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 31 毫秒
1.
Rhodocyclus gelatinosus strain 1 (str. 1), a photoheterotrophic bacterium, used CO as an energy substrate under anaerobic CO/light conditions, and exhibited a diauxic growth response when CO was removed from the culture. Changes in the level of cyclic AMP which occurred in cells during diauxie suggested that the cyclic nucleotide operated as an intracellular control molecule. During CO/light-phase growth, intracellular cyclic AMP was 30 pmol/mg protein, and, as str. 1 adapted for photosynthetic growth after removal of CO, intracellular cyclic AMP levels decreased to 9 pmol/mg protein. Reexposure of a light culture to CO induced synthesis of CO oxidation activity (measured as CO:MV oxidoreductase). If 10 mM cyclic AMP was added with CO, the rate of synthesis of CO:MV oxidoreductase activity increased 25-fold, and str. 1 produced 1,230 units of activity (nmol CO oxidized min-1 mg-1 protein) after only 1 h. With cyclic AMP and no CO, no incerease in CO oxidation activity was seen. Appearance of CO oxidation activity in str. 1 represented de novo protein synthesis and was blocked with chloramphenicol. In addition to stimulating formation of CO oxidative activity, a high level of cyclic AMP in str. 1 during growth with CO appeared to influence photometabolism negatively by repressing bacteriochlorophyll formation.Abbreviations Bchl a bacteriochlorophyll a - MV methyl viologen - CO MV oxidoreductase, carbon monoxide: methyl viologen oxidoreductase  相似文献   

2.
Rhodocyclus gelatinosus 1 grows anaerobically in the dark at the expense of carbon monoxide. Topographical studies with methyl viologen as the membrane probe indicated that CO oxidation and H2 production sites were on the cytoplasmic side of the cell membrane. Membrane-associated hydrogen gas production appeared to be a unidirectional reaction. In the dark, strain 1 whole cells oxidized CO and incorporated about 306 pmol of 32Pi into ATP per min per mg of protein. With CO as the sole energy-yielding substrate, cells grew with a low growth yield coefficient of 3.7 g (dry weight) of cells per mg of CO oxidized.  相似文献   

3.
A comparison of the distribution of CO oxidation activity between soluble and particulate protein fractions obtained after disruption of CO-grown Rhodopseudomonas gelatinosa 1 by French pressure cell breakage and osmotic lysis of spheroplasts suggested that, in situ, the enzyme complex was associated with the cell membrane. An improved, strictly anaerobic method is given for spectrophotometric measurement of CO oxidation activity based on the carbon monoxide:methyl viologen oxidoreductase reaction.  相似文献   

4.
The growth and gas exchange of Seliberia carboxydohydrogena Z-1062 were studied in the regime of turbidostat when the conditions of gaseous nutrition were changed: a decrease in hydrogen concentration and an increase in carbon monoxide concentration, growth on two carbon sources (CO+CO2) and on two energy sources (H2+CO). The inhibition of the bacterial growth by CO was expressed in a decrease of the specific growth rate and in the reduced effectiveness of using a gaseous substrate. When the concentration of carbon monoxide was elevated from 0 to 40% and that of hydrogen was reduced from 80 to 40%, the specific growth rate of the cells was decreased from 0.4 to 0.04 h-1; here, the economic coefficient in terms of hydrogen fell from 3.6 to 0.62 g/g. The CO-oxidizing system of the bacterium was shown to be resistant. The rate of CO oxidation by the culture was from 0.6 to 0.8 L/h per 1 g of the synthesized biomass at the following concentration of gases in the medium (%); H2, 80-40; CO2, 5; O2, 15; CO, 10-40. The rate of CO oxidation by the culture rose when hydrogen concentration was decreased and CO concentration was increased.  相似文献   

5.
Kinetic parameters of the selenium-containing, formate dehydrogenase component of the Escherichia coli formate-hydrogenlyase complex have been determined with purified enzyme. A ping-pong Bi Bi kinetic mechanism was observed. The Km for formate is 26 mM, and the Km for the electron-accepting dye, benzyl viologen, is in the range 1-5 mM. The maximal turnover rate for the formate-dependent catalysis of benzyl viologen reduction was calculated to be 1.7 x 10(5) min-1. Isotope exchange analysis showed that the enzyme catalyzes carbon exchange between carbon dioxide and formate in the absence of other electron acceptors, confirming the ping-pong reaction mechanism. Dissociation constants for formate (12.2 mM) and CO2 (8.3 mM) were derived from analysis of the isotope exchange data. The enzyme catalyzes oxidation of the alternative substrate deuterioformate with little change in the Vmax, but the Km for deuterioformate is approximately three times that of protioformate. This implies formate oxidation is not rate-limiting in the overall coupled reaction of formate oxidation and benzyl viologen reduction. The deuterium isotope effect on Vmax/Km was observed to be approximately 4.2-4.5. Sodium nitrate was found to inhibit enzyme activity in a competitive manner with respect to formate, with a Ki of 7.1 mM. Sodium azide is a noncompetitive inhibitor with a Ki of about 80 microM.  相似文献   

6.
Nostoc sp. (strain Mac) was shown to be capable of using glucose, fructose, or sucrose as a sole source of carbon and energy in the dark. In the light in the absence of exogenously supplied CO(2), this strain exhibited a more versatile metabolism. In addition to the three sugars above, glycerol and acetate served as sole sources of carbon. This photoheterotrophic growth in the absence of exogenously supplied CO(2) appears to involve O(2)-evolving photosynthesis. The action spectrum for photoheterotrophic growth on acetate closely resembles the action spectrum for photosynthesis. The physiology of photoheterotrophic growth was further investigated through determinations of stable carbon isotope ratios and measurements of gas exchanges. These investigations suggest that respired CO(2) from substrate oxidation is assimilated by the photosynthetic machinery.  相似文献   

7.
To examine the potential use of synthesis gas as a carbon and energy source in fermentation processes, Rhodospirillum rubrum was cultured on synthesis gas generated from discarded seed corn. The growth rates, growth and poly-beta-hydroxyalkanoates (PHA) yields, and CO oxidation/H(2) evolution rates were evaluated in comparison to the rates observed with an artificial synthesis gas mixture. Depending on the gas conditioning system used, synthesis gas either stimulated or inhibited CO-oxidation rates compared to the observations with the artificial synthesis gas mixture. Inhibitory and stimulatory compounds in synthesis gas could be removed by the addition of activated charcoal, char-tar, or char-ash filters (char, tar, and ash are gasification residues). In batch fermentations, approximately 1.4 mol CO was oxidized per day per g cell protein with the production of 0.75 mol H(2) and 340 mg PHA per day per g cell protein. The PHA produced from R. rubrum grown on synthesis gas was composed of 86% beta-hydroxybutyrate and 14% beta-hydroxyvalerate. Mass transfer of CO into the liquid phase was determined as the rate-limiting step in the fermentation.  相似文献   

8.
Ferrobacillus ferrooxidans, grown on either elemental sulfur or ferrous sulfate, was able to use either substrate as an energy source for the assimilation of CO(2). In both cases, 0.01 mumole of carbon was incorporated per mumole of oxygen utilized. Glucose inhibited substrate oxidation and CO(2) fixation. Sulfur and iron oxidation were inhibited 5 to 15% and 40 to 50%, respectively, in the presence of 10% glucose. Under the same conditions, CO(2) assimilation was inhibited 50% with elemental sulfur as the energy source, and was almost totally inhibited when ferrous iron was used.  相似文献   

9.
Inducer, inhibitor, and mutant studies on three hydrogenase activities of Rhodospirillum rubrum indicate that they are mediated by three distinct hydrogenase enzymes. Uptake hydrogenase mediates H2 uptake to an unknown physiological acceptor or methylene blue and is maximally synthesized during autotrophic growth in light. Formate-linked hydrogenase is synthesized primarily during growth in darkness or when light becomes limiting, and links formate oxidation to H2 production. Carbon-monoxide-linked hydrogenase is induced whenever CO is present and couples CO oxidation to H2 evolution. The enzymes can be expressed singly or conjointly depending on growth conditions, and the inhibitor or inducer added. All three hydrogenases can use methyl viologen as the mediator for both the H2 evolution and H2 uptake reactions while displaying distinct pH optima, reversibility, and sensitivity to C2H2 gas. Yet, we present evidence that the CO-linked hydrogenase, unlike the uptake hydrogenase, does not link to methylene blue as the electron acceptor. These differences allow conditions to be established to quantitatively assay each hydrogenase independently of the others both in vivo and in vitro.  相似文献   

10.
Carbon Monoxide Oxidation by Methanogenic Bacteria   总被引:31,自引:13,他引:18       下载免费PDF全文
Different species of methanogenic bacteria growing on CO(2) and H(2) were shown to remove CO added to the gas phase. Rates up to 0.2 mumol of CO depleted/min per 10 ml of culture containing approximately 7 mg of cells (wet weight) were observed. Methanobacterium thermoautotrophicum was selected for further study based on its ability to grow rapidly on a completely mineral medium. This species used CO as the sole energy source by disproportionating CO to CO(2) and CH(4) according to the following equation: 4CO + 2H(2)O --> 1CH(4) + 3CO(2). However, growth was slight, and the growth rate on CO was only 1% of that observed on H(2)/CO(2). Growth only occurred with CO concentrations in the gas phase of lower than 50%. Growth on CO agrees with the finding that cell-free extracts of M. thermoautotrophicum contained both an active factor 420 (F(420))-dependent hydrogenase (7.7 mumol/min per mg of protein at 35 degrees C) and a CO-dehydrogenating enzyme (0.2 mumol/min per mg of protein at 35 degrees C) that catalyzed the reduction of F(420) with CO. The properties of the CO-dehydrogenating enzyme are described. In addition to F(420), viologen dyes were effective electron acceptors for the enzyme. The apparent K(m) for CO was higher than 1 mM. The reaction rate increased with increasing pH and displayed an inflection point at pH 6.7. The temperature dependence of the reaction rate followed the Arrhenius equation with an activation energy (DeltaHdouble dagger) of 14.1 kcal/mol (59.0 kJ/mol). The CO dehydrogenase activity was reversibly inactivated by low concentrations of cyanide (2 muM) and was very sensitive to inactivation by oxygen. Carbon monoxide dehydrogenase of M. thermoautotrophicum exhibited several characteristic properties found for the enzyme of Clostridium pasteurianum but differed mainly in that the clostridial enzyme did not utilize F(420) as the electron acceptor.  相似文献   

11.
Extracts of aerobically, CO-autotrophically grown cells of Pseudomonas carboxydovorans were shown to catalyze the oxidation of CO to CO(2) in the presence of methylene blue, pyocyanine, thionine, phenazine methosulfate, or toluylene blue under strictly anaerobic conditions. Viologen dyes and NAD(P)(+) were ineffective as electron acceptors. The same extracts catalyzed the oxidation of formate and of hydrogen gas; the spectrum of electron acceptors was identical for the three substrates, CO, formate, and H(2). The CO- and the formate-oxidizing activities were found to be soluble enzymes, whereas hydrogenase was membrane bound exclusively. The rates of oxidation of CO, formate, and H(2) were measured spectrophotometrically following the reduction of methylene blue. The rate of carbon monoxide oxidation followed simple Michaelis-Menten kinetics; the apparent K(m) for CO was 45 muM. The reaction rate was maximal at pH 7.0, and the temperature dependence followed the Arrhenius equation with an activation energy (DeltaH(0)) of 35.9 kJ/mol (8.6 kcal/mol). Neither free formate nor hydrogen gas is an intermediate of the CO oxidation reaction. This conclusion is based on the differential sensitivity of the activities of formate dehydrogenase, hydrogenase, and CO dehydrogenase to heat, hypophosphite, chlorate, cyanide, azide, and fluoride as well as on the failure to trap free formate or hydrogen gas in coupled optical assays. These results support the following equation for CO oxidation in P. carboxydovorans: CO + H(2)O --> CO(2) + 2 H(+) + 2e(-) The CO-oxidizing activity of P. carboxydovorans differed from that of Clostridium pasteurianum by not reducing viologen dyes and by a pH optimum curve that did not show an inflection point.  相似文献   

12.
Fixation by strain DCB-1 of CO2 carbon into cell material and organic acids occurred during growth on pyruvate both with and without thiosulfate. By using sodium [14C]bicarbonate and sodium [2-14C]pyruvate, the isotopic composition of products and cells was investigated. Up to 70% of cell carbon was derived from CO2. CO2 carbon was also incorporated into succinate, formate, and acetate. Both carbons of acetate underwent exchange reactions with CO2, although the carboxyl-group exchange was twice as fast. Because strain DCB-1 uses CO2 as its major but not sole carbon source while deriving energy from pyruvate metabolism, we describe its metabolism as mixotrophic. Other mixotrophic conditions also supported growth. Lactate or butyrate, which could not support growth in mineral medium, could replace pyruvate as the oxidizable substrate only when acetate was added to the medium.  相似文献   

13.
Fixation by strain DCB-1 of CO2 carbon into cell material and organic acids occurred during growth on pyruvate both with and without thiosulfate. By using sodium [14C]bicarbonate and sodium [2-14C]pyruvate, the isotopic composition of products and cells was investigated. Up to 70% of cell carbon was derived from CO2. CO2 carbon was also incorporated into succinate, formate, and acetate. Both carbons of acetate underwent exchange reactions with CO2, although the carboxyl-group exchange was twice as fast. Because strain DCB-1 uses CO2 as its major but not sole carbon source while deriving energy from pyruvate metabolism, we describe its metabolism as mixotrophic. Other mixotrophic conditions also supported growth. Lactate or butyrate, which could not support growth in mineral medium, could replace pyruvate as the oxidizable substrate only when acetate was added to the medium.  相似文献   

14.
Clostridium thermocellum has the ability to catabolize cellulosic biomass into ethanol, but acetic acid, lactic acid, carbon dioxide, and hydrogen gas (H2) are also produced. The effect of hydrogenase inhibitors (H2, carbon monoxide (CO), and methyl viologen) on product selectivity was investigated. The anticipated effect of these hydrogenase inhibitors was to decrease acetate production. However, shifts to ethanol and lactate production are also observed as a function of cultivation conditions. When the sparge gas of cellobiose-limited chemostat cultures was switched from N2 to H2, acetate declined, and ethanol production increased 350%. In resting cell suspensions, lactate increased when H2 or CO was the inhibitor or when the cells were held at elevated hyperbaric pressure (6.8 atm). In contrast, methyl-viologen-treated resting cells produced twice as much ethanol as the other treatments. The relationship of chemostat physiology to methyl viologen inhibition was revealed by glucose transport experiments, in which methyl viologen decreased the rate of glucose transport by 90%. C. thermocellum produces NAD+ from NADH by H2, lactate, and ethanol production. When the hydrogenases were inhibited, the latter two products increased. However, excess substrate availability causes fructose 1,6-diphosphate, the glycolytic intermediate that triggers lactate production, to increase. Compensatory ethanol production was observed when the chemostat fluid dilution rate or methyl viologen decreased substrate transport. This research highlights the complex effects of high concentrations of dissolved gases in fermentation, which are increasingly envisioned in microbial applications of H2 production for the conversion of synthetic gases to chemicals.  相似文献   

15.
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 degrees 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 microM, 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 microM. 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 microM 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 microM 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   

16.
During the aerobic growth of Streptococcus faecalis strain 10C1, with limiting levels of glucose as the substrate, a molar growth yield (Y) of 58.2 g (dry weight) per mole of glucose was obtained. Under these conditions of growth, glucose was dissimilated primarily to acetate and CO(2). The incorporation of (14)C-glucose into cell material was no greater under aerobic conditions than during anaerobic growth. Assuming an adenosine triphosphate coefficient of 10.5, the aerobic Y cannot be explained solely on the basis of substrate phosphorylation and would appear to substantiate previous enzymatic evidence for oxidative phosphorylation in this cytochromeless species. With mannitol as the substrate, an aerobic Y of 64.6 was obtained. Extracts of mannitol-grown cells contained a nicotinamide adenine dinucleotide (NAD)-linked mannitol-1-phosphate (M-1-P) dehydrogenase. The difference in aerobic Y values with mannitol and glucose as substrates would indicate that the in vivo P/O ratio from the oxidation of reduced NAD generated by the oxidation of M-1-P approximates 0.6. The Y values with pyruvate and glycerol as substrates under aerobic conditions were 15.5 and 24.7, respectively.  相似文献   

17.
The steady-state rate of CO2-dependent O2 evolution by Anabaena variabilis cells in response to illumination was established after a lag phase. The lag phase was shortened (1) in cells incubated with glucose as an oxidizable substrate and (2) upon an increase in light intensity. The lag phase was absent during electron transfer from H2O to p-benzoquinone (in combination with ferricyanide) involving Photosystem II. A lag was observed during electron transfer from H2O to methyl viologen involving Photosystems II and I, but not for electron transfer from N,N,N',N'-tetramethyl-p-phenylenediamine (in combination with ascorbate) to methyl viologen involving only Photosystem I. The lag phases of the light-induced H2O --> CO2 and H2O --> methyl viologen electron transfer reactions showed the same temperature dependences at 10-30 degrees C. The lag was prevented by 3-(3,4-dichlorophenyl)-1,1-dimethylurea at concentrations that caused partial inhibition of photosynthetic O2 evolution. Retardation of cell respiration by a combination of CN- and benzylhydroxamate shortened the lag phase of the H2O --> methyl viologen electron transfer. It is concluded that the lag phase is associated with the electron transfer step between Photosystem II and Photosystem I common for the photosynthetic and respiratory chains and is due to the stimulation of cell respiration during the initial period of illumination as a consequence of an increase in the reduced plastoquinone pool and to subsequent retardation of respiration resulting from the transition of the electron transfer chain to the competitive pathway involving Photosystem I.  相似文献   

18.
Strains of Clostridium thermoaceticum were tested for H2- and CO-dependent growth in a defined medium containing metals, minerals, vitamins, cysteine-sulfide, CO2-bicarbonate, and H2 or CO. Ten of the thirteen strains tested grew at the expense of H2 and CO, and C. thermoaceticum ATCC 39073 was chosen for further study. The doubling times for H2- and CO-dependent growth under chemolithotrophic conditions (the defined medium with nicotinic acid as sole essential vitamin and sulfide as sole reducer) were 25 and 10 h, respectively. Product stiochiometries for chemolithotrophic cultures approximated: 4.1H2 + 2.4CO2----CH3COOH + 0.1 cell C + 0.3 unrecovered C and 6.8CO----CH3COOH + 3.5CO2 + 0.4 cell C + 0.9 unrecovered C. H2-dependent growth produced significantly higher acetate concentrations per unit of biomass synthesized than did CO- or glucose-dependent growth. In contrast, the doubling time for H2-dependent growth under chemolithotrophic conditions (the defined medium without vitamins and sulfide as sole reducer) by Acetogenium kivui ATCC 33488 was 2.7 h; as a sole energy source, CO was not growth supportive for A. kivui. The YH2 values for A. kivui and C. thermoaceticum were 0.91 and 0.46 g of cell dry weight per mol of H2 consumed, respectively; the YCO value for C. thermoaceticum was 1.28 g of cell dry weight per mol of CO consumed. The specific activities of hydrogenase and CO dehydrogenase in both acetogens were influenced by the energy source utilized for growth and were significantly lower in C. thermoaceticum than in A. kivui. With extracts of H2-cultivated cells and benzyl viologen as electron acceptor, the Vmax values for hydrogenase from C. thermoaceticum and A. kivui were 155.7 and 1,670 micromoles of H2 oxidized per min mg of protein, respectively; the Vmax values for CO dehydrogenase from C. thermoaceticum and A. kivui were 90.6 and 2,973 micromoles of CO oxidized per min per mg of protein, respectively.  相似文献   

19.
Paracoccus denitrificans was grown on either unlabelled glucose, [1-13C]glucose or [6-13C]glucose as the sole carbon source for growth. The cells were then incubated with a range of 14C-glucose substrates to compare the 14CO2-evolution rates between cells grown on the glucose and the 13C-labelled glucose. Cells grown on 13C-glucose had significantly faster rates of 14CO2-evolution than those grown on unlabelled glucose. The % yields of 14CO2, per [1-14C]-, [6-14C]- and [U-14C]glucose supplied were also substantially greater than those measured for cells grown on unlabelled glucose. The data indicated that growth of Paracoccus on 13C-enriched glucose substrates resulted in cells with notably different 14C-glucose oxidation metabolism compared to that observed in cells grown on unlabelled glucose.  相似文献   

20.
When incubated in the presence of CO gas, Rubrivivax gelatinosus CBS induces a CO oxidation-H2 production pathway according to the stoichiometry CO + H2O --> CO2 + H2. Once induced, this pathway proceeds equally well in both light and darkness. When light is not present, CO can serve as the sole carbon source, supporting cell growth anaerobically with a cell doubling time of nearly 2 days. This observation suggests that the CO oxidation reaction yields energy. Indeed, new ATP synthesis was detected in darkness following CO additions to the gas phase of the culture, in contrast to the case for a control that received an inert gas such as argon. When the CO-to-H2 activity was determined in the presence of the electron transport uncoupler carbonyl-cyanide m-chlorophenylhydrazone (CCCP), the rate of H2 production from CO oxidation was enhanced nearly 40% compared to that of the control. Upon the addition of the ATP synthase inhibitor N,N'-dicyclohexylcarbodiimide (DCCD), we observed an inhibition of H2 production from CO oxidation which could be reversed upon the addition of CCCP. Collectively, these data strongly suggest that the CO-to-H2 reaction yields ATP driven by a transmembrane proton gradient, but the detailed mechanism of this reaction is not yet known. These findings encourage additional research aimed at long-term H2 production from gas streams containing CO.  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