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 171 毫秒
1.
Butyribacterium methylotrophicum produced more butyrate when grown on lactate than when grown on glucose, and only acetate was detected during growth on pyruvate. Higher levels of NADH were found in butyrate-producing than in acetate-producing cells. The addition of neutral red, an electron-flow modulator, to cells growing on pyruvate altered the carbon and electron flow from acetate plus H2 synthesis to butyrate synthesis. Enzymatic analysis suggested that pyruvate was produced from glucose via an Embden-Meyerhof-Parnas pathway. Pyruvate was further metabolized to butyryl-CoA via, β-hydroxybutyryl-CoA and butyryl-CoA dehydrogenases. Lactate dehydrogenase, unlike butyryl-CoA dehydrogenase, was inducible and detected only in lactate-grown cells. Both of these dehydrogenases utilized 2,6-dichloroindophenol and other artificial electron acceptors but not NAD(P). Ferredoxin–NAD oxidoreductase levels were highest in lactate and lowest in pyruvate-grown cells. Cells contained both a ferredoxin–neutral-red reductase activity and a neutral-red–NAD reductase activity that coupled electron flow to butyrate synthesis. These results showed that butyrate synthesis by B. methylotrophicum was regulated by the carbon source and was dependent on the cellular NADH/NAD ratios, and the levels and direction of ferredoxin- and NAD-linked oxidoreductases. Received: 3 August 1995/Received revision: 31 October 1995/Accepted: 10 November 1995  相似文献   

2.
The Rnf complex is a membrane-bound ferredoxin(Fd):NAD(P)+ oxidoreductase (Fno) that couples Fd oxidation to vectorial H+/Na+ transport across the cytoplasmic membrane. Here, we produced two putative Rnf-complexes from Clostridioides difficile (Cd-Rnf) and Clostridium ljungdahlii (Cl-Rnf) for the first time in Escherichia coli. A redox-responsive low-expression system enabled Rnf assembly in the membranes of E. coli as confirmed by in vitro activity measurements. To study the physiological effects of Rnf on the metabolism of E. coli, we assembled additional Fd-dependent enzymes by plasmid-based multigene expression: (a) an Fd-linked butyrate pathway (But) from C. difficile, (b) an [FeFe]-hydrogenase (Hyd) to modulate the redox state of Fd, and (c) heterologous ferredoxins as electron carriers. The hydrogenase efficiently modulated butyrate formation by H2-mediated Fd reoxidation under nitrogen. In its functionally assembled state, Rnf severely impaired cell growth. Including Hyd in the But/Rnf background, in turn, restored normal growth. Our findings suggest that Rnf mediates reverse electron flow from NADH to Fd, which requires E. coli’s F-type ATPase to function in its reverse, ATP hydrolyzing direction. The reduced Fd is then reoxidized by endogenous Fd:NAD(P)H oxidoreductase (Fpr), which regenerates NADH and, thereby, initiates a futile cycle fueled by ATP hydrolysis. The introduction of hydrogenase interrupts this futile cycle under N2 by providing an efficient NAD(P)+-independent Fd reoxidation route, whereas under H2, Hyd outcompetes Rnf for Fd reduction. This is the first report of an Rnf complex being functionally produced and physiologically investigated in E. coli.  相似文献   

3.
Acidaminococcus fermentans is able to ferment glutamate to ammonia, CO2, acetate, butyrate, and H2. The molecular hydrogen (approximately 10 kPa; E′ = –385 mV) stems from NADH generated in the 3-hydroxybutyryl-CoA dehydrogenase reaction (E°′ = –240 mV) of the hydroxyglutarate pathway. In contrast to growing cells, which require at least 5 mM Na+, a Na+-dependence of the H2-formation was observed with washed cells. Whereas the optimal glutamate fermentation rate was achieved already at 1 mM Na+, H2 formation commenced only at > 10 mM Na+ and reached maximum rates at 100 mM Na+. The acetate/butyrate ratio thereby increased from 2.0 at 1 mM Na+ to 3.0 at 100 mM Na+. A hydrogenase and an NADH dehydrogenase, both of which were detected in membrane fractions, are components of a model in which electrons, generated by NADH oxidation inside of the cytoplasmic membrane, reduce protons outside of the cytoplasmic membrane. The entire process can be driven by decarboxylation of glutaconyl-CoA, which consumes the protons released by NADH oxidation inside the cell. Hydrogen production commences exactly at those Na+ concentrations at which the electrogenic H+/Na+-antiporter glutaconyl-CoA decarboxylase is converted into a Na+/Na+ exchanger. Received: 3 May 1996 / Accepted: 12 August 1996  相似文献   

4.
A psychrotolerant microbial consortium from a low-temperature anaerobic EGSB bioreactor was grown separately on acetate, propionate, butyrate, and H2/CO2 at 30 and 10°C in glass flasks. In the course of the experiments, the cultivation temperature was changed at different time intervals. The initial rates of substrate utilization were higher at 30 than at 10°C. However, the microbial consortium was found to be well adapted to low temperatures; when grown at 10°C for 1.5–5 months, the rates of butyrate, propionate, and H2/CO2 utilization increased steadily. When grown at 30°C for 1.5–2.5 months, this consortium retained its ability to degrade VFA and H2/CO2 at 10°C. However, after long-term (150 days) cultivation at 10°C, its ability to utilize the substrates at 30°C decreased. In the consortium grown in the acetate-containing medium, a Methanosaeta-like methanogen was predominant; in media with propionate and butyrate, besides VFA-degrading bacteria, acetoclastic Methanosaeta-like and hydrogenotrophic Methanospirillum-like methanogenic archaea prevailed. A Methanospirillum-like strain predominated in the H2/CO2-containing medium. The Methanospirillum strain of this microbial community was presumably psychrotolerant. A method based on changes in the cultivation temperature is of practical interest and can be used to start up new bioreactors.  相似文献   

5.
Clostridium acetobutylicum strain P262 utilized lactate at a rapid rate [600 nmol min–1 (mg protein)–1], but lactate could not serve as the sole energy source. When acetate was provided as a co-substrate, the growth rate was 0.05 h–1. Butyrate, carbon dioxide and hydrogen were the end products of lactate and acetate utilization, and the stoichiometry was 1 lactate + 0.4 acetate → 0.7 butyrate + 0.6 H2 + 1 CO2. Lactate-grown cells had twofold lower hydrogenase than glucose-grown cells, and the lactate-grown cells used acetate as an alternative electron acceptor. The cells had a poor affinity for lactate (Ks = 1.1 mM), and there was no evidence for active transport. Lactate utilization was catabolyzed by an inducible NAD-independent lactate dehydrogenase (iLDH) that had a pH optimum of 7.5. The iLDH was fivefold more active with d-lactate than l-lactate, and the K m for d-lactate was 3.2 mM. Lactate-grown cells had little butyraldehyde dehydrogenase activity, and this defect did not allow the conversion of lactate to butanol. Received: 17 October 1994 / Accepted: 30 January 1995  相似文献   

6.
Acetate-degrading anaerobic microorganisms in freshwater sediment were quantified by the most probable number technique. From the highest dilutions a methanogenic, a sulfate-reducing, and a nitrate-reducing microorganism were isolated with acetate as substrate. The methanogen (culture AMPB-Zg) was non-motile and rod-shaped with blunted ends (0.5–1 μm × 3–4 μm long). Doubling times with acetate at 30–35°C were 5.6–8.1 days. The methanogen grew only on acetate. Analysis of the 16S rRNA sequence showed that AMPB-Zg is closely related toMethanosaeta concilii. The isolated sulfate-reducing bacterium (strain ASRB-Zg) was rod-shaped with pointed ends (0.5–0.7 μm × 1.5–3.5 μm long), weakly motile, spore forming, and gram positive. At the optimum growth temperature of 30°C the doubling times with acetate were 3.9–5.3 days. The bacterium grew on a range of organic acids, such as acetate, butyrate, fumarate, and benzoate, but did not grow autotrophically with H2, CO2, and sulfate. The closest relative of strain ASRB-Zg isDesulfotomaculum acetoxidans. The nitrate-reducing bacterium (strain ANRB-Zg) was rod-shaped (0.5–0.7 μm × 0.7–1 μm long), weakly motile, and gram negative. Optimum growth with acetate occurred at 20–25°C. The bacterium grew on a range of organic substrates, such as acetate, butyrate, lactate, and glucose, and did grow autotrophically with H2, CO2, and oxygen but not with nitrate. In the presence of acetate and nitrate, thiosulfate was oxidized to sulfate. Phylogenetically, the closest relative of strain ANRB-Zg isVariovorax paradoxus.  相似文献   

7.
 β-Fructofuranosidase was purified from commercial alkaline protease (Aspergillus oryzae origin). The optimal pH of its transfructosylating activity was more alkaline (pH 8) than that of its hydrolyzing activity (pH 5). In the case of a 24-h reaction with sucrose, the hydrolysis and transfructosylation reaction were optimal at pH 4–5 and pH 8, respectively. In the reaction at pH 8 1-kestose and nystose were the main fructooligosaccharides produced. The transfer ratio was hardly different between pH 5 and pH 8 early in the reaction, but the transfer products (1-kestose and nystose) were decreased at pH 5 as the reaction proceeded because of their hydrolysis. Received: 18 January 1995/Received last revision: 23 August 1995/Accepted: 13 September 1995  相似文献   

8.
 The biofiltration of dimethyl sulphide (Me2S) and other volatile sulphur compounds results in the accumulation of the metabolite sulphuric acid in the carrier material. Regeneration of an acidified (pH 4.7), Hyphomicrobium-MS3-inoculated compost biofilter degrading Me2S was not possible by trickling tap water (days 0–28) or a KH2PO4/K2HPO4 buffer solution (1.26 g PO3- 4 l-1, pH 7) (days 29–47) over the bioreactor at a superficial liquid flow rate of 34 lm-2 day-1. Since the protons produced displaced nutrient cations (Na+, K+, Ca2+, Mg2+, NH+ 4) from the cation-exchange sites on the compost material, 95% of the SO2- 4 was leached as the corresponding sulphate salts and not as sulphuric acid. Concomitantly, the pH of the compost material decreased from 4.7 to 3.9 over the 47 days rinsing period. Moreover, the rinsing procedure resulted in the leaching of essential microbial nutrients from the compost material, such as NH+ 4 (22.3% wash-out over the 47-day rinsing period) and PO3- 4 (39.3% washout over the 28-day tap-water rinsing period). However, mixing limestone powder into the Me2S-degrading compost biofilter was a successful approach to controlling the pH in the optimal range for the inoculum Hyphomicrobium MS3 (pH 6–7). A stoichiometric neutralisation reaction (molar ratio CaCO3/H2SO4=1.1) was observed between the CaCO3 added and the metabolite of the Me2S degradation, while high elimination capacities (above 100 g Me2S m-3 day-1) were obtained over a prolonged (more than 100 days) period. Received: 1 December 1995/Received revision: 26 April 1995 Accepted: 29 April 1996  相似文献   

9.
Since the thermophilic bacterium Moorella sp. HUC22-1 produces 120 mM acetate and 5.2 mM ethanol from H2–CO2, several candidate genes, which were predicted to code for three alcohol dehydrogenases (AdhA, B, C) and one acetaldehyde dehydrogenase (Aldh), were cloned from HUC22-1. The cloned genes were subcloned into a His-tagged expression vector and expressed in Escherichia coli. Recombinant AdhA and B were both dependent on NADP(H) but independent of NAD(H), and their reduction activities from aldehyde to alcohol were higher than their oxidation activities. In contrast with AdhA and B, no activity of AdhC was observed in either reaction. On the other hand, Aldh was active toward both NADP(H) and NAD(H). The enzyme activity of Aldh was directed toward the thioester cleavage and the thioester condensation. When 50 μg of AdhA and 50 μg Aldh were added to the buffer solution (pH 8.0) containing NADPH, NADH and acetyl-CoA at 60°C, 1.6 mM ethanol was produced from 3 mM acetyl-CoA after 90 min. Expression analysis of the mRNAs revealed that the expression level of aldh was threefold higher in the H2–CO2 culture than that in the fructose culture, but levels of adhA, B and C were decreased.  相似文献   

10.
The influence of low temperature (5–29 °C) on the methanogenic activity of non-adapted digested sewage sludge and on temperature/leachate-adapted biomass was assayed by using municipal landfill leachate, intermediates of anaerobic degradation (propionate) and methane precursors (acetate, H2/CO2) as substrates. The temperature dependence of methanogenic activity could be described by Arrhenius-derived models. However, both substrate and adaptation affected the temperature dependence. The adaptation of biomass in a leachate-fed upflow anaerobic sludge-blanket reactor at approximately 20 °C for 4 months resulted in a sevenfold and fivefold increase of methanogenic activity at 11 °C and 22 °C respectively. Both acetate and H2/CO2 were methanized even at 5 °C. At 22 °C, methanogenic activities (acetate 4.8–84 mM) were 1.6–5.2 times higher than those at 11 °C. The half-velocity constant (K s) of acetate utilization at 11 °C was one-third of that at 22 °C while a similar K i was obtained at both temperatures. With propionate (1.1–5.5 mM) as substrate, meth‐anogenic activities at 11 °C were half those at 22 °C. Furthermore, the residual concentration of the substrates was not dependent on temperature. The results suggest that the adaptation of biomass enables the achievement of a high treatment capacity in the anaerobic process even under psychrophilic conditions. Received: 23 December 1996 / Received last revision: 18 June 1997 / Accepted: 23 June 1997  相似文献   

11.
12.
A system is described that combines the fermentation of cellulose to acetate, CH4, and CO2 by Ruminococcus albus and Methanobrevibacter smithii with the fermentation of acetate to CH4 and CO2 by Methanosarcina barkeri to convert cellulose to CH4 and CO2. A cellulose-containing medium was pumped into a co-culture of the cellulolytic R. albus and the H2-using methanogen, Mb. smithii. The effluent was fed into a holding reservoir, adjusted to pH 4.5, and then pumped into a culture of Ms. barkeri maintained at constant volume by pumping out culture contents. Fermentation of 1% cellulose to CH4 and CO2 was accomplished during 132 days of operation with retention times (RTs) of the Ms. barkeri culture of 7.5–3.8 days. Rates of acetate utilization were 9.5–17.3 mmol l−1 day−1 and increased with decreasing RT. The K s for acetate utilization was 6–8 mM. The two-stage system can be used as a model system for studying biological and physical parameters that influence the bioconversion process. Our results suggest that manipulating the different phases of cellulose fermentation separately can effectively balance the pH and ionic requirements of the acid-producing phase with the acid-using phase of the overall fermentation. Received: 7 December 1999 / Received revision: 28 April 2000 / Accepted: 19 May 2000  相似文献   

13.
In a mesophilic (30–35 °C), sulphidogenic, ethanol-fed expanded-granular-sludge-blanket reactor, sulphate, at loading rates of up to 10.0–12.0 g Sl−1␣day−1, was removed with an average efficiency of more than 80%. The pH was between 7.7 and 8.3 and the maximal total dissolved sulphide concentration was up to 20 mM S (650 mg S/l). The alkaline pH was maintained by either a pH-control unit with sodium hydroxide or by stripping part of the sulphide and CO2 from the recycle with nitrogen gas. The superficial upstream liquid velocity (v up) was 3.0–4.5 m/h. The ratio of ethanol to sulphur was near stoichiometry. At alkaline pH, the activity of the acetotrophic sulphate-reducing bacteria, growing on acetate, was strongly enhanced, whereas at pH below 7.7 the acetotrophic sulphate-reducing bacteria were inhibited by aqueous H2S. With regard to the removal efficiency and operational stability, external stripping with N2 and pH control were equally successful. Received: 2 December 1996 / Received revision: 13 March 1997 / Accepted: 15 March 1997  相似文献   

14.
Acetate-grown cells of Methanosarcina barkeri MS were found to form methane from H2:CO2 at the same rate as hydrogen-grown cells. Cells grown on acetate had similar levels of soluble F420-reactive hydrogenase I, and higher levels of cytochrome-linked hydrogenase II compared to hydrogen-grown cells. The hydrogenase I and II activities in the crude extract of acetate-grown cells were separated by differential binding properties to an immobilized Cu2+ column. Hydrogenase II did not react with ferredoxin or F420, whereas hydrogenase I coupled to both ferredoxin and F420. A reconstituted soluble protein system composed of purified CO dehydrogenase, F420-reactive hydrogenase I fraction, and ferredoxin produced H2 from CO oxidation at a rate of 2.5 nmol/min · mg protein. Membrane-bound hydrogenase II coupled H2 consumption to the reduction of CoM-S-S-HTP and the synthesis of ATP. The differential function of hydrogenase I and II is ascribed to ferredoxin-linked hydrogen production from CO and cytochrome b-linked H2 consumption coupled to methanogenesis and ATP synthesis, respectively.  相似文献   

15.
The cellular characteristics of a new methylotrophic, acidogenic, anaerobic bacterium that was first isolated from a sewage digestor in Marburg, Federal Republic of Germany, is described. The Marburg strain is a mesophilic, Gram-positive, nonmotile, pleomorphic rod that performs homoacetic, homobutyric, or heteroacidic fermentations. Cell morphology varies from single or paired straight rods to rudimentary branched rods, club-shaped cells, or oval refractile cells. Cell heat resistance correlated with the presence of a few refractile cells. Electron micrographs of thin sections revealed a thick monolayered cell wall and an atypical spore structure. The DNA base composition was 48.8±0.2 mol% guanosine plus cytosine. Growth required factors in yeast extract; methanol, H2/CO2, glucose, fructose, lactate, and pyruvate were fermented as energy sources. Corrinoid levels varied from 0.35±0.16 to 7.9±1.6 μg/mg cell dry weight when cells 0.1% yeast extract, N2/CO2, 100 mM methanol, and 50 mM Na acetate displayed a 20h doubling time, finalA 540 of 0.9, butyric acid yield of 25 mM, and ≈stoichiometry of 3 mol butyrate formed per 10 mol methanol fermented. The nameButyribacterium (emend.)methylotrophicum sp. nov. is proposed for the Marburg strain.  相似文献   

16.
Washed membranes prepared from H2+CO2- or formate-grown cells of Methanococcus voltae catalyzed the oxidation of coenzyme F420H2 and the reduction of the heterodisulfide (CoB–S–S–CoM) of 2-mercaptoethanesulfonate and 7-mercaptoheptanoylthreonine phosphate, which is the terminal electron acceptor of the methanogenic pathway. The reaction followed a 1:1 stoichiometry according to the equation: F420H2 + COB–S–S–CoM → F420 + CoM–SH + CoB–SH. These findings indicate that the reaction depends on a membrane-bound F420H2-oxidizing enzyme and on the heterodisulfide reductase, which remains partly membrane-bound after cell lysis. To elucidate the nature of the F420H2-oxidizing protein, washed membranes were solubilized with detergent, and the enzyme was purified by sucrose density centrifugation, anion-exchange chromatography, and gel filtration. Several lines of evidence indicate that F420H2 oxidation is catalyzed by a membrane-associated F420-reducing hydrogenase. The purified protein catalyzed the H2-dependent reduction of methyl viologen and F420. The apparent molecular mass and the subunit composition (43, 37, and 27 kDa) are almost identical to those of the F420-reducing hydrogenase that has already been purified from Mc. voltae. Moreover, the N-terminus of the 37-kDa subunit is identical to the amino acid sequence deduced from the fruG gene of the operon encoding the selenium-containing F420-reducing hydrogenase from Mc. voltae. A distinct F420H2 dehydrogenase, which is present in methylotrophic methanogens, was not found in this organism. Received: 18 September 1998 / Accepted: 2 November 1998  相似文献   

17.
In the anaerobic fungus Neocallimastix sp. L2 fermentation of glucose proceeds via the Embden-Meyerhof-Parnas pathway. Enzyme activities leading to the formation of succinate, lactate, ethanol, and formate are associated with the cytoplasmic fraction. The enzymes malic enzyme, NAD(P)H: ferredoxin oxidoreductase, pyruvate: ferredoxin oxidoreductase, hydrogenase, acetate: succinate CoA transferase and succinate thiokinase leading to the formation of H2, CO2, acetate, and ATP are localized in microbodies. Thus, these organelles are identified as hydrogenosomes. In addition, the microbodies contain the O2-scavenging enzymes NADH- and NADPH oxidase, while NAD(P)H peroxidase, catalase, or superoxide dismutase could not be detected. In cell-free extracts from zoospores of Neocallimastix sp. L2 the specific activities of hydrogenosomal enzymes as well as the quantities of these proteins are 2- to 6-fold higher than in mycelium extracts. These findings suggest that hydrogenosomes perform an important role-especially in zoospores — as H2-evolving, ATP-generating and O2-scavenging organelles.Abbrevations DTT Dithiotreitol - PEP Phosphoenol pyruvate  相似文献   

18.
Enrichment cultures inoculated with black mud fermented benzoate according to the stoichiometric equation: 4 C6H5CO2H+18 H2O 15 CH4+13 CO2.Trans-2-hydroxycyclohexanecarboxylate, 2-oxo-cyclohexanecarboxylate, pimelate, caproate, butyrate, acetate, and molecular hydrogen were shown to be regular components of the culture fluid occurring in low concentrations. Inhibition of methanogenesis by chloroform, 4-chlorobutyrate, or 2-bromooctanoate resulted in a cessation of the benzoate breakdown after all intermediates had accumulated. It is proposed that benzoate is fermented via a direct reductive pathway to butyrate, acetate, H2, and CO2, whereafter butyrate is converted to acetate and H2, and the latter substrates are fermented to CH4 and CO2 by methane producers.  相似文献   

19.
The hyperthermophilic and anaerobic bacterium Thermotoga maritima ferments a wide variety of carbohydrates, producing acetate, CO2, and H2. Glucose is degraded through a classical Embden-Meyerhof pathway, and both NADH and reduced ferredoxin are generated. The oxidation of these electron carriers must be coupled to H2 production, but the mechanism by which this occurs is unknown. The trimeric [FeFe]-type hydrogenase that was previously purified from T. maritima does not use either reduced ferredoxin or NADH as a sole electron donor. This problem has now been resolved by the demonstration that this hydrogenase requires the presence of both electron carriers for catalysis of H2 production. The enzyme oxidizes NADH and ferredoxin simultaneously in an approximately 1:1 ratio and in a synergistic fashion to produce H2. It is proposed that the enzyme represents a new class of bifurcating [FeFe] hydrogenase in which the exergonic oxidation of ferredoxin (midpoint potential, −453 mV) is used to drive the unfavorable oxidation of NADH (E0′ = −320 mV) to produce H2 (E0′ = −420 mV). From genome sequence analysis, it is now clear that there are two major types of [FeFe] hydrogenases: the trimeric bifurcating enzyme and the more well-studied monomeric ferredoxin-dependent [FeFe] hydrogenase. Almost one-third of the known H2-producing anaerobes appear to contain homologs of the trimeric bifurcating enzyme, although many of them also harbor one or more homologs of the simpler ferredoxin-dependent hydrogenase. The discovery of the bifurcating hydrogenase gives a new perspective on our understanding of the bioenergetics and mechanism of H2 production and of anaerobic metabolism in general.The order Thermotogales is characterized by the ability of its members to utilize a wide variety of carbohydrates (8). All of these organisms ferment sugars predominantly to acetate, CO2, and H2 (23). They thrive mainly at elevated temperatures, although a new subclass of mesophilic “mesotoga” has also been proposed (19). These properties also make the Thermotoga species excellent candidates for biohydrogen production from plant-based biomass. The genome of the type strain, T. maritima, was one of the first to be sequenced, and this revealed a high degree of lateral gene transfer between archaea and bacteria (17, 18). In addition, T. maritima is part of a structural genomics effort, and the structures of over 100 of its proteins have been determined (20, 21). The organism degrades a wide variety of both simple and complex carbohydrates (4, 5), and the glucose that is produced is oxidized by both classical Embden-Meyerhof (85%) and Entner-Douderhoff (15%) pathways (23). The generation of H2 is accomplished by the enzyme hydrogenase. However, little is known about the bioenergetics of the reaction and the pathways of electron flow from carbohydrate oxidation to H2 formation.Although hydrogenases catalyze the simplest of chemical reactions, the reversible interconversion of protons, electrons, and H2, they are surprisingly complex proteins, some more so than others (33). They can be divided into two major groups, the [NiFe]- and [FeFe]-type hydrogenases, based on the presence of nickel and iron or only iron in their active sites. In general, the physiological roles of the [FeFe] hydrogenases are to evolve H2, while the roles of the [NiFe] enzymes are to oxidize it (33). For example, several Clostridium spp. evolve H2 via a cytoplasmic, monomeric [FeFe] hydrogenase that uses the low-potential redox protein ferredoxin (Fd) (midpoint potential [Em], <−400 mV) as the electron donor (15). In contrast, H2 production using NAD(P)H (E0′ = −320 mV) as the electron donor is thermodynamically unfavorable under physiological conditions because of the more positive redox potential of the pyridine nucleotides (30). Nevertheless, cytoplasmic NAD(P)H-dependent [FeFe] hydrogenases have been reported, although how the endergonic reaction of NAD(P)H-dependent H2 production is accomplished under physiological conditions is not clear (13, 28).During the oxidation of glucose by T. maritima, both Fd and NAD function as physiological electron acceptors (1, 26, 34). NADH is generated via the glyceraldehyde-3-phosphate dehydrogenase reaction of glycolysis, while the pyruvate that is generated by this pathway is oxidized by pyruvate Fd oxidoreductase (POR) to acetyl coenzyme A (acetyl-CoA), producing reduced Fd. Acetyl-CoA is converted to acetate by phosphotransacetylase and acetate kinase with the concomitant production of ATP. This pathway leads to the production of four moles of H2 per mole of glucose, with reductant provided by two moles of NADH and four moles of reduced Fd, together with two moles of acetate and two moles of CO2 (23). The oxidation of reduced Fd and NADH must be directly or indirectly coupled to the reduction of protons to H2 by hydrogenase, but the trimeric cytoplasmic [FeFe] hydrogenase characterized from T. maritima more than a decade ago does not use either T. maritima Fd or NADH as the sole electron donor (10, 31). Consequently, the mechanism by which the oxidation of Fd and NADH is coupled in vivo to H2 production is not known. In this study, we have resolved this long-standing problem by showing that this cytoplasmic enzyme represents a novel type of hydrogenase that requires both physiological electron carriers to be present for the efficient catalysis of H2 production in which both serve as electron donors.  相似文献   

20.
The biochemical mechanisms for growth tolerance to a 100% CO headspace in cultures, and butanol plus ethanol production from CO by Butyribacterium methylotrophicum were assessed in the wild-type and CO-adapted strains. The CO-adapted strain grew on glucose or CO under a 100% CO headspace, whereas, the growth of the wild-type strain was severely inhibited by 100% CO. The CO-adapted strain, unlike the wild-type, also produced butyrate, from either pyruvate or CO. The CO-adapted strain was a metabolic mutant having higher levels of ferredoxin–NAD oxidoreductase activity, which was not inhibited by NADH. Consequently, only the CO-adapted strain can grow on CO because CO oxidation generates reduced ferredoxin which, via the mutated ferredoxin–NAD reductase activity, forms reduced NADH required for catabolism. When the CO-adapted strain was grown at pH 6.0 it produced butanol (0.33 g/l) and ethanol (0.5 g/l) from CO and the cells contained the following NAD-linked enzyme activities (μmol min−1 mg protein−1): butyraldehyde dehydrogenase (227), butanol dehydrogenase (686), acetaldehyde dehydrogenase (82) and ethanol dehydrogenase (129). Received: 15 September 1998 / Received revision: 12 February 1999 / Accepted: 19 February 1999  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