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 78 毫秒
1.
The synthesis of methylmercury by Desulfovibrio desulfuricans LS was investigated on the basis of 14C incorporation from precursors and the measurement of relevant enzyme activities in cell extracts. The previously observed incorporation of C-3 from serine into methylmercury was confirmed by measurement of relatively high activities of serine hydroxymethyltransferase and other enzymes of this pathway. High rates of label incorporation into methylmercury from H14COO- and H14CO3- prompted the assay of enzymes of the acetyl coenzyme A (CoA) synthase pathway. These enzymes were found to be present but at activity levels much lower than those reported for acetogens. Propyl iodide inhibited methylmercury and acetyl-CoA syntheses to similar extents, and methylmercury synthesis was found to compete with acetyl-CoA synthesis for methyl groups. On the basis of these findings, we propose that in methylmercury synthesis by D. desulfuricans LS the methyl group is transferred from CH3-tetrahydrofolate via methylcobalamin. The methyl group may originate from C-3 of serine or from formate via the acetyl-CoA synthase pathway. These pathways are not unique to D. desulfuricans LS, and thus the ability of this bacterium to methylate mercury is most likely associated with the substrate specificity of its enzymes.  相似文献   

2.
The redox properties of purified bisulfite reductases from Desulfovibrio gigas, D. desulfuricans (Norway) and Desulfotomaculum ruminis, containing non-heme iron and siroheme have been studied by EPR spectroscopy. Each enzyme shows ferric siroheme EPR signals which are not completely reduced by dithionite after 20 min, but are readily reduced within 1 min by dithionite plus methyl viologen. With the latter reducing system, each reductase also reveals a variable Beinert “g=1.94” type iron-sulfur signal. Reaction of each reductase with reduced methyl viologen results in reduction of only the siroheme. These results suggest different redox potentials for the iron-sulfur and siroheme moieties, and indicate that their functional properties are similar for each reductase.  相似文献   

3.
Two strains of Desulfovibrio desulfuricans, one known to synthesize monomethylmercury from ionic mercury, were grown to determine methylmercury toxicity and for comparison with an anaerobic strain of Clostridium pasteurianum, a H2 producer, and with the broad-spectrum mercury-resistant Pseudomonas putida strain FB-1, capable of degrading 1 μg of methylmercury to methane and elemental mercury in 2 h. The CH3HgCl resistance of D. desulfuricans strains was 10 times that of P. putida FB-1 and 100 times that of C. pasteurianum. The methylmercury resistance of D. desulfuricans was related to the disappearance of methylmercury from cultures by transformation to dimethylmercury, metacinnabar, methane, and traces of ionic mercury. During a 15-day experiment the kinetics of the two volatile compounds dimethylmercury [(CH3)2Hg] and methane were monitored in the liquid by a specific new technique with purge-and-trap gas chromatography in line with Fourier transform infrared spectroscopy and in the headspace by gas chromatography with flame ionization detection. Insoluble metacinnabar (cubic HgS) of biological origin was detected by X-ray diffractometry in the gray precipitate from the insoluble residue of the pellet of a 1-liter culture spiked with 100 mg of CH3HgCl. This was compared with a 1-liter culture of D. desulfuricans LS spiked with 100 mg of HgCl2. In a further experiment, it was demonstrated that insoluble, decomposable, white dimethylmercury sulfide [(CH3Hg)2S] formed instantly in the reaction of methylmercury with hydrogen sulfide. This organomercurial was extracted with chloroform and identified by gas chromatography in line with mass spectrometry. The D. desulfuricans strains were resistant to high concentrations of methylmercury because they produced insoluble dimethylmercury sulfide, which slowly decomposed under anaerobic conditions to metacinnabar and volatilized to dimethylmercury and methane between pHs 6.2 and 6.5 for high (4.5-g · liter-1) or low (0.09-g · liter-1) sulfate contents. Methane was produced from CH3HgCl at a lower rate than by the broad-spectrum Hg-resistant P. putida strain FB-1.  相似文献   

4.
Phospholipid Composition of Desulfovibrio Species   总被引:9,自引:7,他引:2       下载免费PDF全文
The phospholipids of Desulfovibrio desulfuricans, Norway strain, D. vulgaris, and D. gigas were examined in relationship to their qualitative and quantitative composition. D. desulfuricans and D. vulgaris exhibited an essentially identical phospholipid composition consisting of phosphatidylethanolamine, phosphatidylglycerol, cardiolipin, and lysophosphatidylserine. Phosphatidylserine (10.9%) was present in D. desulfuricans but was not detected in D. vulgaris. D. gigas was found to contain only two phospholipids, phosphatidylethanolamine (30%) and phosphatidylglycerol (70%). An ornithine-containing lipid was detected in D. gigas which was not present in the other two Desulfovibrio species.  相似文献   

5.
The anaerobic sulfate-reducing bacterium Desulfovibrio desulfuricans was grown on medium supplemented with either Kraft lignin or lignosulfonate. Only lignosulfonate contributed to the growth of D. desulfuricans cells, by replacing sulfate, a natural electron acceptor for this microorganism. Kraft lignin added to the culture medium could not substitute for lactate or sulfate, both necessary culture medium components. However, it was found to enhance the viability of D. desulfuricans cells. When changes occurring in lignin during growth of Desulfovibrio cultures were monitored, it was found that both lignin preparations could be partially depolymerized. Spectrophotometric and elemental analysis of biologically treated lignins suggested that both the polyphenolic backbone and lignin functional groups were affected by D. desulfuricans. After treatment, a twofold increase in the sulfur content of Kraft lignin and a minor decrease (14%) in the sulfur content of lignosulfonate were observed. After biological treatment, Kraft lignin and lignosulfonate both bound larger quantities of heavy metals.  相似文献   

6.
The complete genome sequence of the thermophilic sulphur-reducing bacterium, Deferribacter desulfuricans SMM1, isolated from a hydrothermal vent chimney has been determined. The genome comprises a single circular chromosome of 2 234 389 bp and a megaplasmid of 308 544 bp. Many genes encoded in the genome are most similar to the genes of sulphur- or sulphate-reducing bacterial species within Deltaproteobacteria. The reconstructed central metabolisms showed a heterotrophic lifestyle primarily driven by C1 to C3 organics, e.g. formate, acetate, and pyruvate, and also suggested that the inability of autotrophy via a reductive tricarboxylic acid cycle may be due to the lack of ATP-dependent citrate lyase. In addition, the genome encodes numerous genes for chemoreceptors, chemotaxis-like systems, and signal transduction machineries. These signalling networks may be linked to this bacterium''s versatile energy metabolisms and may provide ecophysiological advantages for D. desulfuricans SSM1 thriving in the physically and chemically fluctuating environments near hydrothermal vents. This is the first genome sequence from the phylum Deferribacteres.  相似文献   

7.
The kinetics of oxygen reduction by Desulfovibrio salexigens Mast1 and the role of polyglucose in this activity were examined and compared with those of strains of D. desulfuricans and D. gigas. Oxidation rates were highest at air saturation (up to 40 nmol of O2 min−1 mg of protein−1) and declined with decreasing oxygen concentrations. Studies with cell extracts (CE) indicated that NADH oxidase was entirely responsible for the oxygen reduction in strain Mast1. In D. desulfuricans CSN, at least three independent systems appeared to reduce oxygen. Two were active at all oxygen concentrations (NADH oxidase and NADPH oxidase), and one was maximally active at less than 10 μM oxygen. In contrast to D. gigas and D. salexigens strains, the D. desulfuricans strains also contained NADH peroxidase and NADPH peroxidase activities and did not accumulate polyglucose under nonlimiting growth conditions. At air saturation, initial activities of the oxidases and peroxidases of cells harvested at the end of the log phase were on the order of 20 to 140 nmol of O2 min−1 mg of protein−1. In all strains, these enzymes were relatively stable but were susceptible to inactivation as soon as substrates were added to the assay mixture. Under those conditions, all oxidation activity disappeared after ca. 1 h of incubation. The same finding was observed with whole cells of D. desulfuricans CSN and D. desulfuricans ATCC 27774, but inactivation was less pronounced with cells of D. salexigens Mast1. It appeared that the presence of polyglucose in the whole cells retarded the process of inactivation of NADH oxidase, but this property was lost in crude CE. In spite of the effect of polyglucose on the oxidative potential, oxygen-dependent growth of D. salexigens Mast1 could be demonstrated neither in batch nor in continuous culture.  相似文献   

8.
Sulfate-reducing bacteria have been implicated in inflammatory bowel diseases and ulcerative colitis in humans and there is an interest in inhibiting the growth of these sulfide-producing bacteria. This research explores the use of several chelators of bismuth to determine the most effective chelator to inhibit the growth of sulfate-reducing bacteria. For our studies, Desulfovibrio desulfuricans ATCC 27774 was grown with nitrate as the electron acceptor and chelated bismuth compounds were added to test for inhibition of growth. Varying levels of inhibition were attributed to bismuth chelated with subsalicylate or citrate but the most effective inhibition of growth by D. desulfuricans was with bismuth chelated by deferiprone, 3-hydroxy-1,2-dimethyl-4(1H)-pyridone. Growth of D. desulfuricans was inhibited by 10 μM bismuth as deferiprone:bismuth with either nitrate or sulfate respiration. Our studies indicate deferiprone:bismuth has bacteriostatic activity on D. desulfuricans because the inhibition can be reversed following exposure to 1 mM bismuth for 1 h at 32 °C. We suggest that deferiprone is an appropriate chelator for bismuth to control growth of sulfate-reducing bacteria because deferiprone is relatively nontoxic to animals, including humans, and has been used for many years to bind Fe(III) in the treatment of β-thalassemia.  相似文献   

9.
A protein which contains a cobalt porphyrin was isolated from the sulfate reducer Desulfovibrio gigas. This protein has a molecular weight of approximately 16,700 daltons and is acidic, having an iso-electric point at 3.7. The N-terminal residue was shown to be threonine, and a cobalt analysis gave 0.8 cobalt atoms/molecule, suggesting the presence of a single prosthetic group. The protein has a violet color with absorption bands typical of a metal porphyrin center with maxima at 420 nm, 580 nm with a shoulder at 550 nm. The ratio A420(γ)A588(α) is 2.1. The protein has no electron paramagnetic resonance (e.p.r.) spectrum, and as the visible spectrum suggests, it probably contains diamagnetic CoIII porphyrin. However the cobalt centre appears to be protected from reduction by sodium dithionite or sodium borohydride. Attempts at ligand substitution with strong nucleophiles such as CN, causes a slight spectral shift to higher wavelenghts. The cobalt porphyrin can be extracted from the protein with an acidified acetone solution, indicating that it is not covalently bound to the protein.  相似文献   

10.
11.
A cobalt-porphyrin containing protein has been isolated from the sulfate-reducer Desulfovibrio desulfuricans (Norway). This violet-colored protein has a molecular weight of approx. 13,000 daltons and contains 1 cobalt atom/molecule. The apo-protein was estimated to contain 104 amino-acid residues giving a molecular weight of 11,000 daltons. The UV-visible absorption spectrum of the protein exhibiting maxima at 588,418 and 280 nm with a shoulder at 550 nm is characteristic of metalloporphyrin proteins. The molar extinction coefficients of the cobalt-protein at 588, 418 and 280 nm are 31,330 , 64,670 and 17,200 respectively and its absorbance ratio A280A588 is 0.54. The protein is reduced by dithionite giving a blue-colored reduced form. Important spectral modifications of the chromophore occurred during the reduction including a shift of the Soret peak from 418 to 381 nm and a shift of the α band in the opposite direction from 588 to 593.5 nm. The Co-protein was slowly reduced by the hydrogenase from D.desulfuricans under hydrogen in the presence of cytochrome C3. The reported data suggest that the redox states of the cobalt center of this new electron carrier correspond to the Co(III) and Co(II) states.  相似文献   

12.

The mutual influences of Pseudomonas aeruginosa PAO1 and Desulfovibrio desulfuricans subsp. desulfuricans (ATCC 29577) on their adhesion to stainless steel were investigated in batch and column experiments. It was found that P. aeruginosa promoted the adhesion of D. desulfuricans under conditions of turbulence, but not under quiescent conditions. The enhancement involved the alignment of most D. desulfuricans along P. aeruginosa cells and was attributed to the additional interaction surface area provided by adhered P. aeruginosa to aligning D. desulfuricans cells. A slightly positive effect of pre-adhered D. desulfuricans on the adhesion of P. aeruginosa was found. Under condition of laminar flow, substantially better adhesion of D. desulfuricans to confluent P. aeruginosa biofilms than to steel was observed. The mutual influences are discussed in terms of more favorable adhesion energies and the influence of changed hydraulic conditions due to the roughness of P. aeruginosa biofilms.  相似文献   

13.
To explore the physiological role of tetraheme cytochrome c3 in the sulfate-reducing bacterium Desulfovibrio desulfuricans G20, the gene encoding the preapoprotein was cloned, sequenced, and mutated by plasmid insertion. The physical analysis of the DNA from the strain carrying the integrated plasmid showed that the insertion was successful. The growth rate of the mutant on lactate with sulfate was comparable to that of the wild type; however, mutant cultures did not achieve the same cell densities. Pyruvate, the oxidation product of lactate, served as a poor electron source for the mutant. Unexpectedly, the mutant was able to grow on hydrogen-sulfate medium. These data support a role for tetraheme cytochrome c3 in the electron transport pathway from pyruvate to sulfate or sulfite in D. desulfuricans G20.  相似文献   

14.
Adenylate kinase (AK) mediates the reversible transfer of phosphate groups between the adenylate nucleotides and contributes to the maintenance of their constant cellular level, necessary for energy metabolism and nucleic acid synthesis. The AK were purified from crude extracts of two sulfate-reducing bacteria (SRB), Desulfovibrio (D.) gigas NCIB 9332 and Desulfovibrio desulfuricans ATCC 27774, and biochemically and spectroscopically characterised in the native and fully cobalt- or zinc-substituted forms. These are the first reported adenylate kinases that bind either zinc or cobalt and are related to the subgroup of metal-containing AK found, in most cases, in Gram-positive bacteria. The electronic absorption spectrum is consistent with tetrahedral coordinated cobalt, predominantly via sulfur ligands, and is supported by EPR. The involvement of three cysteines in cobalt or zinc coordination was confirmed by chemical methods. Extended X-ray absorption fine structure (EXAFS) indicate that cobalt or zinc are bound by three cysteine residues and one histidine in the metal-binding site of the “LID” domain. The sequence 129Cys-X5-His-X15-Cys-X2-Cys of the AK from D. gigas is involved in metal coordination and represents a new type of binding motif that differs from other known zinc-binding sites of AK. Cobalt and zinc play a structural role in stabilizing the LID domain.  相似文献   

15.
1. The preparation of a crude extract of Clostridium tetanomorphum containing cobalt porphyrin synthase but little haem-synthase activity is described. 2. The properties of cobalt porphyrin synthase in the clostridial extracts is compared with the properties of a haem synthase present in crude extracts of the yeast Torulopsis utilis. 3. Cobalt porphyrin synthase in extracts of C. tetanomorphum inserts Co2+ ions into the following dicarboxylic porphyrins in descending order of rate of insertion: meso-, deutero- and proto-porphyrins. Esterification renders meso- and deutero-porphyrins inactive as substrates. Neither the tetracarboxylic (coproporphyrin III) nor the octacarboxylic (uroporphyrin III) compounds are converted into cobalt porphyrins by the extract, but the non-enzymic incorporation of Co2+ ions into these two porphyrins is rapid. These extracts are unable to insert Mn2+, Zn2+, Mg2+ or Cu2+ ions into mesoporphyrin. 4. Crude extracts of T. utilis readily insert both Co2+ and Fe2+ ions into deutero-, meso, and proto-porphyrins. Unlike the extracts of C. tetanomorphum, these preparations catalyse the insertion of Co2+ ions into deuteroporphyrin more rapidly than into mesoporphyrin. This parallels the formation of haems by the T. utilis extract. 5. Cobalt porphyrin synthase is present in the particulate fraction of the extracts of C. tetanomorphum but requires a heat-stable factor present in the soluble fraction. This soluble factor can be replaced by GSH. 6. Cobalt porphyrin synthase in the clostridial extract is inhibited by iodoacetamide and to a smaller extent by p-chloromercuribenzoate and N-ethylmaleimide. The haem synthases of T. utilis and Micrococcus denitrificans are also inhibited by various thiol reagents.  相似文献   

16.
Incubation of hepatic microsomes from phenobarbital-treated rats with methyl 2-isopropyl-4-pentenoate results in rapid destruction of the microsomal cytochrome P-450. The destruction does not occur in the absence of NADPH or with methyl 2-isopropylpentanoate. Administration of methyl 2-isopropyl-4-pentenoate to phenobarbital-pretreated rats leads to hepatic accumulation of a “green” pigment which, after methylation and purification, yields an abnormal porphyrin chromatographically and spectroscopically indistinguishable from that similarly obtained with 2-isopropyl-4-pentenamide (allylisopropylacetamide). Field desorption mass spectrometry showed that both abnormal porphyrins exhibited molecular ions at me 730. The mass spectrum of the zinc and copper complexes confirmed this value. Esterification in deuterated methanol of the amide-derived porphyrin showed that only two methyl esters were formed. Finally, methyl 4,5-epoxy-2-isopropylpentanoate and the known metabolites of 2-isopropyl-4-pentenamide were shown not to destroy cytochrome P-450. These results clearly establish that the carbonyl groups of the two destructive substrates are intimately involved in formation of the isolated porphyrin adducts, and exclude participation of the corresponding epoxide metabolites in the destruction of cytochrome P-450.  相似文献   

17.
When grown in the absence of added sulfate, cocultures of Desulfovibrio desulfuricans or Desulfovibrio vulgaris with Methanobrevibacter smithii (Methanobacterium ruminantium), which uses H2 and CO2 for methanogenesis, degraded lactate, with the production of acetate and CH4. When D. desulfuricans or D. vulgaris was grown in the absence of added sulfate in coculture with Methanosarcina barkeri (type strain), which uses both H2-CO2 and acetate for methanogenesis, lactate was stoichiometrically degraded to CH4 and presumably to CO2. During the first 12 days of incubation of the D. desulfuricans-M. barkeri coculture, lactate was completely degraded, with almost stoichiometric production of acetate and CH4. Later, acetate was degraded to CH4 and presumably to CO2. In experiments in which 20 mM acetate and 0 to 20 mM lactate were added to D. desulfuricans-M. barkeri cocultures, no detectable degradation of acetate occurred until the lactate was catabolized. The ultimate rate of acetate utilization for methanogenesis was greater for those cocultures receiving the highest levels of lactate. A small amount of H2 was detected in cocultures which contained D. desulfuricans and M. barkeri until after all lactate was degraded. The addition of H2, but not of lactate, to the growth medium inhibited acetate degradation by pure cultures of M. barkeri. Pure cultures of M. barkeri produced CH4 from acetate at a rate equivalent to that observed for cocultures containing M. barkeri. Inocula of M. barkeri grown with H2-CO2 as the methanogenic substrate produced CH4 from acetate at a rate equivalent to that observed for acetate-grown inocula when grown in a rumen fluid-vitamin-based medium but not when grown in a yeast extract-based medium. The results suggest that H2 produced by the Desulfovibrio species during growth with lactate inhibited acetate degradation by M. barkeri.  相似文献   

18.
Ecological relationships between sulfate-reducing and methane-producing bacteria in mud of Lake Vechten have been studied by continuous culture studies using the chemostat technique. The maximum specific growth rate (μ max) and saturation constant (K s) were, respectively, 0.36 hr−1 and 0.047 mM for lactate-limited growth ofDesulfovibrio desulfuricans and 0,011 hr−1 and 0.17 mM for acetate-limited growth ofMethanobacterium sp. Calculated values for the true molar growth yieldsY G) and maintenance coefficients (m) were 30.6 g bacterial mass/mole of lactate and 0.53 g substrate/g dry wt hr forD. desulfuricans and 37.8 g bacterial mass/mole of acetate and 0.54 g substrate/g dry wt hr forMethanobacterium. No growth ofMethanobacterium was observed at apS2− value (the hydrogen sulfide potential) of more than 11 and there was no effect on the growth atpS2− values above 13. In mixed continuous culture experiments the concentration of acetate decreased in the secondstage growth vessel, whereas that of methane increased stoichiometrically. If the substrate concentration in the reservoirs (S r) was increased from 0.1 to 0.5 mg/ml, the population ofDesulfovibrio increased and that ofMethanobacterium was washed out of the culture vessel, since the concentration of hydrogen sulfide reached apS2− value of 10.5. From the mixed continuous culture experiments a commensalism between the two species can be described, i.e., the acetate-fermentingMethanobacterium benefits from the acetate released byDesulfovibrio which is, in turn, not affected in the presence of the former.  相似文献   

19.
The 5,10,15-tri-(p-dodecanoxyphenyl)-20-(p-hydroxyphenyl)- and 5,10,15,20-tetrakis-(p-dodecanoxyphenyl)-cobalt porphyrin complexes were synthesized, purified and characterized. Silica substrates were functionalized with a covalent 4-ClCH2C6H4SiCl3 monolayer. Additional covalent bonding of the 5,10,15-tri-(p-dodecanoxyphenyl)-20-(p-hydroxyphenyl)-cobalt porphyrin to the silylated substrates was further achieved. The monolayer surface chemical characterization was carried out by X-ray photoelectron measurements. Both the Co 2p and N 1s spectra are evident. The NO2 sensing capability of the present cobalt porphyrin systems, at ppm levels, has been demonstrated.  相似文献   

20.
The degradation of native and pretreated nitrocellulose (NC) by the microscopic fungus Fusarium solani VKM F-819 and a mixed culture of the fungus with a sulfate-reducing bacterium Desulfovibrio desulfuricans VKM B-1388 has been studied. It has been shown that NC pretreatment with UV radiation and ozone promoted its subsequent biodegradation. The degradation of the thus treated NC by a mixed culture of F. solani and D. desulfuricans was the most effective as compared to all other treatment options. The NC nitrogen content decreased from 13.38 to 10.03%; the number average (Mn) and weight average (Mw) molecular masses decreased by three and two times, respectively. These magnitudes were achieved after 5 days of incubation of the pretreated NC. The obtained data can be used to further develop NC degradation technology.  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