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 8 毫秒
1.
Extracts of Ruminococcus albus were not able to convert pyruvate to acetyl phosphate, CO2, and H2 after passage through a diethylaminoethyl (DEAE)-cellulose column. Activity was restored by a brown protein fraction eluted from the column with 0.4 M Cl-. The protein was partially purified and shown to have the spectral and biological characteristics of ferredoxin. R. albus ferredoxin, Clostridium pasteurianum ferredoxin, and methyl viologen restored activity for pyruvate decomposition by DEAE-cellulose-treated R. albus extracts. R. albus or C. pasteurianum ferredoxin restored the ability of DEAE-cellulose-treated C. pasteurianum extracts to form H2 and acetyl phosphate from pyruvate. Ferredoxin-free extracts of R. albus reduced nicotinamide adenine dinucleotide (NAD) when supplemented with R. albus or C. pasteurianum ferredoxin or with methyl viologen. These extracts reduced NADP with H2 poorly unless both ferredoxin and NAD were added, which indicates the presence of an NADH:NADP transhydrogenase. Flavin mononucleotide and flavin adenine dinucleotide were rapidly reduced by H2 by ferredoxin-free extracts in the absence of ferredoxin.  相似文献   

2.
The observation that adenosine triphosphate (ATP)-dependent CO(2) fixation in extracts of chemosynthetic and photosynthetic autotrophs may be regulated in part by adenosine monophosphate (AMP) was extended to the strict autotroph Thiobacillus neapolitanus (X). In addition, this report presents data which include adenosine diphosphate (ADP) in the regulatory role. When the primary CO(2) acceptor, ribose-5-phosphate, was replaced by ribulose-1,5-diphosphate, no inhibition of CO(2) fixation occurred unless the Mg(++) concentration was limiting. A molar ratio of 5:1 AMP or ADP to ATP reduced the specific activity (micromoles of CO(2) fixed per milligram of protein per minute) of the extracts from 0.22 to 0.12 and 0.11, respectively. The reported stimulation of the carboxylative phase of ATP-dependent CO(2) fixation by reduced nicotinamide adenine dinucleotide (NADH(2)) was investigated. Adding NADH(2) to the extracts did not stimulate CO(2) fixation, even at carbonate levels from 0.05 to 30 mumoles, except in the absence of ribose-5-phosphate. Slight increases in CO(2) fixation were noted when the assay system was incubated in air instead of the usual helium atmosphere.  相似文献   

3.
After growth on various carbon sources, sonic extracts of Hydrogenomonas facilis contained ribulosediphosphate (RuDP) carboxylase and phosphoribulokinase (Ru5-P kinase). After very short sonic treatment, a reductive adenosine triphosphate (ATP)-dependent incorporation of (14)CO(2) was also detectable. Reduced nicotinamide adenine dinucleotide (NADH(2)) served as reductant 30-fold more effectively than reduced nicotinamide adenine dinucleotide phosphate (NADPH(2)). Adenosine 5'-phosphate (AMP) and adenosine 5'-pyrophosphate (ADP) inhibited Ru5-P kinase and NADH(2)-, ATP-dependent CO(2) fixation. The levels and duration of CO(2) fixation suggested that it is a cyclic process. The requirement of reduced pyridine nucleotide and ATP and the sensitivity of fixation to AMP and ADP support the conjecture that it occurs via the Calvin cycle. After thorough study of variables affecting catalysis, specific activities (millimicromoles of substrate disappearing per milligram of protein) at 30 C were determined for RuDP carboxylase (C), Ru5-P kinase (K) and ATP-, NADH(2)- dependent CO(2) fixation (CO(2) F) after growth autotrophically on fructose, glucose, ribose, glutamate, lactate, succinate, and acetate. Values for these growth modes were, respectively-for C: 67.3, 51.1, 51.4, 24.6, 2.05, 10.2, 2.25, 1.4; for K: 24.7, 24.0, 23.2, 14.2, 12.8, 12.9, 13.4, 2.8; and for CO(2) F: 4.54, 4.83, 3.10, 2.87, 0.85, 1.51, 0.24, 0.41. The qualitative parallel between values for RuDP carboxylase and CO(2) fixation suggests that one major control point in fixation is the step catalyzed by RuDP carboxylase.  相似文献   

4.
1. Growing cultures of Peptostreptococcus elsdenii and Bacteroides ruminicola incorporate (14)C from [1-(14)C]isobutyrate into the valine of cell protein. With P. elsdenii some of the (14)C is also incorporated into leucine. 2. Crude cell-free extracts of both organisms in the presence of glutamine, carbon dioxide and suitable sources of energy and electrons incorporate (14)C from [1-(14)C]isobutyrate into valine but not into leucine. 3. With extracts of P. elsdenii treated with DEAE-cellulose the reaction is dependent on ATP, CoA, thiamin pyrophosphate, molecular hydrogen and a low-potential electron carrier (ferredoxin, flavodoxin or benzyl viologen). 4. The same extracts incorporate (14)C from NaH(14)CO(3) into valine in the presence of isobutyrate plus ATP, CoA, glutamine and ferredoxin; isobutyryl-CoA or isobutyryl phosphate plus CoA will replace the isobutyrate plus CoA and ATP. With acetyl phosphate in place of isobutyryl phosphate, (14)C is incorporated into alanine. With isovalerate or 2-methylbutyrate in place of isobutyrate, (14)C is incorporated into leucine and isoleucine respectively. 5. When carrier 2-oxoisovalerate is added to the carboxylating system (14)C from [1-(14)C]isobutyrate passes into the oxo acid fraction. 6. It is concluded that these two organisms form valine from isobutyrate by the sequence isobutyrate-->isobutyryl-CoA-->2-oxoisovalerate-->valine and that the reductive carboxylation of isobutyrate is catalysed by a system similar to the pyruvate synthetase of clostridia and photosynthetic bacteria.  相似文献   

5.
Metabolism of lactate as a carbon source by Pseudomonas citronellolis occurred via a nicotinamide adenine dinucleotide (NAD)-independent L-lactate dehydrogenase, which was present in cells grown on DL-lactate but was not present in cells grown on acetate, aspartate, citrate, glucose, glutamate, or malate. The cells also possessed a constitutive, NAD-independent malate dehydrogenase instead of the conventional NAD-dependent malate dehydrogenase instead of the conventional NAD-dependent enzyme in the tricarboxylic acid cycle. Both enzymes were particulate and used dichlorophenolindo-phenol or oxygen as an electron acceptor. In acetate-grown cells, the activity of pyruvate dehydrogenase and NAD phosphate-linked malate enzyme decreased, cells grown on glucose or lactate. This was consistent with the need to maintain a supply of oxalacetate for metabolism of acetate via the tricarboxylic acid cycle. Changes in enzyme activities suggest that gluconeogenesis from noncarbohydrate carbon sources occurs via the malate enzyme (when oxalacetate decarboxylase is inhibited) or a combination of the NAD-independent malate dehydrogenase and oxalacetate decarboxylase.  相似文献   

6.
Methanobacterium ruminantium was shown to possess a nicotinamide adenine dinucleotide phosphate (NADP)-linked factor 420 (F420)-dependent hydrogenase system. This system was also shown to be present in Methanobacterium strain MOH. The hydrogenase system of M. ruminantium also links directly to F420, flavin adenine dinucleotide (FAD), flavin mononucleotide (FMN), methyl viologen, and Fe-3 plus. It has a pH optimum of about 8 and an apparent Km for F420 of about 5 x 10-6 M at pH 8 when NADP is the electron acceptor. The F420-NADP oxidoreductase activity is inactive toward nicotinamide adenine dinucleotide (nad) and no NADPH:NAD or FADH2(FMNH2):NAD transhydrogenase system was detected. Neither crude ferredoxin nor boiled crude extract of Clostridium pasteuranum could replace F420 in the NADP-linked hydrogenase reaction of M. ruminantium. Also, neitther F420 nor a curde "ferredoxin" fraction from M. ruminantium extracts could substitute for ferredoxin in the pyruvate-ferredoxin oxidoreductase reaction of C. pasteurianum.  相似文献   

7.
Bacterial Metabolism of Mevalonic Acid   总被引:5,自引:4,他引:1       下载免费PDF全文
Soluble cell-free extracts of actinomycete S4 grown on media containing mevalonate catalyze acetoacetate formation from mevalonate, mevaldate, and β-hydroxy-β-methylglutaryl-coenzyme A (CoA). Conversion of mevalonate to acetoacetate involves formation of free β-hydroxy-β-methylglutaryl-CoA, but not free mevaldate. The reaction favors mevalonate oxidation, and nicotinamide adenine dinucleotide, rather than nicotinamide adenine dinucleotide phosphate, acts as oxidant.  相似文献   

8.
The two species of 6-phosphogluconate dehydrogenase (EC 1.1.1.43) from Pseudomonas multivorans were resolved from extracts of gluconate-grown bacteria and purified to homogeneity. Each enzyme comprised between 0.1 and 0.2% of the total cellular protein. Separation of the two enzymes, one which is specific for nicotinamide adenine dinucleotide phosphate and the other which is active with nicotinamide adenine dinucleotide or nicotinamide adenine dinucleotide phosphate was facilitated by the marked difference in their respective isoelectric points, which were at pH 5.0 and 6.9. Comparison of the subunit compositions of the two enzymes indicated that they do not share common peptide chains. The enzyme active with nicotinamide adenine dinucleotide was composed of two subunits of about 40,000 molecular weight, and the nicotinamide adenine dinucleotide phosphate-specific enzyme was composed of two subunits of about 60,000 molecular weight. Immunological studies indicated that the two enzymes do not share common antigenic determinants. Reduced nicotinamide adenine dinucleotide phosphate strongly inhibited the 6-phosphogluconate dehydrogenase active with nicotinamide adenine dinucleotide by decreasing its affinity for 6-phosphogluconate. Guanosine-5'-triphosphate had a similar influence on the nicotinamide adenine dinucleotide phosphate-specific 6-phosphogluconate dehydrogenase. These results in conjunction with other data indicating that reduced nicotinamide adenine dinucleotide phosphate stimulates the conversion of 6-phosphogluconate to pyruvate by crude bacterial extracts suggest that in P. multivorans, the relative distribution of 6-phosphogluconate into the pentose phosphate and Entner-Doudoroff pathways might be determined by the intracellular concentrations of reduced nicotinamide adenine dinucleotide phosphate and purine nucleotides.  相似文献   

9.
Dialyzed extracts of Acetobacter suboxydans ATCC 621 catalyze (14)CO(2) assimilation in the presence of phosphoenolpyruvate and a divalent cation. The formation of (14)C-oxalacetate was demonstrated and found not to be dependent upon the presence of orthophosphate or diphosphonucleotides. Oxalacetate synthesis was stimulated by orthophosphate and inhibited by aspartate. All attempts to demonstrate a reversible carboxylation mechanism have failed. (14)C-aspartate was synthesized when phosphoenolpyruvate, H(14)Co(3) (-), pyridoxal phosphate, and glutamate were added to dialyzed extracts. Chromatographic and spectrophotometric analyses and chemical degradation further demonstrate the presence of a reversible aspartate aminotransferase. The function of oxalacetate synthesis in a bacterium that reportedly lacks an operative tricarboxylic acid cycle is discussed.  相似文献   

10.
Cell-free extracts of Bacillus licheniformis were found to contain pyruvate carboxylase which catalyzes the reaction between pyruvate and bicarbonate to yield oxalacetate in the presence of adenosine triphosphate (ATP), acetylcoenzyme A (CoA), and manganese. The plot between the reaction velocity of the carboxylation by the partially purified pyruvate carboxylase (25-fold) and the concentration of pyruvate, bicarbonate, manganese, and ATP did not indicate a pronounced deviation from the Michaelis-Menten hyperbola. The enzyme was inhibited by avidin and aspartate. Biotin partially protected the enzyme from avidin inhibition, whereas the amount of inhibition by aspartate was dependent on the concentration of acetyl-CoA present. The intracellular concentration of acetyl-CoA did not vary significantly enough to allow control of the enzyme by this method. Extracts of 4-hr postexponential-phase cells of B. licheniformis were also found to contain phosphoenolpyruvate carboxykinase, which appears to be under catabolite repression control. It is suggested that the endogenous induction of this enzyme is the determining factor allowing the shift to gluconeogenesis from glycolysis during sporulation of glucose-grown cells.  相似文献   

11.
CO2 fixation by particle-free extracts from Nitrobacter winogradskyi increased by addition of reduced nicotinamide adenine dinucleotide (NADH). Ribulose-1,5-diphosphate, however, increased CO2 fixation, even in the absence of NADH. Phosphoribulokinase (EC 2.7.1.19) was the enzyme of Nitrobacter extracts that was activated specifically by NADH. Pyridoxal-5-phosphate inhibited both CO2 fixation and NADH-activated phosphoribulokinase from Nitrobacter. However, it did not affect phosphoribulokinase from spinach leaves. Since the spinach enzyme had also no requirement for reduced pyridine nucleotides, it appears that pyridoxal phosphate interferes only with the binding of NADH and not with the binding of ribulose-5-phosphate and adenosine-5'-triphosphate. The regulation of phosphoribulokinase activity by NADH provided Nitrobacter with an energy-dependent control mechanism of CO2 assimilation.  相似文献   

12.
Heterotrophic Carbon Metabolism by Beggiatoa alba   总被引:6,自引:0,他引:6       下载免费PDF全文
The assimilation and metabolism of CO(2) and acetate by Beggiatoa alba strain B18LD was investigated. Although B. alba was shown to require CO(2) for growth, the addition of excess CO(2) (as NaHCO(3)) to the medium in a closed system did not stimulate growth. Approximately 24 to 31% of the methyl-labeled acetate and 38 to 46% of the carboxyl-labeled acetate were oxidized to (14)CO(2) by B. alba. The apparent V(max) values for combined assimilation and oxidation of [2-(14)C]acetate by B. alba were 126 to 202 nmol min(-1) mg of protein(-1) under differing growth conditions. The V(max) values for CO(2) assimilation by heterotrophic and mixotrophic cells were 106 and 131 pmol min(-1) mg of protein(-1), respectively. The low V(max) values for CO(2) assimilation, coupled with the high V(max) values for acetate oxidation, suggested that the required CO(2) was endogenously produced from acetate. Moreover, exogenously supplied acetate was required by B. alba for the fixation of CO(2). From 61 to 73% of the [(14)C]acetate assimilated by washed trichomes was incorporated into lipid. Fifty-five percent of the assimilated [2-(14)C]acetate was incorporated into poly-beta-hydroxybutyric acid. This was consistent with chemical data showing that 56% of the heterotrophic cell dry weight was poly-beta-hydroxybutyric acid. Succinate and CO(2) were incorporated into cell wall material, proteins, lipids, nucleic acids, and amino and organic acids, but not into poly-beta-hydroxybutyric acid. Glutamate and succinate were the major stable products after short-term [1-(14)C]acetate assimilation. Glutamate and aspartate were the first stable (14)CO(2) fixation products, whereas glutamate, a phosphorylated compound, succinate, and aspartate were the major stable (14)CO(2) fixation products over a 30-min period. The CO(2) fixation enzymes isocitrate dehydrogenase (nicotinamide adenine dinucleotide phosphate; reversed) and malate dehydrogenase (nicotinamide adenine dinucleotide phosphate; decarboxylating) were found in cell-free extracts of both mixotrophically grown and heterotrophically grown cells. The data indicate that the typical autotrophic CO(2) fixation mechanisms are absent from B. alba B18LD and that the CO(2) and acetate metabolism pathways are probably linked.  相似文献   

13.
A method has been developed for rapidly preparing bundle sheath cell strands from Urochloa panicoides, a phosphoenolpyruvate (PEP) carboxykinase-type C4 plant. These cells catalyzed both HCO3(-)- and oxaloacetate-dependent oxygen evolution; oxaloacetate-dependent oxygen evolution was stimulated by ATP. For this activity oxaloacetate could be replaced by aspartate plus 2-oxoglutarate. Both oxaloacetate- and aspartate plus 2-oxoglutarate-dependent oxygen evolution were accompanied by PEP production and both were inhibited by 3-mercaptopicolinic acid, an inhibitor of PEP carboxykinase. The ATP requirement for oxaloacetate- and aspartate plus 2-oxoglutarate-dependent oxygen evolution could be replaced by ADP plus malate. The increased oxygen evolution observed when malate plus ADP was added with oxaloacetate was accompanied by pyruvate production. These results are consistent with oxaloacetate being decarboxylated via PEP carboxykinase. We suggest that the ATP required for oxaloacetate decarboxylation via PEP carboxykinase may be derived by phosphorylation coupled to malate oxidation in mitochondria. These bundle sheath cells apparently contain diffusion paths for the rapid transfer of compounds as large as adenine nucleotides.  相似文献   

14.
Pyruvate was shown to be the precursor of diacetyl and acetoin in Streptococcus diacetilactis, but dialyzed cell-free extracts of S. diacetilactis and Leuconostoc citrovorum that had been treated with anion-exchange resin to remove coenzyme A (CoA) formed only acetoin from pyruvate in the presence of thiamine pyrophosphate (TPP) and Mg(++) or Mn(++) ions. The ability to produce diacetyl was restored by the addition of acetyl-CoA. Acetyl-phosphate did not replace the acetyl-CoA. Neither diacetyl nor acetoin was formed when the otherwise complete reaction system was modified by using boiled extract or by omitting the extract, pyruvate, TPP, or the metal ions. Free acetaldehyde was not involved in the biosynthesis of diacetyl or acetoin from pyruvate, dialyzed cell-free extracts of the bacteria produced only acetoin (besides CO(2)) from alpha-acetolactate, and acetoin was not involved in the biosynthesis of diacetyl. Only one of the optical isomers present in racemic alpha-acetolactate was attacked by the extracts, and there was no appreciable spontaneous decarboxylation of the alpha-acetolactate at the pH (4.5) used in experiments.  相似文献   

15.
Cell extracts from fermentatively grown Rhodospirillum rubrum reduced about 80 nmol of nicotinamide adenine dinucleotide (NAD) per mg of protein per min under anaerobic conditions with sodium pyruvate. The reaction was specific for pyruvate and NAD; NAD phosphate was not reduced. Results indicated that pyruvate-linked NAD reduction occurred via pyruvate:lipoate oxidoreductase. The reaction required catalytic amounts of both coenzyme A and thiamine pyrophosphate. Addition of sodium arsenite inhibited enzyme activity by 90%. Pyruvate:lipoate oxidoreductase was the only system detected in anaerobic, dark-grown R. rubrum cell extracts which operated to produce reduced NAD. The low activity of the enzyme system suggested that it was not quantitatively important in ATP formation.  相似文献   

16.
Nicotinamide adenine dinucleotide phosphate-specific isocitrate dehydrogenase was extracted from etiolated pea (Pisum sativum L.) seedlings and was purified 65-fold. The purified enzyme exhibits one predominant protein band by polyacrylamide gel electrophoresis, which corresponds to the dehydrogenase activity as measured by the nitro blue tetrazolium technique. The reaction is readily reversible, the pH optima for the forward (nicotinamide adenine dinucleotide phosphate reduction) and reverse reactions being 8.4 and 6.0, respectively. The enzyme has different cofactor and inhibitor characteristics in the two directions. Manganese ions can be used as a cofactor for the reaction in each direction but magnesium ions only act as a cofactor in the forward reaction. Zinc ions, and to a lesser extent calcium ions, inhibit the enzyme at low concentrations when magnesium but not manganese is the metal activator. It is suggested that there is a fundamental difference between magnesium and manganese in the activation of the enzyme. The enzyme shows normal kinetics and the Michaelis contant for each substrate was determined. The inhibition by nucleotides, nucleosides, reaction products, and related compounds was studied. The enzyme shows a linear response to the mole fraction of reduced nicotinamide adenine dinucleotide phosphate when total nicotinamide adenine dinucleotide phosphate (nicotinamide adenine dinucleotide phosphate plus reduced nicotinamide adenine dinucleotide phosphate) is kept constant. Isocitrate in the presence of divalent metal ions will protect the enzyme from inactivation by p-chloromercuribenzoate. Protection is also afforded by manganese ions alone but not by magnesium ions alone There is a concerted inhibition of the enzyme by oxalacetate and glyoxylate.  相似文献   

17.
Acetate assimilation pathway of Methanosarcina barkeri.   总被引:18,自引:11,他引:7       下载免费PDF全文
The pathway of acetate assimilation in Methanosarcina barkeri was determined from analysis of the position of label in alanine, aspartate, and glutamate formed in cells grown in the presence of [14C]acetate and by measurement of enzyme activities in cell extracts. The specific radioactivity of glutamate from cells grown on [1-14C]- or [2-14C]acetate was approximately twice that of aspartate. The methyl and carboxyl carbons of acetate were incorporated into aspartate and glutamate to similar extents. Degradation studies revealed that acetate was not significantly incorporated into the C1 of alanine, C1 or C4 of aspartate, or C1 of glutamate. The C5 of glutamate, however, was partially derived from the carboxyl carbon of acetate. Cell extracts were found to contain the following enzyme activities, in nanomoles per minute per milligram of protein at 37 degrees C: F420-linked pyruvate synthase, 170; citrate synthase, 0.7; aconitase, 55; oxidized nicotinamide adenine dinucleotide phosphate-linked isocitrate dehydrogenase, 75; and oxidized nicotinamide adenine dinucleotide-linked malate dehydrogenase, 76. The results indicate that M. barkeri assimilates acetate into alanine and aspartate via pyruvate and oxaloacetate and into glutamate via citrate, isocitrate, and alpha-ketoglutarate. The data reveal differences in the metabolism of M. barkeri and Methanobacterium thermoautotrophicum and similarities in the assimilation of acetate between M. barkeri and other anaerobic bacteria, such as Clostridium kluyveri.  相似文献   

18.
Glutamate dehydrogenase from Mycoplasma laidlawii   总被引:2,自引:2,他引:0       下载免费PDF全文
Mycoplasma laidlawii possesses a single glutamate dehydrogenase (GDH) with dual coenzyme specificity [specificity for nicotinamide adenine dinucleotide (H) and nicotinamide adenine dinucleotide phosphate (H)]. A purification procedure is reported which results in an enzyme preparation with a specific activity of 79.5 units/mg and which displays only one significant protein band after gel electrophoresis. This one band was determined, by activity staining, to have all of the GDH nucleotide specificities. The molecular weight of the enzyme is 250,000 +/- 10%, and it has a subunit size of about 48,000. The enzyme exhibits measurable activity with aspartate and pyruvate but is inactive with eight other possible substrates. Purine nucleotides do not affect the activity. The K(m) for reduced nicotinamide adenine dinucleotide was 1.8 x 10(-4)m. The optimal substrate concentrations and pH optimum for each of the respective GDH activities are also reported.  相似文献   

19.
The anaerobic oxidation of protoporphyrinogen to protoporphyrin was demonstrated in extracts of Desulfovibrio gigas. Protoporphyrin formation occurred in the presence of nitrite, hydroxylamine, sulfite, thiosulfate, ATP plus sulfate, NAD+, NADP+, flavin adenine dinucleotide, flavin mononucleotide, fumarate, 2,6-dichlorophenol-indophenol, methyl viologen, and 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide. With dialyzed cell extracts, highest activities were observed with sulfite, NAD+, and NADP+ as electron acceptors. The enzyme for protoporphyrinogen oxidation was localized in the membrane of D. gigas and displayed optimal activity at pH 7.3 and 28 degrees C.  相似文献   

20.
Pyruvate-dependent reduction of NADP has been demonstrated in cell extracts of the human gastric pathogen Helicobacter pylori. However, NADP is not a substrate of purified pyruvate:ferredoxin oxidoreductase (PFOR), suggesting that other redox active enzymes mediate this reaction. Here we show that fqrB (HP1164), which is essential and highly conserved among the epsilonproteobacteria, exhibits NADPH oxidoreductase activity. FqrB was purified by nickel interaction chromatography following overexpression in Escherichia coli. The protein contained flavin adenine dinucleotide and exhibited NADPH quinone reductase activity with menadione or benzoquinone and weak activity with cytochrome c, molecular oxygen, and 5,5'-dithio-bis-2-nitrobenzoic acid (DTNB). FqrB exhibited a ping-pong catalytic mechanism, a k(cat) of 122 s(-1), and an apparent K(m) of 14 muM for menadione and 26 muM for NADPH. FqrB also reduced flavodoxin (FldA), the electron carrier of PFOR. In coupled enzyme assays with purified PFOR and FldA, FqrB reduced NADP in a pyruvate- and reduced coenzyme A (CoA)-dependent manner. Moreover, in the presence of NADPH, CO(2), and acetyl-CoA, the PFOR:FldA:FqrB complex generated pyruvate via CO(2) fixation. PFOR was the rate-limiting enzyme in the complex, and nitazoxanide, a specific inhibitor of PFOR of H. pylori and Campylobacter jejuni, also inhibited NADP reduction in cell-free lysates. These capnophilic (CO(2)-requiring) organisms contain gaps in pathways of central metabolism that would benefit substantially from pyruvate formation via CO(2) fixation. Thus, FqrB provides a novel function in pyruvate metabolism and, together with production of superoxide anions via quinone reduction under high oxygen tensions, contributes to the unique microaerobic lifestyle that defines the epsilonproteobacterial group.  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