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.
Whole cells of Desulfobulbus propionicus fermented [1-13C]ethanol to [2-13C] and [3-13C]propionate and [1-13C]-acetate, which indicates the involvement of a randomizing pathway in the formation of propionate. Cell-free extracts prepared from cells grown on lactate (without sulfate) contained high activities of methylmalonyl-CoA: pyruvate transacetylase, acetase kinase and reasonably high activities of NAD(P)-independent L(+)-lactate dehydrogenase NAD(P)-independent pyruvate dehydrogenase, phosphotransacetylase, acetate kinase and reasonably high activity of NAD(P)-independent L(+)-lactate dehydrogenase, fumarate reductase and succinate dehydrogenase. Cell-free extracts catalyzed the conversion of succinate to propionate in the presence of pyruvate, CoA and ATP and the oxaloacetate-dependent conversion of propionate to succinate. After growth on lactate or propionate in the presence of sulfate similar enzyme levels were found except for fumarate reductase which was considerably lower. Fermentative growth on lactate led to higher cytochrome b contents than growth with sulfate as electron acceptor.The labeling studies and the enzyme measurements demonstrate that in Desulfobulbus propionate is formed via a succinate pathway involving a transcarboxylase like in Propionibacterium. The same pathway may be used for the degradation of propionate to acetate in the presence of sulfate.Abbreviations DCPIP 2,6-dichlorophenolindophenol - PEP phosphoenolpyruvate  相似文献   

2.
Under anaerobic conditions, cells of Entamoeba histolytica grown with bacteria produce H2 and acetate while cells grown axenically produce neither. Aerobically, acetate is produced and O2 is consumed by amebae from either type of cells. Centrifuged extracts, 2.4 x 106 x g x min, from both types of cells contain pyruvate synthase (EC 1.2.7.1) and an acetate thiokinase which, together, form a system capable of converting pyruvate to acetate. Pyruvate synthase catalyzes the reaction: pyruvate + CoA leads to CO2 + acetyl-CoA + 2E. Electron acceptors which function with this enzyme are FAD, FMN, riboflavin, ferredoxin, and methyl viologen, but not NAD or NADP. The amebal acetate thiokinase catalyzes the reaction acetyl-CoA + ADP + Pi leads to acetate + ATP + CoA. For this apparently new enzyme we suggest the trivial name acetyl-CoA-synthetase (ADP-forming). Extracts from axenic amebae do not contain hydrogenase, but extracts from cells grown with bacteria do. It is postulated that in bacteria-grown amebae electrons generated at the pyruvate synthase step are utilized anaerobically to produce H2 via the hydrogenase and that the acetyl-CoA is converted to acetate in an energy-conserving step catalyzed by amebal acetyl-CoA synthetase. Aerobically, cells grown under either regimen may utilize the energy-conserving pyruvate-to-acetate pathway since O2 then serves as the ultimate electron acceptor.  相似文献   

3.
A trithionate reductase system was isolated and purified from extracts of Desulfovibrio vulgaris. This system reduced trithionate to thiosulfate and consisted of two proteins. One was bisulfite reductase, an enzyme that reduces bisulfite to trithionate, and the second component was designated TR-1. Both enzymes were required to reduce trithionate to thiosulfate. Flavodoxin and cytochrome c3 from D. vulgaris were tested for their ability to function as electron carriers during trithionate reduction. When molecular hydrogen was the source of electrons for the reduction, both flavodoxin and cytochrome c3 were required. In contrast, when the pyruvate phosphoroclastic system was the reductant, flavodoxin alone participated as the electron carrier. The results indicate that flavodoxin, but not cytochrome c3, interacted with the trithionate reductase system. The cytochrome in the hydrogenase-linked assay functioned as an electron carrier between hydrogenase and flavodoxin.  相似文献   

4.
The small flavoprotein, flavodoxin, isolated from Peptostreptococcus elsdenii, has been covalently coupled to CNBr-activated Sepharose 4B. The immobilized protein replaces ferredoxin as an electron carrier in hydrogen production from dithionite or pyruvate in the presence of ferredoxin-free extracts of P. elsdenii; compared with soluble flavodoxin, its activities in these systems are 13% and 3.5% respectively. Acid treatment reversibly dissociates FMN from the immobilized protein. The dissociation constant of the complex with FMN, determined by fluorimetric titration, is 1.5 (+/- 0.4) nM, and is therefore very little different from that of soluble flavodoxin. Like soluble apoflavodoxin, the immobilized apoprotein is highly specific for flavins with an N-10 side-chain of 5 carbon atoms and a C-5' phosphate group. Approximately half of the flavin impurity in commercial preparations of FMN (12-15% of the total flavin), and similar impurity in synthetic analogues of FMN, is not separated by conventional purification procedures, but it is readily and conveniently removed by affinity chromatography with apoflavodoxin as the immobilized ligand. The immobilized protein is stable for long periods; its capacity for FMN decreases by only 20% after 15 cycles of flavin dissociation and reassociation during several months.  相似文献   

5.
Cénas N  Lê KH  Terrier M  Lederer F 《Biochemistry》2007,46(15):4661-4670
Saccharomyces cerevisiae flavocytochrome b2 (L-lactate:cytochrome c oxido reductase, EC 1.1.2.3) is a homotetramer, with FMN and protoheme IX binding on separate domains. The flavin-binding domains form the enzyme tetrameric core, while the cytochrome b2 domains appear to be mobile around a hinge region (Xia, Z. X., and Mathews, F. S. (1990) J. Mol. Biol. 212, 867-863). The enzyme catalyzes electron transfer from L-lactate to cytochrome c, or to nonphysiological acceptors such as ferricyanide, via FMN and heme b2. The kinetics of this multistep reaction are complex. In order to clarify some of its aspects, the tetrameric FMN-binding domain (FDH domain) has been independently expressed in Escherichia coli (Balme, A., Brunt, C. E., Pallister, R., Chapman, S. K., and Reid, G. A. (1995) Biochem. J. 309, 601-605). We present here an additional characterization of this domain. In our hands, it has essentially the same ferricyanide reductase activity as the holo-enzyme. The comparison of the steady-state kinetics with ferricyanide as acceptor and of the pre-steady-state kinetics of flavin reduction, as well as the kinetic isotope effects of the reactions using L-2-[2H]lactate, indicates that flavin reduction is the limiting step in lactate oxidation. During the oxidation of the reduced FDH domain by ferricyanide, the oxidation of the semiquinone is much faster than the oxidation of two-electron-reduced flavin. This order of reactivity is reversed during flavin to heme b2 transfer in the holo-enzyme. Potentiometric studies of the protein yielded a standard redox potential for FMN at pH 7.0, E(o)7, of -81 mV, a value practically identical to the published value of -90 mV for FMN in holo-flavocytochrome b2. However, as expected from the kinetics of the oxidative half-reaction, the FDH domain was characterized by a significantly destabilized flavin semiquinone state compared with holo-enzyme, with a semiquinone formation constant K of 1.25-0.64 vs 33.5, respectively (Tegoni, M., Silvestrini, M. C., Guigliarelli, B., Asso, M., and Bertrand, P. (1998) Biochemistry, 37, 12761-12771). As in the holo-enzyme, the semiquinone state in the FDH domain is significantly stabilized by the reaction product, pyruvate. We also studied the inhibition exerted in the steady and pre steady states by the reaction product pyruvate and by anions (bromide, chloride, phosphate, acetate), with respect to both flavin reduction and reoxidation. The results indicate that these compounds bind to the oxidized and the two-electron-reduced forms of the FDH domain, and that excess L-lactate also binds to the two-electron-reduced form. These findings point to the existence of a common or strongly overlapping binding site. A comparison of the effect of the anions on WT and R289K holo-flavocytochromes b2 indicates that invariant R289 belongs to this site. According to literature data, it must also be present in other members of the family of L-2-hydroxy acid-oxidizing enzymes.  相似文献   

6.
Flavodoxin from Desulfovibrio vulgaris (Hildenborough) has been expressed at a high level (3-4% soluble protein) in Escherichia coli by subcloning a minimal insert carrying the gene behind the tac promoter of plasmid pDK6. The recombinant protein was readily isolated and its properties were shown to be identical to those of the wild-type protein obtained directly from D. vulgaris, with the exception that the recombinant protein lacks the N-terminal methionine residue. Detailed measurements of the redox potentials of this flavodoxin are reported for the first time. The redox potential, E2, for the couple oxidized flavodoxin/flavodoxin semiquinone at pH 7.0 is -143 mV (25 degrees C), while the value for the flavodoxin semiquinone/flavodoxin hydroquinone couple (E1) at the same pH is -440 mV. The effects of pH on the observed potentials were examined; E2 varies linearly with pH (slope = -59 mV), while E1 is independent of pH at high pH values, but below pH 7.5 the potential becomes less negative with decreasing pH, indicating a redox-linked protonation of the flavodoxin hydroquinone. D. vulgaris apoflavodoxin binds FMN very tightly, with a value of 0.24 nM for the dissociation constant (Kd) at pH 7.0 and 25 degrees C, similar to that observed with other flavodoxins. In addition, the apoflavodoxin readily binds riboflavin (Kd = 0.72 microM; 50 mM sodium phosphate, pH 7.0, 5 mM EDTA at 25 degrees C) and the complex is spectroscopically very similar to that formed with FMN. The redox potentials for the riboflavin complex were determined at pH 6.5 (E1 = -262 mV, E2 = -193 mV; 25 degrees C) and are discussed in the light of earlier proposals that charge/charge interactions between different parts of the flavin hydroquinone play a crucial role in determining E1 in flavodoxin.  相似文献   

7.
Murray TA  Swenson RP 《Biochemistry》2003,42(8):2307-2316
The pathway(s) by which the flavin cofactor binds to the apoflavoprotein is the subject of some debate. The crystal and NMR structures of several different flavodoxins have provided some insight, although there is disagreement about the location of the initial interaction between the flavin mononucleotide (FMN) and the apoflavodoxin and the degree of protein conformational change associated with cofactor binding [Genzor, C. G., Perales-Alcon, A., Sancho, J., and Romero, A. (1996) Nat. Struct. Biol. 3, 329-332; Steensma, E., and van Mierlo, C. P. M. (1998) J. Mol. Biol. 282, 653-666]. Binding kinetics using stopped-flow spectrofluorimetry and phosphate competition studies were used to develop a model for flavin binding to the flavodoxin from Desulfovibrio vulgaris. In the presence of phosphate, the time course of fluorescence quenching associated with FMN binding to apoflavodoxin was biphasic, whereas riboflavin, which lacks the 5'-phosphate group of FMN, displayed monophasic binding kinetics. When the concentration of phosphate in solution was increased, the FMN binding rates of the two phases behaved differently; the rate of one phase decreased, while the rate of the other increased. A similar increase in the single phase associated with riboflavin binding was also observed. This has led to the following model. The binding of the flavin isoalloxazine ring to its subsite is dependent on the presence of a phosphate group in the phosphate-binding subsite. When phosphate is in the buffer solution, FMN can bind in either of two ways: by the initial insertion of the 5'-phosphate group followed by ring binding or, when inorganic phosphate from solution is bound, the insertion of the isoalloxazine ring first. Riboflavin, which lacks the phosphate moiety of FMN, binds only in the presence of inorganic phosphate, presumably due to the binding of this group in the phosphate-binding subsite. These results suggest that cooperative interactions exist between the phosphate subsite and the ring-binding region in the D. vulgaris flavodoxin that are necessary for isoalloxazine ring binding.  相似文献   

8.
Pyruvate:NADP+ oxidoreductase was homogeneously purified from crude extract of Euglena gracilis. The Mr of the enzyme was estimated to be 309,000 by gel filtration. The enzyme migrated as a single protein band with Mr of 166,000 by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, suggesting that the enzyme consists of two identical polypeptides. The absorption spectrum of the native enzyme exhibited maxima at 278, 380, and 430 nm, and a broad shoulder was observed around 480 nm; the maximum at 430 nm was eliminated by reduction of the enzyme with dithionite. Reduction of the enzyme with pyruvate and CoA and reoxidation with NADP+ were proved from changes of absorption spectra. The enzyme contained 2 molecules of FAD and 8 molecules of iron. It was also indicated that the enzyme was thiamine pyrophosphate-dependent. The enzyme was oxygen-sensitive, and the reaction was affected by the presence of oxygen. Pyruvate was the most active substrate, but the enzyme was slightly active for 2-oxobutyrate, 3-hydroxypyruvate, and oxalacetate, but not for glyoxylate and 2-oxoglutarate. The native electron acceptor was NADP+, whereas NAD+ was completely inactive. Methyl viologen, benzyl viologen, FAD, and FMN were utilized as artificial electron acceptors, whereas spinach and Clostridium ferredoxins were inactive. Pyruvate synthesis by reductive carboxylation of acetyl-CoA with NADPH as the electron donor occurred by the reverse reaction of the enzyme. The enzyme also catalyzed a pyruvate-CO2 exchange reaction and electron-transfer reaction from NADPH to other electron acceptors like methyl viologen. These results indicate that pyruvate:NADP+ oxidoreductase in E. gracilis is clearly distinct from either the pyruvate dehydrogenase multienzyme complex or pyruvate:ferredoxin oxidoreductase.  相似文献   

9.
Lactobacillus plantarum P5 grew aerobically in rich media at the expense of lactate; no growth was observed in the absence of aeration. The oxygen-dependent growth was accompanied by the conversion of lactate to acetate which accumulated in the growth medium. Utilization of oxygen with lactate as substrate was observed in buffered suspensions of washed whole cells and in cell-free extracts. A pathway which accounts for the generation of adenosine triphosphate during aerobic metabolism of lactate to acetate via pyruvate and acetyl phosphate is proposed. Each of the enzyme activities involved, nicotinamide adenine dinucleotide independent lactic dehydrogenase, nicotinamide adenine dinucleotide dependent lactic dehydrogenase, pyruvate oxidase, acetate kinase and NADH oxidase were demonstrated in cell-free extracts. The production of pyruvate, acetyl phosphate and acetate was demonstrated using cell-free extracts and cofactors for the enzymes of the proposed pathway.Abbreviations MRS Man, Rogosa and Sharpe (1960) medium modified as in Materials and methods - TY Tryptone Yeast Extract broth - OUL Oxygen uptake with lactate as substrate - DCPIP 2,6-Dichlorophenolindophenol - LDH Lactic dehydrogenase  相似文献   

10.
Two ferredoxins, Fd I and Fd II, were isolated and purified from Desulfovibrio vulgaris Miyazaki. The major component, Fd I, is an iron-sulfur protein of Mr 12,000, composed of two identical subunits. The absorption spectra of Fd I and Fd II have a broad absorption shoulder near 400 nm characteristic of iron-sulfur proteins. The purity index, A400/A280, of Fd I is 0.69, and its millimolar absorption coefficient at 400 nm is 3.73 per Fe. It contains two redox centers with discrete redox behaviors. The amino acid composition and the N-terminal sequence of Fd I are similar to those of Fd III of Desulfovibrio africanus Benghazi and Fd II of Desulfovibrio desulfuricans Norway. Fd I does not serve as an electron carrier for the hydrogenase of D. vulgaris Miyazaki, but it serves as a carrier for pyruvate dehydrogenase of this bacterium. The evolution of H2 from pyruvate was observed by a reconstructed system containing purified hydrogenase, cytochrome C3, Fd I, partially purified pyruvate dehydrogenase, and CoA. The H2-sulfite reducing system can be reconstructed from the purified hydrogenase, cytochrome C3, Fd I and desulfoviridin (sulfite reductase), but the reaction rate is very slow compared to that of the crude extract at the same molar ratio of the components.  相似文献   

11.
1. Adipose tissues from rats fed a balanced diet were incubated in the presence of glucose (20mm) with the following additions: insulin, anti-insulin serum, insulin+acetate, insulin+pyruvate, insulin+lactate, insulin+phenazine methosulphate, insulin+oleate+albumin, insulin+adrenaline+albumin, insulin+6-N-2'-O-dibutyryl 3':5'-cyclic AMP+albumin. 2. Measurements were made of the whole tissue concentrations of adenine nucleotides, hexose phosphates, triose phosphates, glycerol 1-phosphate, 3 phosphoglycerate, 6-phosphogluconate, long-chain fatty acyl-CoA, acid-soluble CoA, citrate, isocitrate, malate and 2-oxoglutarate, and of the release into the incubation medium of lactate, pyruvate and glycerol after 1h of incubation. 3. Fluxes of [(14)C]glucose carbon through the major pathways of glucose metabolism were calculated from the yields of (14)C in various products after 2h of incubation. Fluxes of [(14)C]acetate, [(14)C]pyruvate or [(14)C]lactate carbon in the presence of glucose were also determined. 4. Measurements were also made of the whole-tissue concentrations of metabolites in tissues taken directly from Nembutal-anaesthetized rats. 5. Whole tissue mass-action ratios for phosphofructokinase, phosphoglucose isomerase and the combined (aldolasextriose phosphate isomerase) reaction were similar in vivo and in vitro. The reactants of phosphofructokinase appeared to be far from mass-action equilibrium. In vitro, the reactants of hexokinase also appeared to be far from mass-action equilibrium. 6. Correlation of observed changes in glycolytic flux with changes in fructose 6-phosphate concentration suggested that phosphofructokinase may show regulatory behaviour. The enzyme appeared to be activated in the presence of oleate or adrenaline and to be inhibited in the presence of lactate or pyruvate. 7. Evidence is presented that the reactants of lactate dehydrogenase and glycerol 1-phosphate dehydrogenase may be near to mass-action equilibrium in the cytoplasm. 8. No satisfactory correlations could be drawn between the whole-tissue concentrations of long-chain fatty acyl-CoA, citrate and glycerol 1-phosphate and the observed rates of triglyceride and fatty acid synthesis. Under the conditions employed, the concentration of glycerol 1-phosphate appeared to depend mainly on the cytoplasmic [NAD(+)]/[NADH] ratios. 9. Calculated hexose monophosphate pathway flux rates roughly correlated with fatty acid synthesis rates and with whole tissue [6-phosphogluconate]/[glucose 6-phosphate] ratios. The relative rates of production of NADPH for fatty acid synthesis by the hexose monophosphate pathway and by the ;malic enzyme' are discussed. It is suggested that all NADH produced in the cytoplasm may be used in that compartment for reductive synthesis of fatty acids, lactate or glycerol 1-phosphate.  相似文献   

12.
1. A method was devised for preparing pig heart pyruvate dehydrogenase free of thiamin pyrophosphate (TPP), permitting studies of the binding of [35S]TPP to pyruvate dehydrogenase and pyruvate dehydrogenase phosphate. The Kd of TPP for pyruvate dehydrogenase was in the range 6.2-8.2 muM, whereas that for pyruvate dehydrogenase phosphate was approximately 15 muM; both forms of the complex contained about the same total number of binding sites (500 pmol/unit of enzyme). EDTA completely inhibited binding of TPP; sodium pyrophosphate, adenylyl imidodiphosphate and GTP, which are inhibitors (competitive with TPP) of the overall pyruvate dehydrogenase reaction, did not appreciably affect TPP binding. 2. Initial-velocity patterns of the overall pyruvate dehydrogenase reaction obtained with varying TPP, CoA and NAD+ concentrations at a fixed pyruvate concentration were consistent with a sequential three-site Ping Pong mechanism; in the presence of oxaloacetate and citrate synthase to remove acetyl-CoA (an inhibitor of the overall reaction) the values of Km for NAD+ and CoA were 53+/- 5 muM and 1.9+/-0.2 muM respectively. Initial-velocity patterns observed with varying TPP concentrations at various fixed concentrations of pyruvate were indicative of either a compulsory order of addition of substrates to form a ternary complex (pyruvate-Enz-TPP) or a random-sequence mechanism in which interconversion of ternary intermediates is rate-limiting; values of Km for pyruvate and TPP were 25+/-4 muM and 50+/-10 nM respectively. The Kia-TPP (the dissociation constant for Enz-TPP complex calculated from kinetic plots) was close to the value of Kd-TPP (determined by direct binding studies). 3. Inhibition of the overall pyruvate dehydrogenase reaction by pyrophosphate was mixed non-competitive versus pyruvate and competitive versus TPP; however, pyrophosphate did not alter the calculated value for Kia-TPP, consistent with the lack of effect of pyrophosphate on the Kd for TPP. 4. Pyruvate dehydrogenase catalysed a TPP-dependent production of 14CO2 from [1-14C]pyruvate in the absence of NAD+ and CoA at approximately 0.35% of the overall reaction rate; this was substantially inhibited by phosphorylation of the enzyme both in the presence and absence of acetaldehyde (which stimulates the rate of 14CO2 production two- or three-fold). 5. Pyruvate dehydrogenase catalysed a partial back-reaction in the presence of TPP, acetyl-CoA and NADH. The Km for TPP was 4.1+/-0.5 muM. The partial back-reaction was stimulated by acetaldehyde, inhibited by pyrophosphate and abolished by phosphorylation. 6. Formation of enzyme-bound [14C]acetylhydrolipoate from [3-14C]pyruvate but not from [1-14C]acetyl-CoA was inhibited by phosphorylation. Phosphorylation also substantially inhibited the transfer of [14C]acetyl groups from enzyme-bound [14C]acetylhydrolipoate to TPP in the presence of NADH. 7...  相似文献   

13.
Studies of respiration on glucose in procyclic Trypanosoma congolense in the presence of rotenone, antimycin, cyanide, salicylhydroxamic acid and malonate have indicated the presence of NADH dehydrogenase, cytochrome b-c1, cytochrome aa3, trypanosome alternate oxidase and NADH fumarate reductase/succinate dehydrogenase pathway that contributes electrons to coenzyme Q of the respiratory chain. The rotenone sensitive NADH dehydrogenase, the trypanosome alternate oxidase, and cytochrome aa3 accounted for 24.5 +/- 6.5, 36.2 +/- 4.2 and 54.1 +/- 5.5% respectively of the total respiration. Activities of lactate dehydrogenase, NAD(+)-linked malic enzyme and pyruvate kinase were less than 6 nanomoles/min/mg protein suggesting that they play a minor role in energy metabolism of the parasite. Phosphoenolpyruvate carboxykinase, pyruvate dehydrogenase, succinate dehydrogenase, NADP(+)-linked malic enzyme, NADH fumarate reductase, malate dehydrogenase, and alpha-ketoglutarate dehydrogenase and glycerol kinase on the other hand had specific activities greater than 60 nanomoles/min/mg protein. These enzyme activities could account for the production of pyruvate, acetate, succinate and glycerol. The results further show that the amount of glycerol produced was 35-48% of the combined total of pyruvate, acetate and succinate produced. It is apparent that some of the glycerol 3-phosphate produced in glycolysis in the presence of salicylhydroxamic acid is dephosphorylated to form glycerol while the rest is oxidised via cytochrome aa3 to form acetate, succinate and pyruvate.  相似文献   

14.
The complete amino acid sequence for the 148-amino acid flavodoxin from Desulfovibrio vulgaris was determined to be: H3N+-Met-Pro-Lys-Ala-Leu-Ile-Val-Tyr-Gly-Ser-Thr-Thr-Gly-Asn-Thr-Glu-Tyr-Thr-Ala-Glu-Thr-Ile-Ala-Arg-Glu-Leu-Ala-Asn-Ala-Gly-Tyr-Glu-Val-Asp-Ser-Arg-Asp-Ala-Ala-Ser-Val-Glu-Ala-Gly-Gly-Leu-Phe-Glu-Gly-Phe-Asp-Leu-Val-Leu-Leu-Gly-Cys-Ser-Thr-Trp-Gly-Asp-Asp-Ser-Ile-Glu-Leu-Gln-Asp-Asp-Phe-Ile-Pro-Leu-Phe-Asp-Ser-Leu-Glu-Glu-Thr-Gly-Ala-Gln-Gly-Arg-Lys-Val-Ala-Cys-Phe-Gly-Cys-Gly-Asp-Ser-Ser-Tyr-Glu-Tyr-Phe-Cys-Gly-Ala-Val-Asp-Ala-IleGlu-Glu-Lys-Leu-Lys-Asn-Leu-Gly-Ala-Glu-Ile-Val-Gln-Asp-Gly-Leu-Arg-Ile-Asp-Gly-Asp-Pro-Arg-Ala-Ala-Arg-Asp-Asp-Ile-Val-Gly-Try-Ala-His-Asp-Val-Arg-Gly-Ala-Ile-COO. This protein is of interest as it was the first flavoenzyme for which high resolution x-ray diffraction studies were published (Watenpaugh, K.D., Sieker, L.C., and Jensen, L.H. (1973) Proc. NAtl. Acad. Sci. U.S.A. 70, 3857-3860). Ser(10), Thr(12), Asn(14), and Thr(15) were shown to bind the phosphate of the FMN while the isoalloxazine ring is positioned between Trp(60) and Tyr(98).  相似文献   

15.
1. The role of pyruvate carboxylation in the net synthesis of tricarboxylic acid-cycle intermediates during acetate metabolism was studied in isolated rat hearts perfused with [1-14C]pyruvate. 2. The incorporation of the 14C label from [1-14C]pyruvate into the tricarboxylic acid-cycle intermediates points to a carbon input from pyruvate via enzymes in addition to pyruvate dehydrogenase and citrate synthase. 3. On addition of acetate, the specific radioactivity of citrate showed an initial maximum at 2 min, with a subsequent decline in labelling. The C-6 of citrate (which is removed in the isocitrate dehydrogenase reaction) and the remainder of the molecule showed differential labelling kinetics, the specific radioactivity of C-6 declining more rapidly. Since this carbon is lost in the isocitrate dehydrogenase reaction, the results are consistent with a rapid inactivation of pyruvate dehydrogenase after the addition of acetate, which was confirmed by measuring the 14CO2 production from [1-14C]pyruvate. 4. The results can be interpreted to show that carboxylation of pyruvate to the C4 compounds of the tricarboxylic acid cycle occurs under conditions necessitating anaplerosis in rat myocardium, although the results do not identify the enzyme involved. 5. The specific radioactivity of tissue lactate was too low to allow it to be used as an indicator of the specific radioactivity of the intracellular pyruvate pool. The specific radioactivity of alanine was three times that of lactate. When the hearts were perfused with [1-14C]lactate, the specific radioactivity of alanine was 70% of that of pyruvate. The results suggest that a subcompartmentation of lactate and pyruvate occurs in the cytosol.  相似文献   

16.
Ishikita H 《Biochemistry》2008,47(15):4394-4402
The redox potential of the flavin mononucleotide (FMN) hydroquinones for one-electron reduction in the Desulfovibrio vulgaris ( D. vulgaris) flavodoxin ( E sq/hq for FMNH (*)/FMNH (-)) was calculated using the crystal structure of the relevant hydroquinone form and compared to the results of the Clostridium beijerinckii ( C. beijerinckii) flavodoxin. In D. vulgaris and C. beijerinckii flavodoxins, the protein side chain causes significant downshifts of 170 and 240 mV in E sq/hq, respectively. In the C. beijerinckii flavodoxin, the E sq/hq downshift because of the protein side chain is essentially compensated by the counter influence of the protein backbone ( E sq/hq upshift of 260 mV). However, in the D. vulgaris flavodoxin, the corresponding protein backbone influence on E sq/hq is significantly small, i.e., less than half of that in the C. beijerinckii flavodoxin. In particular, there is a significant difference in the influence of the protein backbone of the so-called 60s loop region between the two flavodoxins. The E sq/hq difference can be best explained by the lower compensation of the side chain influence by the backbone influence in the D. vulgaris flavodoxin than in the C. beijerinckii flavodoxin.  相似文献   

17.
The purified carbon monoxide dehydrogenase from Clostridium thermoaceticum is the only protein required to catalyze an exchange reaction between carbon monoxide and the carbonyl group of acetyl-CoA. This exchange requires that the CO dehydrogenase bind the methyl, the carbonyl, and the CoA groups of acetyl-CoA, then equilibrate the carbonyl with CO in the solution and re-form acetyl-CoA. CoA is not necessary for the exchange and, in fact, inhibits the reaction. These studies support the view that CO dehydrogenase is the condensing enzyme that forms acetyl-CoA from its component parts. Carbon dioxide also exchanges with the C-1 of acetyl-CoA, but at a much lower rate than does CO. At 50 degrees C and pH 5.3, the optimal pH, the turnover number is 70 mol of CO exchanged per min/mol of enzyme. Low potential electron carriers are stimulatory. The Km app for stimulation by ferredoxin is 50-fold less than the value for flavodoxin. Neither ATP or Pi stimulate the exchange. The EPR spectrum of the CO-reacted enzyme is markedly changed by binding of CoA or acetyl-CoA. Arginine residues of the CO dehydrogenase appear to be involved in the active site, possibly by binding acetyl-CoA. Mersalyl acid, methyl iodide, 5,5-dithiobis-(2-nitrobenzoate), and sodium dithionite inhibit the exchange reaction. A scheme is presented to account for the role of CO dehydrogenase in the exchange reaction and in the synthesis of acetate.  相似文献   

18.
Carbon monoxide dehydrogenase (CO dehydrogenase) has been purified from the homoacetate-fermenting bacterium, Clostridium thermoaceticum. By use of 63Ni, it has been determined that the dehydrogenase is a metallo nickel enzyme. Nickel was rapidly taken up by the organism and most of the ingested metal was found to be incorporated into CO dehydrogenase. As estimated by gel filtration, the native enzyme has a molecular weight of 410,000. Ferredoxin and a membrane-bound b-type cytochrome, both obtained from C. thermoaceticum, are rapidly reduced by the enzyme in the presence of carbon monoxide and both are considered to be native electron carriers. FMN and Desulfovibrio vulgaris cytochrome c3 were also reduced by the enzyme, while spinach ferredoxin, FAD, NAD, and NADP were not. CO dehydrogenase activity was not appreciably affected by propyl iodide, methyl iodide, carbon tetrachloride, or metal chelators, but was reversibly inhibited by KCN. A method for the in situ assay of CO dehydrogenase in polyacrylamide gels is presented.  相似文献   

19.
F Shimizu  M Ogata  T Yagi  S Wakabayashi  H Matsubara 《Biochimie》1989,71(11-12):1171-1177
Rubredoxin was purified from Desulfovibrio vulgaris Miyazaki. It was sequenced and some of its properties determined. Rubredoxin is composed of 52 amino acids. It is highly homologous to that from D. vulgaris Hildenborough. Its N-terminal methionyl residue is partially formylated. The millimolar absorption coefficients of the rubredoxin at 489 nm and 280 nm are 8.1 and 18.5, respectively, and the standard redox potential is +5 mV, which is slightly higher than those of other rubredoxins. Rubredoxin, as well as cytochrome c-553, was reduced with lactate by the action of lactate dehydrogenase of this organism, and the reaction was stimulated with 2-methyl-1,4-naphthoquinone. It is suggested that rubredoxin, in collaboration with membranous quinone, functions as a natural electron carrier for cytoplasmic lactate dehydrogenase of this organism, whereas cytochrome c-553 plays the same role for periplasmic lactate dehydrogenase.  相似文献   

20.
The chemical characterization, EPR properties, and mechanism of pyruvate:flavodoxin (ferredoxin) oxidoreductase from Klebsiella pneumoniae and Clostridium thermoaceticum have been investigated. A simple, specific, and sensitive assay and an efficient purification (based on the high affinity of these enzymes for a dye attached to agarose) are reported. The observed iron content of 8 atoms/subunit is twice that reported by others, whereas the contents of lipoate and flavin are less than 0.1 mol/subunit, in agreement with previous reports. Spectroscopic evidence suggests that the iron is present in Fe4S4(2+,1+) clusters. Reduction of the enzyme requires the presence of CoA as well as 1.1 pyruvate/subunit, which is very nearly the theoretical amount required the reduce two Fe4S(2+,1+) clusters. In the absence of CoA, stoichiometric amounts of pyruvate are decarboxylated, but the Fe/S centers are not reduced. We conclude that the K. pneumoniae and C. thermoaceticum enzymes are adapted to rapid reduction of low potential 1-e- carriers, similar to the pyruvate oxidoreductase of Halobacterium (Kerscher, L., and Oesterhelt, D. (1977) FEBS Lett. 83, 197-201), but different in that an Fe/S center-radical pair is used in the latter enzyme in place of the pair of Fe4S4 centers we find. The K. pneumoniae and C. thermoaceticum oxidoreductases appear to be mechanistically closely related to the Clostridium acidiurici enzyme (Uyeda, K., and Rabinowitz, J. C. (1971) J. Biol. Chem. 246, 3111-3119), differing as a class from the lipoate-containing, pyridine nucleotide-reducing enzyme present in aerobes (Reed, L. J. (1974) Accts. Chem. Res. 2, 740-746). The function of the Klebsiella enzyme is to supply electrons to nitrogenase. This is accomplished in vitro with purified components via a nif-specific flavodoxin or other low potential 1-e- carriers such as viologen dyes or ferredoxins. The in vivo molar ratio of nitrogenase to the physiological reduction system, estimated from activity measurements of individual components in crude extracts, was 0.4:0.03:2:1 pyruvate oxidoreductase:flavodoxin:nitrogenase component II:nitrogenase component I.  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