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.
Mixotrophic Growth of Hydrogenomonas eutropha   总被引:1,自引:0,他引:1  
Mixotrophic growth conditions were established by the addition of lactate to cultures of Hydrogenomonas eutropha growing autotrophically in a gaseous environment of H(2), O(2), and CO(2) (6:2:1). The specific growth rate of mixotrophic cultures was double that of the autotrophic cultures, and lactate disappearance paralleled growth. Growth yields in mixotrophic cultures were significantly greater than those in heterotrophic cultures for equal quantities of lactate consumed. The magnitude of the increase in yield was directly proportional to the absolute growth rate at the time of lactate addition to the starting autotrophic culture and to the time under mixotrophic conditions. The specific activities of hydrogenase and ribulose diphosphate carboxylase decreased during mixotrophic growth; the total activities increased somewhat. The results suggested that the complete autotrophic and heterotrophic physiologies functioned simultaneously under mixotrophic contions.  相似文献   

2.
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   

3.
Carbon dioxide and oxygen concentrations have a profound effect on the lag period of chemoautotrophically grown Hydrogenomonas eutropha. Minimum lag periods and high growth rates were obtained in shaken flask cultures with a prepared gas mixture containing 70% H(2), 20% O(2), and 10% CO(2). However, excessively long lag periods resulted when the same gas mixture was sparged through the culture. The lag period was shortened in sparged cultures by decreasing both the pO(2) and the pCO(2), indicating that gas medium equilibration had not occurred in shaken cultures. The lag period was completely eliminated at certain concentrations of O(2) and CO(2). The optimum pO(2) was 0.05 atm, but the optimum pCO(2) varied according to the pH of the medium and physiological age of the inoculum. At pH 6.4, the pCO(2) required to obtain immediate growth of exponential, postexponential, and stationary phase inocula at equal specific rates was 0.02, 0.05, and 0.16 atm, respectively. With each 0.3-unit increase in the pH of the medium, a 50% decrease in the CO(2) concentration was needed to permit growth to occur at the same rate. The pCO(2) changes required to compensate for the pH changes of the medium had the net effect of maintaining a constant bicarbonate ion concentration. Initial growth of H. eutropha was therefore indirectly related to pCO(2) and directly dependent upon a constant bicarbonate ion concentration.  相似文献   

4.
The enthalpy balance model of growth uses measurements of the rates of heat and CO(2) production to quantify rates of decarboxylation, oxidative phosphorylation and net anabolism. Enthalpy conversion efficiency (eta(H)) and the net rate of conservation of enthalpy in reduced biosynthetic products (R(SG)DeltaH(B)) can be calculated from metabolic heat rate (q) and CO(2) rate (R(CO2)). eta(H) is closely related to carbon conversion efficiency and the efficiency of conservation of available electrons in biosynthetic products. R(SG)DeltaH(B) and eta(H) can be used, together with biomass composition, to describe the rate and efficiency of growth of plant tissues. q is directly related to the rate of O(2) consumption and the ratio q:R(CO2) is inversely related to the respiratory quotient. We grew seedlings of Eucalyptus globulus at 16 and 28 degrees C for four to six weeks, then measured q and R(CO2) using isothermal calorimetry. Respiratory rate at a given temperature was increased by a lower growth temperature but eta(H) was unaffected. Enthalpy conversion efficiency - and, therefore, carbon conversion efficiency - decreased with increasing temperature from 15 to 35 degrees C. The ratio of oxidative phosphorylation to oxygen consumption (P/O ratio) was inferred in vivo from eta(H) and by assuming a constant ratio of growth to maintenance respiration with changing temperature. The P/O ratio decreased from 2.1 at 10-15 degrees C to less than 0.3 at 35 degrees C, suggesting that decreased efficiency was not only due to activity of the alternative oxidase pathway. In agreement with predictions from non-equilibrium thermodynamics, growth rate was maximal near 25 degrees C, where the calculated P/O ratio was about half maximum. We propose that less efficient pathways, such as the alternative oxidase pathway, are necessary to satisfy the condition of conductance matching whilst maintaining a near constant phosphorylation potential. These conditions minimize entropy production and maximize the efficiency of mitochondrial energy conversions as growing conditions change, while maintaining adequate finite rates of energy processing.  相似文献   

5.
The photosynthetic oxygen evolution of Chlorella vulgaris (Beijer.) cells taken from phosphate-deficient (-P) and control cultures was measured during 8 days of culture growth. Under inorganic carbon concentration (50 microM) in the measuring cell suspension and irradiance (150 micromol m(-2) s(-1)), the same as during culture growth, there were no marked differences in the photosynthetic O2 evolution rate between the -P cells and the controls. The much slower growth of -P cultures indicated that the utilization of absorbed photosynthetically active radiation (PAR) in the CO2 assimilation and biomass production were in -P cells less efficient than in the controls. Alga cells under the phosphorus stress utilized more of the absorbed PAR in the nitrate reduction than the control cells. However, under conditions of more efficient CO2 supply (inorganic carbon concentration 150 microM, introducing of exogenous carbonic anhydrase to the measuring cell suspension) and under increased irradiance (500 micromol m(-2) s(-1)), the photosynthetic O2 evolution in -P cells reached a higher rate than in the controls. The results suggest that in -P cells the restricted CO2 availability limits the total photosynthetic process. But under conditions more favorable for the CO2 uptake and under high irradiance, the -P cells may reveal a higher photosynthetic oxygen evolution rate than the controls. It is concluded that an increased potential activity of the photosynthetic light energy absorption and conversion in the C. vulgaris cells from -P cultures is a sign of acclimation to phosphorus stress by a sun-type like adaptation response of the photosynthetic apparatus.  相似文献   

6.
Two energy-generating hydrogenases enable the aerobic hydrogen bacterium Ralstonia eutropha (formerly Alcaligenes eutrophus) to use molecular hydrogen as the sole energy source. The complex synthesis of the nickel-iron-containing enzymes has to be efficiently regulated in response to H(2), which is available in low amounts in aerobic environments. H(2) sensing in R. eutropha is achieved by a hydrogenase-like protein which controls the hydrogenase gene expression in concert with a two-component regulatory system. In this study we show that the H(2) sensor of R. eutropha is a cytoplasmic protein. Although capable of H(2) oxidation with redox dyes as electron acceptors, the protein did not support lithoautotrophic growth in the absence of the energy-generating hydrogenases. A specifically designed overexpression system for R. eutropha provided the basis for identifying the H(2) sensor as a nickel-containing regulatory protein. The data support previous results which showed that the sensor has an active site similar to that of prototypic [NiFe] hydrogenases (A. J. Pierik, M. Schmelz, O. Lenz, B. Friedrich, and S. P. J. Albracht, FEBS Lett. 438:231-235, 1998). It is demonstrated that in addition to the enzymatic activity the regulatory function of the H(2) sensor is nickel dependent. The results suggest that H(2) sensing requires an active [NiFe] hydrogenase, leaving the question open whether only H(2) binding or subsequent H(2) oxidation and electron transfer processes are necessary for signaling. The regulatory role of the H(2)-sensing hydrogenase of R. eutropha, which has also been investigated in other hydrogen-oxidizing bacteria, is intimately correlated with a set of typical structural features. Thus, the family of H(2) sensors represents a novel subclass of [NiFe] hydrogenases denoted as the "regulatory hydrogenases."  相似文献   

7.
Toluene degradation occurred concomitantly with sulfate reduction in anaerobic microcosms inoculated with contaminated subsurface soil from an aviation fuel storage facility near the Patuxent River (Md.). Similar results were obtained for enrichment cultures in which toluene was the sole carbon source. Several lines of evidence suggest that toluene degradation was directly coupled to sulfate reduction in Patuxent River microcosms and enrichment cultures: (i) the two processes were synchronous and highly correlated, (ii) the observed stoichiometric ratios of moles of sulfate consumed per mole of toluene consumed were consistent with the theoretical ratio for the oxidation of toluene to CO2 coupled with the reduction of sulfate to hydrogen sulfide, and (iii) toluene degradation ceased when sulfate was depleted, and conversely, sulfate reduction ceased when toluene was depleted. Mineralization of toluene was confirmed in experiments with [ring-U-14C]toluene. The addition of millimolar concentrations of amorphous Fe(OH)3 to Patuxent River microcosms and enrichment cultures either greatly facilitated the onset of toluene degradation or accelerated the rate once degradation had begun. In iron-amended microcosms and enrichment cultures, ferric iron reduction proceeded concurrently with toluene degradation and sulfate reduction. Stoichiometric data and other observations indicate that ferric iron reduction was not directly coupled to toluene oxidation but was a secondary, presumably abiotic, reaction between ferric iron and biogenic hydrogen sulfide.  相似文献   

8.
Toluene degradation occurred concomitantly with sulfate reduction in anaerobic microcosms inoculated with contaminated subsurface soil from an aviation fuel storage facility near the Patuxent River (Md.). Similar results were obtained for enrichment cultures in which toluene was the sole carbon source. Several lines of evidence suggest that toluene degradation was directly coupled to sulfate reduction in Patuxent River microcosms and enrichment cultures: (i) the two processes were synchronous and highly correlated, (ii) the observed stoichiometric ratios of moles of sulfate consumed per mole of toluene consumed were consistent with the theoretical ratio for the oxidation of toluene to CO2 coupled with the reduction of sulfate to hydrogen sulfide, and (iii) toluene degradation ceased when sulfate was depleted, and conversely, sulfate reduction ceased when toluene was depleted. Mineralization of toluene was confirmed in experiments with [ring-U-14C]toluene. The addition of millimolar concentrations of amorphous Fe(OH)3 to Patuxent River microcosms and enrichment cultures either greatly facilitated the onset of toluene degradation or accelerated the rate once degradation had begun. In iron-amended microcosms and enrichment cultures, ferric iron reduction proceeded concurrently with toluene degradation and sulfate reduction. Stoichiometric data and other observations indicate that ferric iron reduction was not directly coupled to toluene oxidation but was a secondary, presumably abiotic, reaction between ferric iron and biogenic hydrogen sulfide.  相似文献   

9.
The effects of methanol addition and consumption on chloroform degradation rate and product distribution in methanogenic methanol enrichment cultures and in cultures of Methanosarcina barkeri 227 were investigated. Degradation of chloroform with initial concentrations up to 27.3 microM in enrichment cultures and 4.8 microM in pure cultures was stimulated by the addition of methanol. However, methanol consumption was inhibited by as little as 2.5 microM chloroform in enrichment cultures and 0.8 microM chloroform in pure cultures, suggesting that the presence of methanol, not its exact concentration or consumption rate, was the most significant variable affecting chloroform degradation rate. Methanol addition also significantly increased the number of moles of dichloromethane produced per mole of chloroform consumed. In enrichment cultures, the number of moles of dichloromethane produced per mole of chloroform consumed ranged from 0.7 (methanol consumption essentially uninhibited) to 0.35 (methanol consumption significantly inhibited) to less than 0.2 (methanol not added to the culture). In pure cultures, the number of moles of dichloromethane produced per mole of chloroform consumed was 0.47 when methanol was added and 0.24 when no methanol was added. Studies with [14C]chloroform in both enrichment and pure cultures confirmed that methanol metabolism stimulated dichloromethane production compared with CO2 production. The results indicate that while the addition of methanol significantly stimulated chloroform degradation in both methanogenic methanol enrichment cultures and cultures of M. barkeri 227, the prospects for use of methanol as a growth substrate for anaerobic chloroform-degrading systems may be limited unless the increased production of undesirable chloroform degradation products and the inhibition of methanol consumption can be mitigated.  相似文献   

10.
In cell suspension of Desulfovibrio desulfuricans B-1388, oxidation of CO as the only energy source is associated with reduction of SO42-. After a 2-h incubation of cells in 8% CO, 81% of the gas is converted. Oxidation of 1 mole CO results in formation of 0.23 mole H2S. Intracellular ATP content increases from 2.5 (control) to 8.3 nmoles/mg (during CO conversion). Dinitrophenol inhibits sulfate reduction and CO oxidation. CO dehydrogenase was detected in cytoplasmic and membrane cell fractions (59 and 34%, respectively).  相似文献   

11.
Oxygenation of carbon monoxide by bovine heart cytochrome c oxidase   总被引:1,自引:0,他引:1  
Cytochrome c oxidase (ferrocytochrome c:oxygen oxidoreductase, EC 1.9.3.1), as the terminal enzyme of the mammalian mitochondrial electron transport chain, has long been known to catalyze the reduction of dioxygen to water. We have found that when reductively activated in the presence of dioxygen, the enzyme will also catalyze the oxidation of carbon monoxide to its dioxide. Two moles of carbon dioxide is produced per mole of dioxygen, and similar rates of production are observed for 1- and 2-electron-reduced enzyme. If 13CO and O2 are used to initiate the reaction, then only 13CO2 is detected as a product. With 18O2 and 12CO, only unlabeled and singly labeled carbon dioxide are found. No direct evidence was obtained for a water-gas reaction (CO + H2O----CO2 + H2) of the oxidase with CO. The CO oxygenase activity is inhibited by cyanide, azide, and formate and is not due to the presence of bacteria. Studies with scavengers of partially reduced dioxygen show that catalase decreases the rate of CO oxygenation.  相似文献   

12.
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   

13.
The O(2)-tolerant [NiFe] hydrogenases of Ralstonia eutropha are capable of H(2) conversion in the presence of ambient O(2). Oxygen represents not only a challenge for catalysis but also for the complex assembling process of the [NiFe] active site. Apart from nickel and iron, the catalytic center contains unusual diatomic ligands, namely two cyanides (CN(-)) and one carbon monoxide (CO), which are coordinated to the iron. One of the open questions of the maturation process concerns the origin and biosynthesis of the CO group. Isotope labeling in combination with infrared spectroscopy revealed that externally supplied gaseous (13)CO serves as precursor of the carbonyl group of the regulatory [NiFe] hydrogenase in R. eutropha. Corresponding (13)CO titration experiments showed that a concentration 130-fold higher than ambient CO (0.1 ppmv) caused a 50% labeling of the carbonyl ligand in the [NiFe] hydrogenase, leading to the conclusion that the carbonyl ligand originates from an intracellular metabolite. A novel setup allowed us to the study effects of CO depletion on maturation in vivo. Upon induction of CO depletion by addition of the CO scavenger PdCl(2), cells cultivated on H(2), CO(2), and O(2) showed severe growth retardation at low cell concentrations, which was on the basis of partially arrested hydrogenase maturation, leading to reduced hydrogenase activity. This suggests gaseous CO as a metabolic precursor under these conditions. The addition of PdCl(2) to cells cultivated heterotrophically on organic substrates had no effect on hydrogenase maturation. These results indicate at least two different pathways for biosynthesis of the CO ligand of [NiFe] hydrogenase.  相似文献   

14.
Nostoc punctiforme ATCC 29133 is a nitrogen-fixing, heterocystous cyanobacterium of symbiotic origin. During nitrogen fixation, it produces molecular hydrogen (H(2)), which is recaptured by an uptake hydrogenase. Gas exchange in cultures of N. punctiforme ATCC 29133 and its hydrogenase-free mutant strain NHM5 was studied. Exchange of O(2), CO(2), N(2), and H(2) was followed simultaneously with a mass spectrometer in cultures grown under nitrogen-fixing conditions. Isotopic tracing was used to separate evolution and uptake of CO(2) and O(2). The amount of H(2) produced per molecule of N(2) fixed was found to vary with light conditions, high light giving a greater increase in H(2) production than N(2) fixation. The ratio under low light and high light was approximately 1.4 and 6.1 molecules of H(2) produced per molecule of N(2) fixed, respectively. Incubation under high light for a longer time, until the culture was depleted of CO(2), caused a decrease in the nitrogen fixation rate. At the same time, hydrogen production in the hydrogenase-deficient strain was increased from an initial rate of approximately 6 micro mol (mg of chlorophyll a)(-1) h(-1) to 9 micro mol (mg of chlorophyll a)(-1) h(-1) after about 50 min. A light-stimulated hydrogen-deuterium exchange activity stemming from the nitrogenase was observed in the two strains. The present findings are important for understanding this nitrogenase-based system, aiming at photobiological hydrogen production, as we have identified the conditions under which the energy flow through the nitrogenase can be directed towards hydrogen production rather than nitrogen fixation.  相似文献   

15.
Azotobacter vinelandii can grow with a variety of organic carbon sources and fix N2 without the need for added H2. However, due to an active H2-oxidizing system, H2-dependent mixotrophic growth in an N-free medium was demonstrated when mannose was provided as the carbon source. There was no appreciable growth with either H2 or mannose alone. Both the growth rate and the cell yield were dependent on the concentrations of both substrates, H2 and mannose. Cultures growing mixotrophically with H2 and mannose consumed approximately 4.8 mmol of O2 and produced 4.6 mmol of CO2 per mmol of mannose consumed. In the absence of H2, less CO2 was produced, less O2 was consumed, and cell growth was negligible. The rate of acetylene reduction in mixotrophic cultures was comparable to the rate in cultures grown in N-free sucrose medium. The rate of [14C]mannose uptake of cultures with H2 was greater than with argon, whereas [14C]sucrose uptake was unaffected by the addition of H2; therefore, the role of H2 in mixotrophic metabolism may be to provide energy for mannose uptake. A. vinelandii is not an autotroph, as attempts to grow the organism chemoautotrophically with H2 or to detect ribulose bisphosphate carboxylase activity were unsuccessful.  相似文献   

16.
Cell suspensions of Methanosarcina barkeri, grown on acetate, catalyzed the conversion of carbon monoxide and H2O to CO2 and H2 in stoichiometric amounts when methane formation was inhibited by bromoethanesulfonate. The specific activity was 80-120 nmol min-1 mg protein-1 at 5% CO in the gas phase. CO oxidation was coupled with the phosphorylation of ADP as indicated by a rapid increase of the intracellular ATP level upon start of the reaction. At least 0.1 mol ATP was formed/mol CO consumed. The onset of CO oxidation was also accompanied by an increase of the proton motive force (delta p) from 100 mV to 150 mV (inside negative). Addition of the uncoupler tetrachlorosalicylanilide to CO-metabolizing cells led to a rapid decrease of the ATP level and of delta p, and to an increase of the CO oxidation rate up to 70%. In the presence of the proton-translocating ATPase inhibitor N,N'-dicyclohexylcarbodiimide the phosphorylation of ADP was inhibited and CO oxidation slowed down, whereas delta p was almost unaffected. Inhibition of CO oxidation under these conditions was relieved by the addition of the protonophore tetrachlorosalicylanilide. The results indicate that in acetate-grown M. barkeri the free-energy change associated with the formation of CO2 and H2 from CO and H2O (delta G degrees = -20 kJ/mol) can be used to drive the phosphorylation of ADP and that the coupling proceeds via a chemiosmotic mechanism. A possible role of the carbon monoxide oxidation reaction as an energy-conserving site in acetate fermentation to CH4 and CO2 is discussed.  相似文献   

17.
Mining genomic databases to identify novel hydrogen producers   总被引:7,自引:0,他引:7  
The realization that fossil fuel reserves are limited and their adverse effect on the environment has forced us to look into alternative sources of energy. Hydrogen is a strong contender as a future fuel. Biological hydrogen production ranges from 0.37 to 3.3 moles H(2) per mole of glucose and, considering the high theoretical values of production (4.0 moles H(2) per mole of glucose), it is worth exploring approaches to increase hydrogen yields. Screening the untapped microbial population is a promising possibility. Sequence analysis and pathway alignment of hydrogen metabolism in complete and incomplete genomes has led to the identification of potential hydrogen producers.  相似文献   

18.
The tetrameric cytoplasmic [NiFe] hydrogenase (SH) of Ralstonia eutropha couples the oxidation of hydrogen to the reduction of NAD(+) under aerobic conditions. In the catalytic subunit HoxH, all six conserved motifs surrounding the [NiFe] site are present. Five of these motifs were altered by site-directed mutagenesis in order to dissect the molecular mechanism of hydrogen activation. Based on phenotypic characterizations, 27 mutants were grouped into four different classes. Mutants of the major class, class I, failed to grow on hydrogen and were devoid of H(2)-oxidizing activity. In one of these isolates (HoxH I64A), H(2) binding was impaired. Class II mutants revealed a high D(2)/H(+) exchange rate relative to a low H(2)-oxidizing activity. A representative (HoxH H16L) displayed D(2)/H(+) exchange but had lost electron acceptor-reducing activity. Both activities were equally affected in class III mutants. Mutants forming class IV showed a particularly interesting phenotype. They displayed O(2)-sensitive growth on hydrogen due to an O(2)-sensitive SH protein.  相似文献   

19.
The growth constant and Y (sucrose) (grams of cells per mole of sucrose) for NH(3)-grown cultures of Clostridium pasteurianum were 1.7 times those of N(2)-grown cultures, whereas the rate of sucrose utilized per gram of cells per hour was similar for both conditions. The Y (sucrose) of chemostat cultures grown on limiting NH(3) under argon at generation times equal to those of N(2)-fixing cultures was less than that of cultures grown on excess NH(3), but cells of NH(3)-limited cultures contained the N(2)-fixing system in high concentration. The concentration of the N(2)-fixing system in whole cells, when measured with adenosine triphosphate (ATP) nonlimiting, was more than twofold greater than the amount needed for the N(2) actually fixed. Thus, energy production from sucrose, and not the concentration of the N(2)-fixing system nor the maximal rate at which N(2) could be fixed, was the limiting factor for growth of N(2)-fixing cells. Either NH(3) or some product of NH(3) metabolism partially regulated the rate of sucrose metabolism since, when cultures fixing N(2), growing on NH(3), or growing on limiting NH(3) in the absence of N(2) were deprived of their nitrogen source, the rate of sucrose catabplism decreased. Calculations showed that the rate of ATP production was the growth rate-limiting factor in cells grown on N(2), and that the increased sucrose requirement of N(2)-fixing cultures in part reflected the energy demand of N(2) fixation. Calculations indicated that whole cells require about 20 moles of ATP for the fixation of 1 mole of N(2) to 2 moles of NH(3).  相似文献   

20.
The photosynthetic green sulfur bacterium Chlorobaculum tepidum assimilates CO(2) and organic carbon sources (acetate or pyruvate) during mixotrophic growth conditions through a unique carbon and energy metabolism. Using a (13)C-labeling approach, this study examined biosynthetic pathways and flux distributions in the central metabolism of C. tepidum. The isotopomer patterns of proteinogenic amino acids revealed an alternate pathway for isoleucine synthesis (via citramalate synthase, CimA, CT0612). A (13)C-assisted flux analysis indicated that carbons in biomass were mostly derived from CO(2) fixation via three key routes: the reductive tricarboxylic acid (RTCA) cycle, the pyruvate synthesis pathway via pyruvate:ferredoxin oxidoreductase, and the CO(2)-anaplerotic pathway via phosphoenolpyruvate carboxylase. During mixotrophic growth with acetate or pyruvate as carbon sources, acetyl-CoA was mainly produced from acetate (via acetyl-CoA synthetase) or citrate (via ATP citrate lyase). Pyruvate:ferredoxin oxidoreductase converted acetyl-CoA and CO(2) to pyruvate, and this growth-rate control reaction is driven by reduced ferredoxin generated during phototrophic growth. Most reactions in the RTCA cycle were reversible. The relative fluxes through the RTCA cycle were 80~100 units for mixotrophic cultures grown on acetate and 200~230 units for cultures grown on pyruvate. Under the same light conditions, the flux results suggested a trade-off between energy-demanding CO(2) fixation and biomass growth rate; C. tepidum fixed more CO(2) and had a higher biomass yield (Y(X/S), mole carbon in biomass/mole substrate) in pyruvate culture (Y(X/S) = 9.2) than in acetate culture (Y(X/S) = 6.4), but the biomass growth rate was slower in pyruvate culture than in acetate culture.  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