenzoate (4-CBA 12 g l-1) (Ajithkumar 1998). In this study, by delineating the initial biochemical steps involved in the degradation of these compounds, we investigated how this strain can do so well. Resting cells, permeabilised cells as well as cell-free extracts failed to dechlorinate both 3-CBA and 4-CBA under anaerobic conditions, whereas the former two readily degraded both compounds under aerobic conditions. Accumulation of any intermediary metabolite was not observed during growth as well as reaction with resting cells under highly aerated conditions. However, on modification of reaction conditions, 3-chlorocatechol (3-CC) and 4-chlorocatechol (4-CC) accumulated in 3-CBA and 4-CBA flasks, respectively. Fairly high titres of pyrocatechase II (chlorocatechol 1,2-dioxygenase) activity were obtained in extracts of cells grown on 3-CBA and 4-CBA. Meta-pyrocatechase (catechol 2,3-dioxygenase) activity against4-CC and catechol, but not against 3-CC, was also detected in low titres. Accumulation of small amounts of 2-chloro-5-hydroxy muconic semialdehyde, the meta-cleavage product of 4-CC, was detected in the medium, when 4-CBA concentration was 4 mM or greater, indicating the presence of a minor meta-pathway in strain 3mT. However, 3-CBA exclusively, and more than 99% of 4-CBA were degraded through the formation of the respective chlorocatechol, via a modified ortho-pathway. This defies the traditional view that the microbes that follow chlorocatechol pathways are not very good degraders of chlorobenzoates. 4-Hydroxybenzoatewas readily (and 3-hydroxybenzoate to a lesser extent) degraded by the strain, through the formation of protocatechuate and gentisate, respectively, as intermediary dihydroxy metabolites.  相似文献   

10.
Pseudomonas aeruginosa JB2 was isolated from a polychlorinated biphenyl-contaminated soil by enrichment culture containing 2-chlorobenzoate as the sole carbon source. Strain JB2 was subsequently found also to grow on 3-chlorobenzoate, 2,3- and 2,5-dichlorobenzoates, 2,3,5-trichlorobenzoate, and a wide range of other mono- and dihalogenated benzoic acids. Cometabolism of 2,4-dichlorobenzoate was also observed. Chlorocatechols were the central intermediates of all chlorobenzoate catabolic pathways. Degradation of 2-chlorobenzoate was routed through 3-chlorocatechol, whereas 4-chlorocatechol was identified from the metabolism of both 2,3- and 2,5-dichlorobenzoate. The initial attack on chlorobenzoates was oxygen dependent and most likely mediated by dioxygenases. Although plasmids were not detected in strain JB2, spontaneous mutants were detected in 70% of glycerol-grown colonies. The mutants were all of the following phenotype: benzoate+, 3-chlorobenzoate+, 2-chlorobenzoate-, 2,3-dichlorobenzoate-, 2,5-dichlorobenzoate-. While chlorocatechols were oxidized by the mutants at wild-type levels, oxidation of 2-chloro- and 2,3- and 2,5-dichlorobenzoates was substantially diminished. These findings suggested that strain JB2 possessed, in addition to the benzoate dioxygenase, a halobenzoate dioxygenase that was necessary for the degradation of chlorobenzoates substituted in the ortho position.  相似文献   

11.
Pseudomonas aeruginosa JB2 was isolated from a polychlorinated biphenyl-contaminated soil by enrichment culture containing 2-chlorobenzoate as the sole carbon source. Strain JB2 was subsequently found also to grow on 3-chlorobenzoate, 2,3- and 2,5-dichlorobenzoates, 2,3,5-trichlorobenzoate, and a wide range of other mono- and dihalogenated benzoic acids. Cometabolism of 2,4-dichlorobenzoate was also observed. Chlorocatechols were the central intermediates of all chlorobenzoate catabolic pathways. Degradation of 2-chlorobenzoate was routed through 3-chlorocatechol, whereas 4-chlorocatechol was identified from the metabolism of both 2,3- and 2,5-dichlorobenzoate. The initial attack on chlorobenzoates was oxygen dependent and most likely mediated by dioxygenases. Although plasmids were not detected in strain JB2, spontaneous mutants were detected in 70% of glycerol-grown colonies. The mutants were all of the following phenotype: benzoate+, 3-chlorobenzoate+, 2-chlorobenzoate-, 2,3-dichlorobenzoate-, 2,5-dichlorobenzoate-. While chlorocatechols were oxidized by the mutants at wild-type levels, oxidation of 2-chloro- and 2,3- and 2,5-dichlorobenzoates was substantially diminished. These findings suggested that strain JB2 possessed, in addition to the benzoate dioxygenase, a halobenzoate dioxygenase that was necessary for the degradation of chlorobenzoates substituted in the ortho position.  相似文献   

12.
We have previously shown that the bacterium Sphingomonas paucimobilis BPSI-3, isolated from PCB-contaminated soil, can degrade halogenated biphenyls, naphthalenes, catechols and benzoic acids. However, before such an organism can be used in bioremediation, it is important to characterise the degradation products and determine the degradation pathways to ensure that compounds more toxic or mobile than the original contaminants are not produced. In the degradation of 4-chlorobiphenyl, S. paucimobilis BPSI-3 produces a novel chlorinated picolinic acid. In this paper, we show that 4-chlorobenzoate is an intermediate in this degradation and, through 15N-labelling, that 5-chloropicolinate is the only nitrogenous metabolite isolated under the extraction conditions used. The position of the chlorine indicates that degradation of 4-chlorocatechol occurs exclusively via a 2,3-extradiol cleavage. These data allow us to postulate a more definitive catabolic pathway for the biodegradation of 4-chlorobiphenyl to 5-chloro-2-hydroxymuconic acid semialdehyde via 4-chlorobenzoate in S. paucimobilis BPSI-3. Received 19 April 1999/ Accepted in revised form 23 July 1999  相似文献   

13.
 A denitrifying bacterial biofilm population established on a polypropylene substratum of a fixed-film reactor was characterized by microscopy, scanning electron microscopy and immunofluorescence after 120 days of operation. The reactor, operated at pH 7.0, 22°C, and −180 mV with synthetic wastewater containing methanol/nitrate, achieved a denitrification rate of 0.24 mol NO- 3 l-1 day-1 with a removal efficiency for nitrate of 95%–99% at an organic loading rate of 0.325 mol methanol l-1 day-1. The gas produced contained 2%–3% (v/v) methane and 3%–4% (v/v) carbon dioxide in addition to nitrogen. The biofilm contained mainly cells of Methanobrevibacter arboriphilus antigenically related to strain DC, short, flagellated, gram-negatively staining rods of Pseudomonas sp. antigenically related to Pseudomonas stutzeri strain AN11, non-identified pink-pigmented rods and small lemon-shaped cells with mono- and bipolar appendages resembling prosthecate Hyphomicrobium sp. The biofilm analysis provided evidence for a syntrophy between the denitrifying, methylotrophic, bacterial consortium and hydrogenotrophic methanogens, which were identified by antigenic fingerprinting with 17 antibody probes. Received: 11 July 1994/Received revision: 23 September 1994/Accepted: 28 September 1994  相似文献   

14.
Summary The anaerobic degradation of phenol under denitrifying conditions by a bacterial consortium was studied both in batch and continuous cultures. Anaerobic degradation was dependent on NOf3 p– and concentrations up to 4 mm phenol were degraded within 2–5 days. During continuous growth in a fermenter, steady states could be maintained at eight dilution rates (D) corresponding to residence times between 12.5 and 50 h. Culture wash-out occurred at D=0.084 h–1. The kinetic parameters obtained for anaerobic degradation of phenol under denitrifying conditions by the consortium were: maximam specific growth rate = 0.091 h–1; saturation constant = 4.91 mg phenol/l; true growth yield = 0.57 mg dry wt/mg phenol; maintenance coefficient = 0.013 mg phenol/mg dry wt per hour. The Haldane model inhibition constant was estimated from batch culture data giving a value of 101 mg/l. The requirement of CO2 for the anaerobic degradation of phenol with NOf3 p– indicates that phenol carboxylation to 4-hydroxybenzoate was the first step of phenol degradation by this culture. 4-Hydroxybenzoate, proposed as an intermediate of phenol carboxylation under these conditions, was detected only in continuous cultures at very low growth rates (D=0.02 h–1), but was never detected as a free intermediary metabolite either in batch or in continuous cultures. Correspondence to: N. Khoury  相似文献   

15.
Microprofiles of O2 and NO3- were measured simultaneously in freshwater sediment with microsensors which were completely free from electrical interference because of coaxial designs. Depth profiles of nitrification (NO3- production) and denitrification (NO3- consumption) were subsequently determined by computer simulation of the measured microprofiles. The nitrifying bacterial community responded very quickly to changes in environmental conditions, and new steady-state microprofiles of O2 and NO3- were usually approached within a few hours after perturbation. Nitrification started quickly after introduction of O2 in previously anoxic layers, suggesting prolonged survival of the nitrifiers during anaerobiosis. Changes in the availability of O2 and NH4+ greatly affected the nitrification profile, and there was a high rate of coupled nitrification-denitrification under conditions in which nitrification occurred right above the oxic-anoxic interface. Addition of C2H2 rapidly removed the NO3- peaks, indicating that NO3- production was due mainly to autotrophic nitrification.  相似文献   

16.
Abstract Ferulic and syringic acids are methoxylated aromatic compounds that often serve as models of the subunits of lignin. Although these compounds have important implications for global carbon cycles, there is limited information on their fate in anoxic environments. Enrichment cultures were established on these two model compounds under methanogenic, sulfidogenic, and denitrifying conditions, using a Raritan River (New Jersey) marsh sediment as the inoculum. All cultures completely degraded ∼1.5 mm of both substrates. Methane production in the methanogenic cultures corresponded to the stoichiometric values expected for complete mineralization to CO2 and CH4. Sulfate and nitrate reduction in their respective cultures were both greater than 60% of the amounts predicted for complete mineralization. Aromatic intermediates of ferulic and syringic acid metabolism were identified, and pathways of degradation under sulfidogenic and denitrifying conditions are proposed. Syringic acid is sequentially O-demethylated to gallic acid under both sulfate and nitrate-reducing conditions before ring cleavage occurs. Ferulic acid undergoes propenoate side chain reduction, O-demethylation, removal of an acetate moiety from the side chain, and decarboxylation to form catechol. Catechol is further degraded under sulfidogenic conditions. Under denitrifying conditions, ferulic acid undergoes loss of an acetate moiety, prior to O-demethylation, to form protocatechuic acid, the last product detected before ring cleavage. Received: 23 February 1996; Revised: 20 May 1996; Accepted: 24 May 1996  相似文献   

17.
The anaerobic degradation ofp-cresol was studied with one sediment source under three reducing conditions—denitrifying, sulfidogenic, and methanogenic. Loss ofp-cresol (1 mM) in all the anaerobic systems took initially 3 to 4 weeks. In acclimated culturesp-cresol was degraded in less than a week.p-Cresol was completely metabolized under denitrifying, sulfidogenic, and methanogenic conditions, with formation of nitrogen gas, loss of sulfate, and formation of methane and carbon dioxide, respectively.p-Cresol metabolism proceeded throughp-hydroxybenzal-dehyde andp-hydroxybenzoate under denitrifying and methanogenic conditions. These compounds were rapidly degraded in cultures acclimated top-cresol under all three reducing conditions. These results suggest that the initial pathway ofp-cresol degradation is the same under denitryfying, sulfidogenic, and methanogenic conditions and proceeds via oxidation of the methyl substituent top-hydroxybenzaldehyde andp-hydroxybenzoate. The initial rate ofp-hydroxybenzaldehyde degradation was high in both the unacclimated cultures and in the cultures acclimated top-cresol, suggesting that this step is nonspecific. Benzoate was additionally detected as a metabolite followingp-hydroxybenzoate in the methanogenic cultures, but not in the denitrifying or sulfidogenic cultures. The degradation pathway therefore may diverge afterp-hydroxybenzoate formation depending on which electron acceptor is available.  相似文献   

18.
Bacterial biofilters used in marine recirculation aquaculture systems need improvements to enhance nitrogen removal efficiency. Relatively little is known about biofilter autochthonous population structure and function. The present study was aimed at isolating and characterizing an autochthonous denitrifying bacterium from a marine biofilter installed at a recirculation aquaculture system. Colonization of four different media in a marine fish farm was followed by isolation of various denitrifying strains and molecular classification of the most promising one, strain T2, as a novel member of the Pseudomonas fluorescens cluster. This strain exhibits high metabolic versatility regarding N and C source utilization and environmental conditions for growth. It removed nitrate through aerobic assimilatory metabolism at a specific rate of 116.2 mg NO3-N g dw−1 h−1. Dissimilatory NO3-N removal was observed under oxic conditions at a limited rate, where transient NO2-N formed represented 22% (0.17 mg L−1) of the maximum transient NO2-N observed under anoxic conditions. Dissimilatory NO3-N removal under anoxic conditions occurred at a specific rate of 53.5 mg NO3-N g dw−1 h−1. The isolated denitrifying strain was able to colonize different materials, such as granular activated carbon (GAC), Filtralite and Bioflow plastic rings, which allow the development of a prototype bioreactor for strain characterization under dynamic conditions and mimicking fish-farm operating conditions.  相似文献   

19.
O- 2 produced by the autoxidation of respiratory chain electron carriers, and other cellular reductants, inactivates bacterial and mammalian iron-sulfur-containing (de)hydratases including the citric acid cycle enzyme aconitase. Release of the solvent-exposed iron atom and oxidation of the [4Fe-4S]2+ cluster accompanies loss of catalytic activity. Rapid reactivation is achieved by iron-sulfur cluster reduction and Fe2+ insertion. Inactivation-reactivation is a dynamic and cyclical process which modulates aconitase and (de)hydratase activities in Escherichia coli and mammalian cells. The balance of inactive and active aconitase provides a sensitive measure of the changes in steady-statO- 2 levels occuring in living cells and mitochondria under stress conditions. Aconitases are also inactivated by other oxidants including O2, H2O2, NO., and ONOO which are associated with inflammation, hyperoxia and other pathophysiological conditions. Loss of aconitase activity during oxidant stress may impair energy production, and the liberation of reactive iron may further enhance oxidative damage. Iron-sulfur center cycling may also serve adaptive functions by modulating gene expression or by signaling metabolic quiescence.  相似文献   

20.
The growth of a denitrifying Pseudomonas strain on benzoic acid and 2-aminobenzoic acid (anthranilic acid) has been studied. The organism grew aerobically on benzoate, 2-aminobenzoate, and gentisate, but not on catechol or protocatechuic acid. These and other findings suggest that aerobic degradation of benzoic acid was via gentisic acid. Under completely anaerobic conditions in the presence of nitrate, benzoate and 2-aminobenzoate (5 mM each) were oxidized to CO2 with the concurrent reduction of NO 3 - to NO 2 - . Only after complete NO 3 - consumption was NO 2 - reduced to N2. Cells contained a NADP-specific 2-oxoglutaate dehydrogenase, in contrast to a NAD-specific pyruvate dehydrogenase. During anaerobic metabolism of [carboxyl-14C]benzoic acid, 16% of the label of metabolized benzoic acid was incorporated into cell material; this excludes intermediary decarboxylation during anaerobic metabolism. Extracts catalysed the activation of benzoic acid and a variety of its derivatives to the respective aryl-coenzyme A thioesters, ATP being cleaved to AMP and PPi; two synthetase activites were present. Extracts from 2-aminobenzoate-grown cells catalyzed a NADH-dependent reduction of 2-aminobenzoyl-CoA (100 nmol·min-1·mg-1 cell protein) to an unidentified CoA thioester, with a stoichiometric release of NH3 and a stoichiometry of 3 mol NADH oxidized per mol 2-aminobenzyol-CoA reduced when tested under aerobic conditions. The 2-aminobenzoyl-CoA reductase activity was lacking in anaerobic benzoate-grown cells and in aerobic cells. This is taken as evidence that 2-aminobenzoyl-CoA reductase is a key enzyme in a novel reductive pathway of anaerobic 2-aminobenzoic acid metabolism.Dedicated to Prof. Charles W. Evans  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