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.
The metabolism of Desulfovibrio vulgaris Hildenborough grown on medium containing lactate or pyruvate plus a high concentration of sulfate (36 mM) was studied. Molecular growth yields were 6.7 +/- 1.3 and 10.1 +/- 1.7 g/mol for lactate and pyruvate, respectively. Under conditions in which the energy source was the sole growth-limiting factor, we observed the formation of 0.5 mol of hydrogen per mol of lactate and 0.1 mol of hydrogen per mol of pyruvate. The determination of metabolic end products revealed that D. vulgaris produced, in addition to normal end products (acetic acid, carbon dioxide, hydrogen sulfide) and molecular hydrogen, 2 and 5% of ethanol per mol of lactate and pyruvate, respectively. Power-time curves of growth of D. vulgaris on lactate and pyruvate were obtained, by the microcalorimetric Tian-Calvet apparatus. The enthalpies (delta Hmet) associated with the oxidation of these substrates and calculated from growth thermograms were -36.36 +/- 5 and -70.22 +/- 3 kJ/mol of lactate and pyruvate, respectively. These experimental values were in agreement with the homologous values assessed from the theoretical equations of D. vulgaris metabolism of both lactate and pyruvate. The hydrogen production by this sulfate reducer constitutes an efficient regulatory system of electrons, from energy source through the pathway of sulfate reduction. This hydrogen value may thus facilitate interactions between this strain and other environmental microflora, especially metagenic bacteria.  相似文献   

2.
When grown in the absence of added sulfate, cocultures of Desulfovibrio desulfuricans or Desulfovibrio vulgaris with Methanobrevibacter smithii (Methanobacterium ruminantium), which uses H(2) and CO(2) for methanogenesis, degraded lactate, with the production of acetate and CH(4). When D. desulfuricans or D. vulgaris was grown in the absence of added sulfate in coculture with Methanosarcina barkeri (type strain), which uses both H(2)-CO(2) and acetate for methanogenesis, lactate was stoichiometrically degraded to CH(4) and presumably to CO(2). 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 CH(4). Later, acetate was degraded to CH(4) and presumably to CO(2). 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 H(2) was detected in cocultures which contained D. desulfuricans and M. barkeri until after all lactate was degraded. The addition of H(2), but not of lactate, to the growth medium inhibited acetate degradation by pure cultures of M. barkeri. Pure cultures of M. barkeri produced CH(4) from acetate at a rate equivalent to that observed for cocultures containing M. barkeri. Inocula of M. barkeri grown with H(2)-CO(2) as the methanogenic substrate produced CH(4) 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 H(2) produced by the Desulfovibrio species during growth with lactate inhibited acetate degradation by M. barkeri.  相似文献   

3.
A mutant of Methanosarcina barkeri (Fusaro) is able to grow on pyruvate as the sole carbon and energy source. During growth, pyruvate is converted to CH4 and CO2, and about 1.5 mol of ATP per mol of CH4 is formed (A.-K. Bock, A. Prieger-Kraft, and P. Schönheit, Arch. Microbiol. 161:33-46, 1994). The pyruvate-utilizing mutant of M. barkeri could also grow on pyruvate when methanogenesis was completely inhibited by bromoethanesulfonate (BES). The mutant grew on pyruvate (80 mM) in the presence of 2 mM BES with a doubling time of about 30 h up to cell densities of about 400 mg (dry weight) of cells per liter. During growth on pyruvate, the major fermentation products were acetate and CO2 (about 0.9 mol each per mol of pyruvate). Small amounts of acetoin, acetolactate, alanine, leucine, isoleucine, and valine were also detected. CH4 was not formed. The molar growth yield (Yacetate) was about 9 g of cells (dry weight) per mol of acetate, indicating an ATP yield of about 1 mol/mol of acetate formed. Growth on pyruvate in the presence of BES was limited; after six to eight generations, the doubling times increased and the final cell densities decreased. After 9 to 11 generations, growth stopped completely. In the presence of BES, suspensions of pyruvate-grown cells fermented pyruvate to acetate, CO2, and H2. CH4 was not formed. Conversion of pyruvate to acetate, in the complete absence of methanogenesis, was coupled to ATP synthesis. Dicyclohexylcarbodiimide, an inhibitor of H(+)-translocating ATP synthase, did not inhibit ATP formation. In the presence of dicyclohexylcarbodiimide, stoichiometries of up to 0.9 mol of ATP per mol of acetate were observed. The uncoupler arsenate completely inhibited ATP synthesis, while the rates of acetate, CO2, and H2 formation were stimulated up to fourfold. Cell extracts of M. barkeri grown on pyruvate under nonmethenogenic conditions contained pyruvate: ferredoxin oxidoreductase (0.5 U/mg), phosphate acetyltransferase (12 U/mg), and acetate kinase (12 U/mg). From these data it is concluded that ATP was synthesized by substrate level phosphorylation during growth of the M. barkeri mutant on pyruvate in the absence of methanogenesis. This is the first report of growth of a methanogen under nonmethanogenic conditions at the expense of a fermentative energy metabolism.  相似文献   

4.
We compared the metabolism of methanol and acetate when Methanosarcina barkeri was grown in the presence and absence of Desulfovibrio vulgaris. The sulfate reducer was not able to utilize methanol or acetate as the electron donor for energy metabolism in pure culture, but was able to grow in coculture. Pure cultures of M. barkeri produced up to 10 mumol of H(2) per liter in the culture headspace during growth on acetate or methanol. In coculture with D. vulgaris, the gaseous H(2) concentration was 相似文献   

5.
The activities of pure and mixed cultures of Desulfovibrio vulgaris and Methanosarcina barkeri in the exponential growth phase were monitored by measuring changes in dissolved-gas concentration by membrane-inlet mass spectrometry. M. barkeri grown under H2-CO2 or methanol produced limited amounts of methane and practically no hydrogen from either substrate. The addition of CO resulted in a transient H2 production concomitant with CO consumption. Hydrogen was then taken up, and CH4 production increased. All these events were suppressed by KCN, which inhibited carbon monoxide dehydrogenase activity. Therefore, with both substrates, H2 appeared to be an intermediate in CO reduction to CH4. The cells grown on H2-CO2 consumed 4 mol of CO and produced 1 mol of CH4. Methanol-grown cells reduced CH3OH with H2 resulting from carbon monoxide dehydrogenase activity, and the ratio was then 1 mol of CH4 to 1 mol of CO. Only 12CH4 and no 13CH4 was obtained from 13CO, indicating that CO could not be the direct precursor of CH4. In mixed cultures of D. vulgaris and M. barkeri on lactate, an initial burst of H2 was observed, followed by a lower level of production, whereas methane synthesis was linear with time. Addition of CO to the mixed culture also resulted in transient extra H2 production but had no inhibitory effect upon CH4 formation, even when the sulfate reducer was D. vulgaris Hildenborough, whose periplasmic iron hydrogenase is very sensitive to CO. The hydrogen transfer is therefore probably mediated by a less CO-sensitive nickel-iron hydrogenase from either of both species.  相似文献   

6.
The activities of pure and mixed cultures of Desulfovibrio vulgaris and Methanosarcina barkeri in the exponential growth phase were monitored by measuring changes in dissolved-gas concentration by membrane-inlet mass spectrometry. M. barkeri grown under H2-CO2 or methanol produced limited amounts of methane and practically no hydrogen from either substrate. The addition of CO resulted in a transient H2 production concomitant with CO consumption. Hydrogen was then taken up, and CH4 production increased. All these events were suppressed by KCN, which inhibited carbon monoxide dehydrogenase activity. Therefore, with both substrates, H2 appeared to be an intermediate in CO reduction to CH4. The cells grown on H2-CO2 consumed 4 mol of CO and produced 1 mol of CH4. Methanol-grown cells reduced CH3OH with H2 resulting from carbon monoxide dehydrogenase activity, and the ratio was then 1 mol of CH4 to 1 mol of CO. Only 12CH4 and no 13CH4 was obtained from 13CO, indicating that CO could not be the direct precursor of CH4. In mixed cultures of D. vulgaris and M. barkeri on lactate, an initial burst of H2 was observed, followed by a lower level of production, whereas methane synthesis was linear with time. Addition of CO to the mixed culture also resulted in transient extra H2 production but had no inhibitory effect upon CH4 formation, even when the sulfate reducer was D. vulgaris Hildenborough, whose periplasmic iron hydrogenase is very sensitive to CO. The hydrogen transfer is therefore probably mediated by a less CO-sensitive nickel-iron hydrogenase from either of both species.  相似文献   

7.
Desulfovibrio vulgaris Madison and Thermodesulfobacterium commune contained functionally distinct hydrogenase activities, one which exchanged 3H2 into 3H2O and was inhibited by carbon monoxide and a second activity which produced H2 in the presence of CO. Cell suspensions of D. vulgaris used either lactate, pyruvate, or CO as the electron donor for H2 production in the absence of sulfate. Both sulfidogenic species produced and consumed hydrogen as a trace gas during growth on lactate or pyruvate as electron donors and on thiosulfate or sulfate as electron acceptors. Higher initial levels of hydrogen were detected during growth on lactate-sulfate than on pyruvate-sulfate. D. vulgaris but not T. commune also produced and then consumed CO during growth on organic electron donors and sulfate or thiosulfate. High partial pressures of exogenous H2 inhibited growth and substrate consumption when D. vulgaris was cultured on pyruvate alone but not when it was metabolizing pyruvate plus sulfate or lactate plus sulfate. The data are discussed in relation to supporting two different models for the physiological function of H2 metabolism during growth of sulfidogenic bacteria on organic electron donors plus sulfate. A trace H2 transformation model is proposed for control of redox processes during growth on either pyruvate or lactate plus sulfate, and an obligate H2 cycling model is proposed for chemiosmotic energy coupling during growth on CO plus sulfate.  相似文献   

8.
This paper shows that differences in growth behavior of Escherichia coli strain HB101 and strain HB101[pGEc47] can be related to yeast extract-enriched medium rather than plasmid properties. An optimal medium for growth of E. coli HB101[pGEc47] was designed based on the individual yield coefficients for specific medium components (NH4+ 6 g g-1, PO43- 14 g g-1, SO42- 50 g g-1). The yield coefficient for L-leucine depends on the glucose content of the medium (20 g g-1 for 3% glucose, 40 g g-1 for 1% glucose) and the yield coefficient for L-proline depends on the cultivation mode (20 g g-1 for batch cultivation, 44 g g-1 for continuous cultivation). Growth on defined medium after medium optimization is as rapid as on complex medium (0. 42-0.45 h-1). The critical dilution rate (DR) in the defined medium above which undesired production of acetic acid occurs is in the range of 0.23-0.26 h-1.  相似文献   

9.
This article aims to study hydrogen production/consumption in Desulfovibrio (D.) desulfuricans strain New Jersey, a sulfate reducer isolated from a medium undergoing active biocorrosion and to compare its hydrogen metabolism with two other Desulfovibrio species, D. gigas and D. vulgaris Hildenborough. Hydrogen production was followed during the growth of these three bacterial species under different growth conditions: no limitation of sulfate and lactate, sulfate limitation, lactate limitation, pyruvate/sulfate medium and in the presence of molybdate. Hydrogen production/consumption by D. desulfuricans shows a behavior similar to that of D. gigas but a different one from that of D. vulgaris, which produces higher quantities of hydrogen on lactate/sulfate medium. The three species are able to increase the hydrogen production when the sulfate became limiting. Moreover, in a pyruvate/sulfate medium hydrogen production was lower than on lactate/sulfate medium. Hydrogen production by D. desulfuricans in presence of molybdate is extremely high. Hydrogenases are key enzymes on production/consumption of hydrogen in sulfate reducing organisms. The specific activity, number and cellular localization of hydrogenases vary within the three Desulfovibrio species used in this work, which could explain the differences observed on hydrogen utilization.  相似文献   

10.
Interspecies hydrogen transfer was studied in Desulfovibrio vulgaris-Methanosarcina barkeri mixed cultures. Experiments were performed under batch and continuous growth culture conditions. Lactate or pyruvate was used as an energy source. In batch culture and after 30 days of simultaneous incubation, these organisms were found to yield 1.5 mol of methane and 1.5 mol of carbon dioxide per mol of lactate fermented. When M. barkeri served as the hydrogen acceptor, growth yields of D. vulgaris were higher compared with those obtained on pyruvate without any electron acceptor other than protons. In continuous culture, all of the carbon derived from the oxidation of lactate was recovered as methane and carbon dioxide, provided the dilution rate was minimal. Increasing the dilution rate induced a gradual accumulation of acetate, causing acetate metabolism to cease at above μ = 0.05 h−1. Under these conditions all of the methane produced originated from carbon dioxide. The growth yields of D. vulgaris were measured when sulfate or M. barkeri was the electron acceptor. Two key observations resulted from the present study. First, although sulfate was substituted by M. barkeri, metabolism of D. vulgaris was only slightly modified. The coculture-fermented lactate produced equimolar quantities of carbon dioxide and methane. Second, acetogenesis and methane formation from acetate were completely separable.  相似文献   

11.
Minimal growth of Methanosarcina barkeri strain DM occurred when sulfide was omitted fromthe growth medium, and addition of either sodium sulfate or coenzyme M to sulfide-depleted media failed to restore growth. Optimal growth occurred in the presence of 1.25 mM added sulfide, giving a molar growth yield (YCH4) of 4.4 mg (dry weight) of cells per mmol of CH4 produced. Increasing sulfide to 12.5 mM led to decrease in YCH4 (1.9 mg [dry weight]/mmol of CH4), in the specific growth rate and in be intracellular levels of adenosine triphosphate. However, the specific rate of methane production increased. The data suggested that at elevated sulfide levels (12.5 mM) the decrease in YCH4 might be a result of an increase in the relative energy needed for maintnenace and of uncoupling of growth from energy production.  相似文献   

12.
Minimal growth of Methanosarcina barkeri strain DM occurred when sulfide was omitted fromthe growth medium, and addition of either sodium sulfate or coenzyme M to sulfide-depleted media failed to restore growth. Optimal growth occurred in the presence of 1.25 mM added sulfide, giving a molar growth yield (YCH4) of 4.4 mg (dry weight) of cells per mmol of CH4 produced. Increasing sulfide to 12.5 mM led to decrease in YCH4 (1.9 mg [dry weight]/mmol of CH4), in the specific growth rate and in be intracellular levels of adenosine triphosphate. However, the specific rate of methane production increased. The data suggested that at elevated sulfide levels (12.5 mM) the decrease in YCH4 might be a result of an increase in the relative energy needed for maintnenace and of uncoupling of growth from energy production.  相似文献   

13.
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   

14.
The effect of low substrate concentrations on the metabolic pathway and sulfur isotope fractionation during sulfate reduction was investigated for Archaeoglobus fulgidus strain Z. This archaeon was grown in a chemostat with sulfate concentrations between 0.3 mM and 14 mM at 80°C and with lactate as the limiting substrate. During sulfate reduction, lactate was oxidized to acetate, formate, and CO2. This is the first time that the production of formate has been reported for A. fulgidus. The stoichiometry of the catabolic reaction was strongly dependent on the sulfate concentration. At concentrations of more than 300 μM, 1 mol of sulfate was reduced during the consumption of 1 mol of lactate, whereas only 0.6 mol of sulfate was consumed per mol of lactate oxidized at a sulfate concentration of 300 μM. Furthermore, at low sulfate concentrations acetate was the main carbon product, in contrast to the CO2 produced at high concentrations. We suggest different pathways for lactate oxidation by A. fulgidus at high and low sulfate concentrations. At about 300 μM sulfate both the growth yield and the isotope fractionation were limited by sulfate, whereas the sulfate reduction rate was not limited by sulfate. We suggest that the cell channels more energy for sulfate uptake at sulfate concentrations below 300 to 400 μM than it does at higher concentrations. This could explain the shift in the metabolic pathway and the reduced growth yield and isotope fractionation at low sulfate levels.  相似文献   

15.
We performed a comparative study of the growth energetics of some species of Desulfovibrio by measuring microcalorimetric and molar growth yield values. Lactate and pyruvate were used as energy sources for sulfate reduction. On lactate-sulfate media Desulfovibrio desulfuricans Norway, Desulfovibrio gigas, and Desulfovibrio africanus exhibited molar growth yields of 4.1 +/- 0.6, 3.7 +/- 1.7, and 1.8 +/- 0.1 g/mol, respectively, whereas on pyruvate-sulfate media the molar growth yields were higher (8.5 +/- 0.8, 7.7 +/- 1.6, and 3.5 +/- 0.5 g/mol, respectively). Thus, we found that D. africanus was the least efficient species in converting energy into cell material. The uncoupling of energy in this strain was obvious since its catabolic activities were high compared with those of the two other strains. The enthalpy changes associated with lactate and pyruvate metabolism were -49 +/- 0.7 and -70.2 +/- 6.0 jK/mol, respectively, for D. desulfuricans, -76.6 +/- 1.8 and -91.2 +/- 1.1 kJ/mol, respectively, for D. gigas, and -78.8 +/- 7.2 and -88.0 +/- 6.2 kJ/mol, respectively, for D. africanus. D. gigas and D. africanus produced only acetate, CO2 and hydrogen sulfide as metabolic end products. In addition to these normal end products, D. desulfuricans Norway produced a small amount of butanol. This butanol production was interpreted as reflecting a regulatory system of electron flow during the catabolism of both substrates. Such metabolism was comparable to that reported for D. vulgaris, which lost part of the reducing power of its energy sources through hydrogen evolution.  相似文献   

16.
Gluconeogenesis predominates in periportal regions of the liver lobule   总被引:2,自引:0,他引:2  
Rates of gluconeogenesis from lactate were calculated in periportal and pericentral regions of the liver lobule in perfused rat livers from increases in O2 uptake due to lactate. When lactate (0.1-2.0 mM) was infused into livers from fasted rats perfused in either anterograde or the retrograde direction, a good correlation (r = 0.97) between rates of glucose production and extra O2 uptake by the liver was observed as expected. Rates of oxygen uptake were determined subsequently in periportal and pericentral regions of the liver lobule by placing miniature oxygen electrodes on the liver surface and measuring the local change in oxygen concentration when the flow was stopped. Basal rates of oxygen uptake of 142 +/- 11 and 60 +/- 4 mumol X g-1 X h-1 were calculated for periportal and pericentral regions, respectively. Infusion of 2 mM lactate increased oxygen uptake by 71 mumol X g-1 X h-1 in periportal regions and by 29 mumol X g-1 X h-1 in pericentral areas of the liver lobule. Since the stoichiometry between glucose production and extra oxygen uptake is well-established, rates of glucose production in periportal and pericentral regions of the liver lobule were calculated from local changes in rates of oxygen uptake for the first time. Maximal rates of glucose production from lactate (2 mM) were 60 +/- 7 and 25 +/- 4 mumol X g-1 X h-1 in periportal and pericentral zones of the liver lobule, respectively. The lactate concentrations required for half-maximal glucose synthesis were similar (0.4-0.5 mM) in both regions of the liver lobule in the presence or absence of epinephrine (0.1 microM). In the presence of epinephrine, maximal rates of glucose production from lactate were 79 +/- 5 and 59 +/- 3 mumol X g-1 X h-1 in periportal and pericentral regions, respectively. Thus, gluconeogenesis from lactate predominates in periportal areas of the liver lobule during perfusion in the anterograde direction; however, the stimulation by added epinephrine was greatest in pericentral areas. Differences in local rates of glucose synthesis may be due to ATP availability, as a good correlation between basal rates of O2 uptake and rates of gluconeogenesis were observed in both regions of the liver lobule in the presence and absence of epinephrine. In marked contrast, when livers were perfused in the retrograde direction, glucose production was 28 +/- 5 mumol X g-1 X h-1 in periportal areas and 74 +/- 6 mumol X g-1 X h-1 in pericentral regions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
The objective was to determine ovarian follicular fluid concentrations of glucose, lactate, and pyruvate in relation to follicle size in buffalo and sheep. The effect of varying concentrations of these substances on in vitro oocyte maturation, oocyte protein content, and granulosa and cumulus cell growth was also investigated. Follicular fluid was aspirated from various sizes of follicles (from ovaries without a dominant follicle) collected from adult, cycling nonpregnant buffalo (Bubalus bubalis) and sheep (Ovis aries) during the breeding season. Overall, mean (+/-S.E.M.) concentrations (mM) were glucose 2.42+/-0.31 and 1.40+/-0.22, lactate 7.56+/-2.61 and 10.42+/-1.64, and pyruvate 0.02+/-0.01 and 0.002+/-0.00, in buffalo and sheep, respectively. In both species, as follicles became larger, concentrations of glucose significantly increased, lactate significantly decreased, but pyruvate was not affected. Oocyte maturation was higher (P<0.05) in medium containing supra-physiological concentrations of either glucose (5 mM), or pyruvate (10 mM) alone, or physiological concentrations of glucose, lactate and pyruvate in combination, compared to supra-physiological concentrations of lactate (15 mM) alone, or sub- or supra-physiological concentrations of glucose, lactate and pyruvate in combination (both species). The protein content of oocytes was not significantly affected by the concentration of glucose, lactate, and pyruvate in the maturation medium. However, growth of granulosa and cumulus cells was higher (P<0.05) in medium containing supra-physiological concentrations of glucose (5 mM) alone, or pyruvate (10 mM) alone, or physiological, or supra-physiological concentrations of glucose, lactate and pyruvate in combination, compared to supra-physiological concentrations of lactate (15 mM) alone, or sub-physiological concentrations of glucose, lactate and pyruvate in combination (both species). In conclusion, concentrations of glucose, pyruvate and lactate in the medium had cell type-specific effects on oocyte maturation, and on growth of granulosa and cumulus cells. Furthermore, glucose and pyruvate were the principal energy sources for oocytes and follicular somatic cells in buffalo and sheep.  相似文献   

18.
The yield coefficients forDesulfovibrio vulgaris andD. gigas varied with the electron donoracceptor combinations and with the bacterial strain. The only evidence for electron transport coupled formation of adenosine triphosphate (ATP) was with sulfate as the electron acceptor. WithD. vulgaris the ATP formation coupling to electron flow with pyruvate oxidation was 1:4 electrons and with lactate oxidation was 1:8 electrons. WithD. gigas these ratios were 1:8 electrons and 1:16 electrons for the oxidation of pyruvate and lactate. The clearest resolution of energy coupling was withD. vulgaris growing on formatesulfate medium where 2 ATP appear to be formed with the transfer of electrons from formate to adenosine phosphosulfate and one ATP with the transfer of electrons from formate to sulfite.  相似文献   

19.
Desulfovibrio vulgaris Hildenborough is a well-studied sulfate reducer that can reduce heavy metals and radionuclides [e.g., Cr(VI) and U(VI)]. Cultures grown in a defined medium had a lag period of approximately 30 h when exposed to 0.05 mM Cr(VI). Substrate analyses revealed that although Cr(VI) was reduced within the first 5 h, growth was not observed for an additional 20 h. The growth lag could be explained by a decline in cell viability; however, during this time small amounts of lactate were still utilized without sulfate reduction or acetate formation. Approximately 40 h after Cr exposure (0.05 mM), sulfate reduction occurred concurrently with the accumulation of acetate. Similar amounts of hydrogen were produced by Cr-exposed cells compared to control cells, and lactate was not converted to glycogen during non-growth conditions. D. vulgaris cells treated with a reducing agent and then exposed to Cr(VI) still experienced a growth lag, but the addition of ascorbate at the time of Cr(VI) addition prevented the lag period. In addition, cells grown on pyruvate displayed more tolerance to Cr(VI) compared to lactate-grown cells. These results indicated that D. vulgaris utilized lactate during Cr(VI) exposure without the reduction of sulfate or production of acetate, and that ascorbate and pyruvate could protect D. vulgaris cells from Cr(VI)/Cr(III) toxicity. J.D. Wall and M.W. Fields are both affiliated to the Virtual Institute of Microbial Stress and Survival (). M.E. Clark and S.B. Thieman contributed equally to this work.  相似文献   

20.
We compared the metabolism of methanol and acetate when Methanosarcina barkeri was grown in the presence and absence of Desulfovibrio vulgaris. The sulfate reducer was not able to utilize methanol or acetate as the electron donor for energy metabolism in pure culture, but was able to grow in coculture. Pure cultures of M. barkeri produced up to 10 μmol of H2 per liter in the culture headspace during growth on acetate or methanol. In coculture with D. vulgaris, the gaseous H2 concentration was ≤2 μmol/liter. The fractions of 14CO2 produced from [14C]methanol and 2-[14C]acetate increased from 0.26 and 0.16, respectively, in pure culture to 0.59 and 0.33, respectively, in coculture. Under these conditions, approximately 42% of the available electron equivalents derived from methanol or acetate were transferred and were utilized by D. vulgaris to reduce approximately 33 μmol of sulfate per 100 μmol of substrate consumed. As a direct consequence, methane formation in cocultures was two-thirds that observed in pure cultures. The addition of 5.0 mM sodium molybdate or exogenous H2 decreased the effects of D. vulgaris on the metabolism of M. barkeri. An analysis of growth and carbon and electron flow patterns demonstrated that sulfate-dependent interspecies H2 transfer from M. barkeri to D. vulgaris resulted in less methane production, increased CO2 formation, and sulfide formation from substrates not directly utilized by the sulfate reducer as electron donors for energy metabolism and growth.  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