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.
Acetate and CO2 assimilation by Methanothrix concilii.   总被引:5,自引:2,他引:3       下载免费PDF全文
Biosynthetic pathways in Methanothrix concilii, a recently isolated aceticlastic methanogen, were studied by 13C-nuclear magnetic resonance spectroscopy. Labeling patterns of amino acids, lipids, and carbohydrates were determined. Similar to other methanogens, acetate was carboxylated to pyruvate, which was further converted to amino acids by various biosynthetic pathways. The origin of carbon atoms in glutamate, proline, and arginine clearly showed that an incomplete tricarboxylic acid cycle operating in the oxidative direction was used for their biosynthesis. Isoleucine was synthesized via citramalate, which is a typical route for methanogens. As with Methanosarcina barkeri, an extensive exchange of the label between the carboxyl group of acetate and CO2 was observed. Lipids predominantly contained diphytanyl chains, the labeling of which indicated that biosynthesis proceeded through mevalonic acid. Labeling of the C-1,6 of glucose from [2-13C]acetate is consistent with a glucogenic route for carbohydrate biosynthesis. Except for the different origins of the methyl group of methionine, the metabolic properties of Methanothrix concilii are closely related to those of Methanosarcina barkeri.  相似文献   

2.
The rate of oxidation of the carboxyl and methyl carbons of [14C]methionine to CO2 by homogenates of liver, kidney cortex, pancreas, muscle and small intestinal mucosa was studied in two breeds of sheep (Merino and Poll Dorset Horn) at three ages (2 weeks, 3 months, 4 years). Sodium alpha-keto-gamma- methiolbutyrate (0 X 4 mM) stimulated production of CO2 from the carboxyl carbon of methionine, but not from the methyl carbon. Sodium pyruvate did not affect the recovery of CO2 from either carboxyl or methyl of methionine. Sodium formate (15 mM) suppressed the conversion of the methyl carbon of methionine to CO2 by liver and kidney homogenates to 4 and 50%, respectively, of control values, but did not affect the percentage of carboxyl carbon of methionine recovered in CO2 with either tissue. With addition of S-methyl-L-cysteine (40 mM) and 3- methylthiopropionate (10 mM) the percentage of methyl and carboxyl carbons recovered in CO2 was reduced to about 20% of control values in homogenates of both tissues. Activity per gram of tissue was higher in liver and kidney cortex than in pancreas, intestinal mucosa, or muscle, with no significant differences due to breed (Merino or Poll Dorset Horn) or sex (ewe, ram or wether) of sheep. Conversion of both the carboxyl and methyl carbons to CO2 by liver was significantly lower in 2-week-old lambs than in older animals (P less than 0.01). The activity of other tissues was not markedly affected by age. Results are discussed in relation to evidence of alternative pathways of methionine catabolism, and capacities of the tissues of the sheep to catabolize methionine by alternative pathways.  相似文献   

3.
The catabolism of methanol, formate, or carbon monoxide to acetate or butyrate or both was examined in two acetogenic bacteria. Butyribacterium methylotrophicum simultaneously transformed methanol and formate mainly to butyrate with concomitant H2 and CO2 production and consumption. In contrast, methanol plus CO was primarily converted to acetate, and only slight amounts of CO2 were produced. In vivo 13C nuclear magnetic resonance analysis of [13C]methanol transformation by B. methylotrophicum indicated that methanol was predominantly incorporated into the methyl of acetate. 13CO2 was produced and then consumed, and butyrate was formed from the condensation of two acetate precursors. The analysis of the position of acetate labeled by a given 13C single-carbon substrate when B. methylotrophicum or Acetobacterium woodii was grown in the presence of a second one-carbon substrate indicated two trends: when methanol was consumed, CO, CO2, or formate predominantly labeled the acetate carboxyl; when CO was consumed, CO2 and formate were principally funneled into the acetate methyl group, and CO remained a better carboxyl precursor. These data suggest a model of acetate synthesis via the combined operation of two readily reversible single-carbon pathways which are linked by CO2.  相似文献   

4.
Total synthesis of acetate from CO2 by Clostridium acidiurici during fermentations of hypoxanthine has been shown to involve synthesis of glycine from methylenetetrahydrofolate, CO2, and NH3. The glycine is converted to serine by the addition of methylenetetrahydrofolate, and the resulting serine is converted to pyruvate, which is decarboxylated to form acetate. Since CO2 is converted to methylenetetrahydrofolate, both carbons of the acetate are derived from CO2. The evidence supporting this pathway is based on (i) the demonstration that glycine decarboxylase is present in C. acidiurici, (ii) the fact that glycine is synthesized by crude extracts at a rate which is rapid enough to account for the in vivo synthesis of acetate from CO2, (iii) the fact that methylenetetrahydrofolate is an intermediate in the formation of both carbons of acetate from CO2, and (iv) the fact that the alpha carbon of glycine is the source of the carboxyl group of acetate. Evidence is presented that this synthesis of acetate does not involve carboxylation of a methyl corrinoid enzyme such as occurs in Clostridium thermoaceticum and Clostridium formicoaceticum. Thus, there are two different mechanisms for the total synthesis of acetate from CO2 by clostridia.  相似文献   

5.
Iida K  Kajiwara M 《The FEBS journal》2007,274(19):5090-5095
The metabolic pathways leading from l-[2-13C]aspartic acid, [2-13C]glycine and l-[methyl-13C]methionine to vitamin B12 were investigated, focusing on the biosynthetic pathways leading to the aminopropanol moiety of vitamin B12 and on the role of the Shemin pathway leading to delta-aminolevulinic acid (a biosynthetic intermediate of tetrapyrrole), by means of feeding experiments with Propionibacterium shermanii in combination with 13C-NMR spectroscopy. The 13C-methylene carbons of l-[2-(13)C]aspartic acid, which is transformed to [2-13C]glycine via l-[2-13C]threonine, and [2-13C]glycine added to the culture medium served mainly to enrich the seven methyl carbons of the corrin ring through C-methylation by S-adenosyl-l-[methyl-13C]methionine derived from catabolically generated l-[methyl-13C]methionine in the presence of tetrahydrofolic acid. The results indicate that the catabolism of these amino acids predominates over pathways leading to (2R)-1-amino-2-propanol or delta-aminolevulinic acid in P. shermanii. Feeding of l-[methyl-13C]methionine efficiently enriched all seven methyl carbons. In the cases of [2-13C]glycine and l-[methyl-13C]methionine, the 13C-enrichment ratio of the methyl carbon at C-25 (the site of the first C-methylation) was less than those of the other six methyl carbons, probably due to the influence of endogenous d-glucose in P. shermanii. The almost identical 13C-enrichment ratios of the other six methyl carbons indicated that these C-methylations during vitamin B12 biosynthesis were completed before the amino acids were completely consumed. However, in the case of l-[2-13C]aspartic acid, the 13C-enrichment ratios of five methyl carbons were low and similar, whereas the last two sites of C-methylation (C-53 and C-35) were not labeled, presumably because of complete consumption of the smaller amount of added label. The ratios of 13C-incorporation into the seven methyl carbons are influenced by the conditions of amino acid feeding experiments in a manner that is dependent upon the order of C-methylation in the corrin ring of vitamin B12.  相似文献   

6.
The effect of methionine supplementation on glycine and serine metabolism was studied in vitamin B-12-deficient rats which received only 0.2% methionine in the diet. In the perfused liver, incorporation of the C-2 of glycine to the C-3 of serine was increased by addition of methionine to the perfusate. The oxidation of [1-14C]glycine to 14CO2 was however depressed. Unlike methionine, glycine did not have any significant effect on the liver folate coenzyme distribution. Oxidation of [3-14C]serine to 14CO2 both in vivo and in perfused liver was increased by methionine. A major portion of the C-3 radioactivity however was recovered in glucose. Data presented indicate that the rate of oxidation of [2-14C]histidine to 14CO2 is a more sensitive indicator of folate deficiency than the rate of oxidation of [3-14C]serine to 14CO2 although both are presumably tetrahydrofolate dependent.  相似文献   

7.
T D Hsu  M F Lux    H L Drake 《Journal of bacteriology》1990,172(10):5901-5907
The acetogen Clostridium thermoaceticum generates growth-essential CO2 equivalents from carboxylated aromatic compounds (e.g., 4-hydroxybenzoate), and these CO2 equivalents are likely integrated into the acetogenic pathway (T. Hsu, S. L. Daniel, M. F. Lux, and H. L. Drake, J. Bacteriol. 172:212-217, 1990). By using 4-hydroxybenzoate as a model substrate, an assay was developed to study the expression and activity of the decarboxylase involved in the activation of aromatic carboxyl groups. The aromatic-dependent decarboxylase was induced by carboxylated aromatic compounds in the early stages of growth and was not repressed by glucose or other acetogenic substrates; nonutilizable carboxylated aromatic compounds did not induce the decarboxylase. The decarboxylase activity displayed saturation kinetics at both whole-cell and cell extract levels, was sensitive to oxidation, and was not affected by exogenous energy sources. However, at the whole-cell level, metabolic inhibitors decreased the decarboxylase activity. Supplemental biotin or avidin did not significantly affect decarboxylation. The aromatic-dependent decarboxylase was specific for benzoates with a hydroxyl group in the para position of the aromatic ring; the meta position could be occupied by various substituent groups (-H, -OH, -OCH3, -Cl, or -F). The carboxyl carbon from [carboxyl-14C] vanillate went primarily to 14CO2 in short-term decarboxylase assays. During growth, the aromatic carboxyl group went primarily to CO2 under CO2-enriched conditions. However, under CO2-limited conditions, the aromatic carboxyl carbon went nearly totally to acetate, with equal distribution between the carboxyl and methyl carbons, thus demonstrating that acetate could be totally synthesized from aromatic carboxyl groups. In contrast, when cocultivated (i.e., supplemented) with CO under CO2-limited conditions, the aromatic carboxyl group went primarily to the methyl carbon of acetate.  相似文献   

8.
Patterns of (sup14)CO(inf2) incorporation into molecular components of the thermophilic cyanobacterial mat communities inhabiting hot springs located in Yellowstone National Park and Synechococcus sp. strain C1 were investigated. Exponentially growing Synechococcus sp. strain C1 partitioned the majority of incorporated (sup14)CO(inf2) into protein, low-molecular-weight metabolites, and lipid fractions (45, 22, and 18% of total incorporated carbon, respectively). In contrast, mat cores from various hot springs predominantly accumulated polyglucose during periods of illumination (between 77 and 85% of total incorporated (sup14)CO(inf2)). Although photosynthetically active, mat photoautotrophs do not appear to be rapidly growing, since we also detected only limited synthesis of macromolecules associated with growth (i.e., protein and rRNA). To test the hypothesis that polysaccharide reserves are fermented in situ under the dark anaerobic conditions cyanobacterial mats experience at night, mat cores were prelabeled with (sup14)CO(inf2) under illuminated conditions and then transferred to dark anaerobic conditions. Radiolabel in the polysaccharide fraction decreased by 74.7% after 12 h, of which 58.5% was recovered as radiolabeled acetate, CO(inf2), and propionate. These results indicate tightly coupled carbon fixation and fermentative processes and the potential for significant transfer of carbon from primary producers to heterotrophic members of these cyanobacterial mat communities.  相似文献   

9.
NaCN was tested for its inhibitory effects on growth of and metabolism by Methanosarcina barkeri 227. NaCN (10 microM) inhibited catabolism of acetate methyl groups to CH4 and CO2 but did not inhibit methanogenesis from methanol, CO2, methylamine, or trimethylamine. NaCN also inhibited the assimilation of methanol or CO2 (as the sole carbon source) into cell carbon and stimulated the assimilation of acetate. These results suggest that inhibition by NaCN was a result of its action as an inhibitor of in vivo CO dehydrogenase. The results also implicate CO dehydrogenase in the oxidation of acetate but not methanol methyl groups to CO2.  相似文献   

10.
Acetate turnover in the methanogenic freshwater anoxic sediments of Lake Vechten, The Netherlands, and in anoxic sediments from the Tamar Estuary, United Kingdom, and the Grosser Jasmunder Bodden, Germany, the latter two dominated by sulfate reduction, was determined. Stable isotopes and radioisotopes, inhibitors (chloroform and fluoroacetate), and methane flux were used to provide independent estimates of acetate turnover. Pore water acetate pool sizes were determined by gas chromatography with a flame ionization detector, and stable isotope-labeled acetate was determined by gas chromatography-mass spectrometry. The appearance of acetates with a different isotope labeling pattern from that initially added demonstrated that isotopic exchange occurred during methanogenic acetate metabolism. The predominant exchange processes were (i) D-H exchange in the methyl group and (ii) (sup13)C-(sup12)C exchange at the carboxyl carbon. These exchanges are most probably caused by the activity of the enzyme complex carbon monoxide dehydrogenase and subsequent methyl group dehydrogenation by tetrahydromethanopterine or a related enzyme. The methyl carbon was not subject to exchange during transformation to methane, and hence acetate with the methyl carbon labeled will provide the most reliable estimate of acetate turnover to methane. Acetate turnover rate estimates with these labels were consistent with independent estimates of acetate turnover (acetate accumulation after inhibition and methane flux). Turnover rates from either radioisotope- or stable isotope-labeled methyl carbon isotopes are, however, dependent on accurate determination of the acetate pool size. The additions of large amounts of stable isotope-labeled acetate elevate the acetate pool size, stimulating acetate consumption and causing deviation from steady-state kinetics. This can, however, be overcome by the application of a non-steady-state model. Isotopic exchange in sediments dominated by sulfate reduction was minimal.  相似文献   

11.
The effect on veratryl alcohol production of supplementing cultures of the lignin-degrading fungus Phanerochaete chrysosporium with different methyl-(sup2)H(inf3)-labelled methyl precursors has been investigated. Both chloromethane (CH(inf3)Cl) and l-methionine caused earlier initiation of veratryl alcohol biosynthesis, but S-adenosyl-l-methionine (SAM) retarded the formation of the compound. A high level of C(sup2)H(inf3) incorporation into both the 3- and 4-O-methyl groups of veratryl alcohol occurred when either l-[methyl-(sup2)H(inf3)]methionine or C(sup2)H(inf3)Cl was present, but no significant labelling was detected when S-adenosyl-l-[methyl-(sup2)H(inf3)]methionine was added. Incorporation of C(sup2)H(inf3) from C(sup2)H(inf3)Cl was strongly antagonized by the presence of unlabelled l-methionine; conversely, incorporation of C(sup2)H(inf3) from l-[methyl-(sup2)H(inf3)]methionine was reduced by CH(inf3)Cl. These results suggest that l-methionine is converted either directly or via an intermediate to CH(inf3)Cl, which is utilized as a methyl donor in veratryl alcohol biosynthesis. SAM is not an intermediate in the conversion of l-methionine to CH(inf3)Cl. In an attempt to identify the substrates for O methylation in the metabolic transformation of benzoic acid to veratryl alcohol, the relative activities of the SAM- and CH(inf3)Cl-dependent methylating systems on several possible intermediates were compared in whole mycelia by using isotopic techniques. 4-Hydroxybenzoic acid was a much better substrate for the CH(inf3)Cl-dependent methylation system than for the SAM-dependent system. The CH(inf3)Cl-dependent system also had significantly increased activities toward both isovanillic acid and vanillyl alcohol compared with the SAM-dependent system. On the basis of these results, it is proposed that the conversion of benzoic acid to veratryl alcohol involves para hydroxylation, methylation of 4-hydroxybenzoic acid, meta hydroxylation of 4-methoxybenzoic acid to form isovanillic acid, and methylation of isovanillic acid to yield veratric acid.  相似文献   

12.
Csypyrones B1, B2 and B3 are α-pyrones that can be obtained from Aspergillus oryzae expressing CsyB, which is a type III polyketide synthase. We investigated the biosynthesis of the csypyrone B compounds using [1-13C] and [2-13C] acetate feeding experiments. 13C NMR analyses of the methyl esters of the csypyrone B compounds fed with the 13C-labeled acetates showed that the carboxyl carbons of the csypyrone B side-chains were derived from the C-2 methyl carbon of the acetate. These results indicated that fatty acyl starters are involved in the CsyB reaction and that the csypyrone B compounds are formed by the oxidation of side-chains by the host fungus.  相似文献   

13.
The cellulose model compound methyl 4-O-methyl-beta-D-glucopyranosyl-(1-->4)-beta-D-glucopyranoside (6) was synthesised in high overall yield from methyl beta-D-cellobioside. The compound was crystallised from methanol to give colourless prisms, and the crystal structure was determined. The monoclinic space group is P2(1) with Z=2 and unit cell parameters a=6.6060 (13), b=14.074 (3), c=9.3180 (19) A, beta=108.95(3) degrees. The structure was solved by direct methods and refined to R=0.0286 for 2528 reflections. Both glucopyranoses occur in the 4C(1) chair conformation with endocyclic bond angles in the range of standard values. The relative orientation of both units described by the interglycosidic torsional angles [phi (O-5' [bond] C-1' [bond] O-4 [bond] C-4) -89.1 degrees, Phi (C-1' [bond] O-4 [bond] C-4 [bond] C-5) -152.0 degrees] is responsible for the very flat shape of the molecule and is similar to those found in other cellodextrins. Different rotamers at the exocyclic hydroxymethyl group for both units are present. The hydroxymethyl group of the terminal glucose moiety displays a gauche-trans orientation, whereas the side chain of the reducing unit occurs in a gauche-gauche conformation. The solid state (13)C NMR spectrum of compound 6 exhibits all 14 carbon resonances. By using different cross polarisation times, the resonances of the two methyl groups and C-6 carbons can easily be distinguished. Distinct differences of the C-1 and C-4 chemical shifts in the solid and liquid states are found.  相似文献   

14.
The synthesis of amino acids by Methanobacterium omelianskii   总被引:2,自引:1,他引:1       下载免费PDF全文
1. Methanobacterium omelianskii was grown on (14)CO(2) and unlabelled ethanol, or on [1-(14)C]- or [2-(14)C]-ethanol and unlabelled carbon dioxide. The cell protein was hydrolysed and certain of the amino acids were isolated and degraded. 2. Carbon from both carbon dioxide and ethanol is used for biosynthesis of amino acids, and in most cases ethanol is incorporated as a C(2) unit. Ethanol carbon atoms and carbon dioxide carbon atoms apparently enter the same range of compounds. Ethanol and carbon dioxide are equally important as sources of cell carbon. 3. The origins of carbon atoms of aspartate, alanine, glycine, serine and threonine are consistent with the synthesis of these amino acids, by pathways known to exist in aerobic organisms, from pyruvate arising by a C(2)+C(1) condensation. The proportion of total radioactivity found in C-1 of lysine, proline, methionine and valine is consistent with synthesis of these amino acids by pathways similar to those found in Escherichia coli. Isoleucine is probably formed by carboxylation of a C(5) precursor formed entirely from ethanol. Glutamate is formed by an unknown pathway.  相似文献   

15.
Bacterial strains from human feces that reduce CO2 to acetic acid.   总被引:2,自引:2,他引:0       下载免费PDF全文
We used dilutions of fecal suspensions from a human volunteer to enrich cultures for bacteria that reduce CO2 to acetate in the colon. The soluble enrichment substrates used were glucose, methanol, formate, and vanillate, which were used with a gas phase that contained 80% N2 and 20% CO2. The gaseous enrichment substrates used were 80% H2-20% CO2 and 50% CO-50% CO2. We isolated three different strains that produced acetate from CO2. One strain produced acetate from methanol, vanillate, H2-CO2, glucose, and other sugars. The other two strains did not form acetate from methanol or vanillate. Both of the latter strains formed acetate from glucose and other sugars, but only one of these strains formed acetate from H2-CO2. Both of these strains cometabolized formate. However, none of the enrichment cultures or pure cultures used CO or formate as a substrate for growth. The two strains that produced acetate from H2 and CO2 grew slowly when the gases alone were used as substrates, but they rapidly cometabolized H2 and CO2 when they were grown with organic substrates. The ability of all of the strains to produce acetate from CO2 and/or other one-carbon precursors was verified by determining the radioactivity of the methyl and carboxyl groups of the acetate formed after growth with 14CO2 or other radioactively labeled one-carbon precursors.  相似文献   

16.
Methanosarcina barkeri was grown by acetate fermentation in complex medium (N2 gas phase). The molar growth yield was 1.6–1.9 g cells/mol methane formed. Under these conditions 63–82% of the methane produced byMethanosarcina strains was derived from the methyl carbon of acetate, indicating that some methane was derived from other media components. Growth was not demonstrated in complex media lacking acetate or mineral acetate medium containing acetate but lacking H2/CO2, methanol, or trypticase and yeast extract. Acetate metabolism byM. barkeri strain MS was further exmined in mineral acetate medium containing H2/CO2 and/or methanol, but lacking cysteine. Under these conditions, more methane was derived from the methyl carbon of acetate than from the carboxyl carbon. Methanogenesis from the methyl group increased with increasing acetate concentration. The methyl carbon contributed up to 42% of the methane formed with H2/CO2 and up to 5% with methanol. Methanol stimulated the oxidation of the methyl group of acetate to CO2. The average rates of methane formation from acetate were 1.3 nomol/min ·ml/culture (0.04mg2 cell dry weight) in defined media (gas phase H2/CO2) and complex media (gas phase N2). Acetate contributed up to 60% of cell carbon formed under the growth conditions examined. Similar quantities of cell carbon were derived from the methyl and carboxyl carbons of acetate, suggesting incorporation of this compound as a two-carbon unit. Incorporated acetate was not preferentially localized in lipid material, as 70% of the incorporated acetate was found in the wall and protein cell fractions. Acetate catabolism was stimulated by pregrowing of cultures in media containing acetate, while acetate anabolism was not influenced. The results are discussed in terms of the differences between the mechanisms of acetate catabolism and anabolism.Abbreviations CH3-S-CoM methyl coenzyme M - TCA trichloroacetic acid - CoM coenzyme M (2-mercaptoethane sulfonic acid) - Eo standard potential change (pH 7) - F420 Factor 420, a low redox electron carrier - Go standard free energy change (pH 7) - kJ kilojoules (=0.24 kilocalories) - PBBW Weimer's phosphate-buffered basal medium - X unknown C1 carrier  相似文献   

17.
Clostridium formicoaceticum ferments fructose labeled with (14)C in carbon 1, 4, 5, or 6 via the Embden Meyerhof pathway. In fermentations of fructose in the presence of (14)CO(2), acetate is formed labeled equally in both carbons. Extracts convert the methyl groups of 5-methyltetrahydrofolate and methyl-B(12) to the methyl group of acetate in the presence of pyruvate. Formate dehydrogenase, 10-formyltetrahydrofolate synthetase, 5,10-methenyltetrahydrofolate cyclohydrolase, 5,10-methylenetetrahydrofolate dehydrogenase, and 5,10-methylenetetrahydrofolate reductase are present in extracts of C. formicoaceticum. These enzymes are needed for the conversion of CO(2) to 5-methyltetrahydrofolate. It is proposed that acetate is totally synthesized from CO(2) via the reactions catalyzed by the enzymes listed above and that 5-methyltetra-hydrofolate and a methylcorrinoid are intermediates in this synthesis.  相似文献   

18.
The nuclear magnetic resonance technique was used to study carbon dioxide reduction by the pig large-intestinal flora. Washed bacterial cell suspensions were incubated for 6 and 15 h under 13CO2 and H2 as the gas phase and with a buffer containing NaH13CO3 and cellobiose and amino acids (casein hydrolysate) as substrates. Methane was produced in all incubation media. Significant amounts of single- as well as multiple-labelled acetate and butyrate were formed, demonstrating synthesis of acetate from H2 + CO2. Propionate was labelled mainly on the carboxyl group, which was attributed to an enzymatic exchange of the carboxyl group of propionate with 13CO2. These results indicate that the reduction of CO2 to acetate may be an important pathway for microbial production of acetate in the pig large intestine even in the presence of methanogenesis.  相似文献   

19.
Sediments from mercury-contaminated and uncontaminated reaches of the Carson River, Nevada, were assayed for sulfate reduction, methanogenesis, denitrification, and monomethylmercury (MeHg) degradation. Demethylation of [(sup14)C]MeHg was detected at all sites as indicated by the formation of (sup14)CO(inf2) and (sup14)CH(inf4). Oxidative demethylation was indicated by the formation of (sup14)CO(inf2) and was present at significant levels in all samples. Oxidized/reduced demethylation product ratios (i.e., (sup14)CO(inf2)/(sup14)CH(inf4) ratios) generally ranged from 4.0 in surface layers to as low as 0.5 at depth. Production of (sup14)CO(inf2) was most pronounced at sediment surfaces which were zones of active denitrification and sulfate reduction but was also significant within zones of methanogenesis. In a core taken from an uncontaminated site having a high proportion of oxidized, coarse-grain sediments, sulfate reduction and methanogenic activity levels were very low and (sup14)CO(inf2) accounted for 98% of the product formed from [(sup14)C]MeHg. There was no apparent relationship between the degree of mercury contamination of the sediments and the occurrence of oxidative demethylation. However, sediments from Fort Churchill, the most contaminated site, were most active in terms of demethylation potentials. Inhibition of sulfate reduction with molybdate resulted in significantly depressed oxidized/reduced demethylation product ratios, but overall demethylation rates of inhibited and uninhibited samples were comparable. Addition of sulfate to sediment slurries stimulated production of (sup14)CO(inf2) from [(sup14)C]MeHg, while 2-bromoethanesulfonic acid blocked production of (sup14)CH(inf4). These results reveal the importance of sulfate-reducing and methanogenic bacteria in oxidative demethylation of MeHg in anoxic environments.  相似文献   

20.
The oxidation of [(sup14)C]methyl bromide ([(sup14)C]MeBr) to (sup14)CO(inf2) was measured in field experiments with soils collected from two strawberry plots fumigated with mixtures of MeBr and chloropicrin (CCl(inf3)NO(inf2)). Although these fumigants are considered potent biocides, we found that the highest rates of MeBr oxidation occurred 1 to 2 days after injection when the fields were tarped, rather than before or several days after injection. No oxidation of MeBr occurred in heat-killed soils, indicating that microbes were the causative agents of the oxidation. Degradation of MeBr by chemical and/or biological processes accounted for 20 to 50% of the loss of MeBr during fumigation, with evasion to the atmosphere inferred to comprise the remainder. In laboratory incubations, complete removal of [(sup14)C]MeBr occurred within a few days, with 47 to 67% of the added MeBr oxidized to (sup14)CO(inf2) and the remainder of counts associated with the solid phase. Chloropicrin inhibited the oxidation of MeBr, implying that use of this substance constrains the extent of microbial degradation of MeBr during fumigation. Oxidation was by direct bacterial attack of MeBr and not of methanol, a product of the chemical hydrolysis of MeBr. Neither nitrifying nor methane-oxidizing bacteria were sufficiently active in these soils to account for the observed oxidation of MeBr, nor could the microbial degradation of MeBr be linked to cooxidation with exogenously supplied electron donors. However, repeated addition of MeBr to live soils resulted in higher rates of its removal, suggesting that soil bacteria used MeBr as an electron donor for growth. To support this interpretation, we isolated a gram-negative, aerobic bacterium from these soils which grew with MeBr as a sole source of carbon and energy.  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