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.
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   

2.
Pyruvate:ferredoxin oxidoreductase (PFOR) catalyzes the oxidative decarboxylation of pyruvate to acetyl-CoA and CO(2). The catalytic proficiency of this enzyme for the reverse reaction, pyruvate synthase, is poorly understood. Conversion of acetyl-CoA to pyruvate links the Wood-Ljungdahl pathway of autotrophic CO(2) fixation to the reductive tricarboxylic acid cycle, which in these autotrophic anaerobes is the stage for biosynthesis of all cellular macromolecules. The results described here demonstrate that the Clostridium thermoaceticum PFOR is a highly efficient pyruvate synthase. The Michaelis-Menten parameters for pyruvate synthesis by PFOR are: V(max) = 1.6 unit/mg (k(cat) = 3.2 s(-1)), K(m)(Acetyl-CoA) = 9 micrometer, and K(m)(CO(2)) = 2 mm. The intracellular concentrations of acetyl-CoA, CoASH, and pyruvate have been measured. The predicted rate of pyruvate synthesis at physiological concentrations of substrates clearly is sufficient to support the role of PFOR as a pyruvate synthase in vivo. Measurements of its k(cat)/K(m) values demonstrate that ferredoxin is a highly efficient electron carrier in both the oxidative and reductive reactions. On the other hand, rubredoxin is a poor substitute in the oxidative direction and is inept in donating electrons for pyruvate synthesis.  相似文献   

3.
Furdui C  Ragsdale SW 《Biochemistry》2002,41(31):9921-9937
Pyruvate:ferredoxin oxidoreductase (PFOR) catalyzes the coenzyme A (CoA)-dependent oxidative decarboxylation of pyruvate. In many autotrophic anaerobes, PFOR links the Wood-Ljungdahl pathway to glycolysis and to cell carbon synthesis. Herein, we cloned and sequenced the M. thermoacetica PFOR, demonstrating strong structural homology with the structurally characterized D. africanus PFOR, including the presence of three [4Fe-4S] clusters per monomeric unit. The PFOR reaction includes a hydroxyethyl-thiamin pyrophosphate (HE-TPP) radical intermediate, which forms rapidly after PFOR reacts with pyruvate. This step precedes electron transfer from the HE-TPP radical intermediate to an intramolecular [4Fe-4S] cluster. We show that CoA increases the rate of this redox reaction by 10(5)-fold. Analysis by Marcus theory indicates that, in the absence of CoA, this is a true electron-transfer reaction; however, in its presence, electron transfer is gated by an adiabatic event. Analysis by the Eyring equation indicates that entropic effects dominate this rate enhancement. Our results indicate that the energy of binding CoA contributes minimally to the rate increase since the thiol group of CoA lends over 40 kJ/mol to the reaction, whereas components of CoA that afford most of the cofactor's binding energy contribute minimally. Major conformational changes also do not appear to explain the rate enhancement. We propose several ways that CoA can accomplish this rate increase, including formation of a highly reducing adduct with the HE-TPP radical to increase the driving force for electron transfer. We also consider the possibility that CoA itself forms part of the electron-transfer pathway.  相似文献   

4.
Rubredoxin (Rd) from the moderately thermophilic green sulfur bacterium Chlorobium tepidum was found to function as an electron acceptor for pyruvate ferredoxin oxidoreductase (PFOR). This enzyme, which catalyzes the conversion of pyruvate to acetyl-CoA and CO(2), exhibited an absolute dependence upon the presence of Rd. However, Rd was incapable of participating in the pyruvate synthase or CO(2) fixation reaction of C. tepidum PFOR, for which two different reduced ferredoxins are employed as electron donors. These results suggest a specific functional role for Rd in pyruvate oxidation and provide the initial indication that the two important physiological reactions catalyzed by PFOR/pyruvate synthase are dependent on different electron carriers in the cell. The UV-visible spectrum of oxidized Rd, with a monomer molecular weight of 6500, gave a molar absorption coefficient at 492 nm of 6.89 mM(-1) cm(-1) with an A(492)/A(280) ratio of 0.343 and contained one iron atom/molecule. Further spectroscopic studies indicated that the CD spectrum of oxidized C. tepidum Rd exhibited a unique absorption maximum at 385 nm and a shoulder at 420 nm. The EPR spectrum of oxidized Rd also exhibited unusual anisotropic resonances at g = 9.675 and g = 4.322, which is composed of a narrow central feature with broader shoulders to high and low field. The midpoint reduction potential of C. tepidum Rd was determined to be -87 mV, which is the most electronegative value reported for Rd from any source.  相似文献   

5.
Oxidative decarboxylation of pyruvate forming acetyl-coenzyme A is a crucial step in many metabolic pathways. In most anaerobes, this reaction is carried out by pyruvate-ferredoxin oxidoreductase (PFOR), an enzyme normally oxygen sensitive except in Desulfovibrio africanus (Da), where it shows an abnormally high oxygen stability. Using site-directed mutagenesis, we have specified a disulfide bond-dependent protective mechanism against oxidative conditions in Da PFOR. Our data demonstrated that the two cysteine residues forming the only disulfide bond in the as-isolated PFOR are crucial for the stability of the enzyme in oxidative conditions. A methionine residue located in the environment of the proximal [4Fe-4S] cluster was also found to be essential for this protective mechanism. In vivo analysis demonstrated unambiguously that PFOR in Da cells as well as two other Desulfovibrio species was efficiently protected against oxidative stress. Importantly, a less active but stable Da PFOR in oxidized cells rapidly reactivated when returned to anaerobic medium. Our work demonstrates the existence of an elegant disulfide bond-dependent reversible mechanism, found in the Desulfovibrio species to protect one of the key enzymes implicated in the central metabolism of these strict anaerobes. This new mechanism could be considered as an adaptation strategy used by sulfate-reducing bacteria to cope with temporary oxidative conditions and to maintain an active dormancy.  相似文献   

6.
Pyruvate-ferredoxin oxidoreductases (PFOR) are unique among thiamine pyrophosphate (ThDP)-containing enzymes in giving rise to a rather stable cofactor-based free-radical species upon the decarboxylation of their first substrate, pyruvate. We have obtained snapshots of unreacted and partially reacted (probably as a tetrahedral intermediate) pyruvate-PFOR complexes at different time intervals. We conclude that pyruvate decarboxylation involves very limited substrate-to-product movements but a significant displacement of the thiazolium moiety of ThDP. In this respect, PFOR seems to differ substantially from other ThDP-containing enzymes, such as transketolase and pyruvate decarboxylase. In addition, exposure of PFOR to oxygen in the presence of pyruvate results in significant inhibition of catalytic activity, both in solution and in the crystals. Examination of the crystal structure of inhibited PFOR suggests that the loss of activity results from oxime formation at the 4' amino substituent of the pyrimidine moiety of ThDP.  相似文献   

7.
The nifJ gene codes for pyruvate:ferredoxin oxidoreductase (PFOR), which reduces ferredoxin during fermentative catabolism of pyruvate to acetyl-coenzyme A (acetyl-CoA). A nifJ knockout mutant was constructed that lacks one of two pathways for the oxidation of pyruvate in the cyanobacterium Synechococcus sp. strain PCC 7002. Remarkably, the photoautotrophic growth rate of this mutant increased by 20% relative to the wild-type (WT) rate under conditions of light-dark cycling. This result is attributed to an increase in the quantum yield of photosystem II (PSII) charge separation as measured by photosynthetic electron turnover efficiency determined using fast-repetition-rate fluorometry (F(v)/F(m)). During autofermentation, the excretion of acetate and lactate products by nifJ mutant cells decreased 2-fold and 1.2-fold, respectively. Although nifJ cells displayed higher in vitro hydrogenase activity than WT cells, H(2) production in vivo was 1.3-fold lower than the WT level. Inhibition of acetate-CoA ligase and pyruvate dehydrogenase complex by glycerol eliminated acetate production, with a resulting loss of reductant and a 3-fold decrease in H(2) production by nifJ cells compared to WT cells. Continuous electrochemical detection of dissolved H(2) revealed two temporally resolved phases of H(2) production during autofermentation, a minor first phase and a major second phase. The first phase was attributed to reduction of ferredoxin, because its level decreased 2-fold in nifJ cells. The second phase was attributed to glycolytic NADH production and decreased 20% in nifJ cells. Measurement of the intracellular NADH/NAD(+) ratio revealed that the reductant generated by PFOR contributing to the first phase of H(2) production was not in equilibrium with bulk NADH/NAD(+) and that the second phase corresponded to the equilibrium NADH-mediated process.  相似文献   

8.
Regulation of heart muscle pyruvate dehydrogenase kinase   总被引:31,自引:25,他引:6       下载免费PDF全文
1. The activity of pig heart pyruvate dehydrogenase kinase was assayed by the incorporation of [(32)P]phosphate from [gamma-(32)P]ATP into the dehydrogenase complex. There was a very close correlation between this incorporation and the loss of pyruvate dehydrogenase activity with all preparations studied. 2. Nucleoside triphosphates other than ATP (at 100mum) and cyclic 3':5'-nucleotides (at 10mum) had no significant effect on kinase activity. 3. The K(m) for thiamin pyrophosphate in the pyruvate dehydrogenase reaction was 0.76mum. Sodium pyrophosphate, adenylyl imidodiphosphate, ADP and GTP were competitive inhibitors against thiamin pyrophosphate in the dehydrogenase reaction. 4. The K(m) for ATP of the intrinsic kinase assayed in three preparations of pig heart pyruvate dehydrogenase was in the range 13.9-25.4mum. Inhibition by ADP and adenylyl imidodiphosphate was predominantly competitive, but there was nevertheless a definite non-competitive element. Thiamin pyrophosphate and sodium pyrophosphate were uncompetitive inhibitors against ATP. It is suggested that ADP and adenylyl imidodiphosphate inhibit the kinase mainly by binding to the ATP site and that the adenosine moiety may be involved in this binding. It is suggested that thiamin pyrophosphate, sodium pyrophosphate, adenylyl imidodiphosphate and ADP may inhibit the kinase by binding through pyrophosphate or imidodiphosphate moieties at some site other than the ATP site. It is not known whether this is the coenzyme-binding site in the pyruvate dehydrogenase reaction. 5. The K(m) for pyruvate in the pyruvate dehydrogenase reaction was 35.5mum. 2-Oxobutyrate and 3-hydroxypyruvate but not glyoxylate were also substrates; all three compounds inhibited pyruvate oxidation. 6. In preparations of pig heart pyruvate dehydrogenase free of thiamin pyrophosphate, pyruvate inhibited the kinase reaction at all concentrations in the range 25-500mum. The inhibition was uncompetitive. In the presence of thiamin pyrophosphate (endogenous or added at 2 or 10mum) the kinase activity was enhanced by low concentrations of pyruvate (25-100mum) and inhibited by a high concentration (500mum). Activation of the kinase reaction was not seen when sodium pyrophosphate was substituted for thiamin pyrophosphate. 7. Under the conditions of the kinase assay, pig heart pyruvate dehydrogenase forms (14)CO(2) from [1-(14)C]pyruvate in the presence of thiamin pyrophosphate. Previous work suggests that the products may include acetoin. Acetoin activated the kinase reaction in the presence of thiamin pyrophosphate but not with sodium pyrophosphate. It is suggested that acetoin formation may contribute to activation of the kinase reaction by low pyruvate concentrations in the presence of thiamin pyrophosphate. 8. Pyruvate effected the conversion of pyruvate dehydrogenase phosphate into pyruvate dehydrogenase in rat heart mitochondria incubated with 5mm-2-oxoglutarate and 0.5mm-l-malate as respiratory substrates. It is suggested that this effect of pyruvate is due to inhibition of the pyruvate dehydrogenase kinase reaction in the mitochondrion. 9. Pyruvate dehydrogenase kinase activity was inhibited by high concentrations of Mg(2+) (15mm) and by Ca(2+) (10nm-10mum) at low Mg(2+) (0.15mm) but not at high Mg(2+) (15mm).  相似文献   

9.
The radical intermediate of pyruvate:ferredoxin oxidoreductase (PFOR) from Moorella thermoacetica was characterized using electron paramagnetic resonance (EPR) spectroscopy at X-band and D-band microwave frequencies. EPR spectra, obtained with various combinations of isotopically labeled substrate (pyruvate) and coenzyme (thiamine pyrophosphate (TPP)), were analyzed by spectral simulations. Parameters obtained from the simulations were compared with those predicted from electronic structure calculations on various radical structures. The g-values and 14N/15N-hyperfine splittings obtained from the spectra are consistent with a planar, hydroxyethylidene-thiamine pyrophosphate (HE-TPP) pi-radical, in which spin is delocalized onto the thiazolium sulfur and nitrogen atoms. The 1H-hyperfine splittings from the methyl group of pyruvate and the 13C-hyperfine splittings from C2 of both pyruvate and TPP are consistent with a model in which the pyruvate-derived oxygen atom of the HE-TPP radical forms a hydrogen bond. The hyperfine splitting constants and g-values are not compatible with those predicted for a nonplanar, sigma/n-type cation radical.  相似文献   

10.
Sulfate reducers have developed a multifaceted adaptative strategy to survive against oxidative stresses. Along with this oxidative stress response, we recently characterized an elegant reversible disulfide bond-dependent protective mechanism in the pyruvate:ferredoxin oxidoreductase (PFOR) of various Desulfovibrio species. Here, we searched for thiol redox systems involved in this mechanism. Using thiol fluorescent labeling, we show that glutathione is not the major thiol/disulfide balance-controlling compound in four different Desulfovibrio species and that no other plentiful low molecular weight thiol can be detected. Enzymatic analyses of two thioredoxins (Trxs) and three thioredoxin reductases allow us to propose the existence of two independent Trx systems in Desulfovibrio vulgaris Hildenborough (DvH). The TR1/Trx1 system corresponds to the typical bacterial Trx system. We measured a TR1 apparent K(m) value for Trx1 of 8.9 μM. Moreover, our results showed that activity of TR1 was NADPH-dependent. The second system named TR3/Trx3 corresponds to an unconventional Trx system as TR3 used preferentially NADH (K(m) for NADPH, 743 μM; K(m) for NADH, 5.6 μM), and Trx3 was unable to reduce insulin. The K(m) value of TR3 for Trx3 was 1.12 μM. In vitro experiments demonstrated that the TR1/Trx1 system was the only one able to reactivate the oxygen-protected form of Desulfovibrio africanus PFOR. Moreover, ex vivo pulldown assays using the mutant Trx1(C33S) as bait allowed us to capture PFOR from the DvH extract. Altogether, these data demonstrate that PFOR is a new target for Trx1, which is probably involved in the protective switch mechanism of the enzyme.  相似文献   

11.
Inhibition of pyruvate dehydrogenase complex by moniliformin.   总被引:2,自引:0,他引:2       下载免费PDF全文
The mechanism for the inhibition of pyruvate dehydrogenase complex from bovine heart by moniliformin was investigated. Thiamin pyrophosphate proved to be necessary for the inhibitory action of moniliformin. The inhibition reaction was shown to be time-dependent and to follow first-order and saturation kinetics. Pyruvate protected the pyruvate dehydrogenase complex against moniliformin inactivation. Extensive dialysis of the moniliformin-inactivated complex only partially reversed inactivation. Moniliformin seems to act by inhibition of the pyruvate dehydrogenase component of the enzyme complex and not by acting on the dihydrolipoamide transacetylase or dehydrogenase components, as shown by monitoring the effect of moniliformin on each component individually. On the basis of these results, a suicide inactivator mechanism for moniliformin on pyruvate dehydrogenase is proposed.  相似文献   

12.
Nonenzymatic decarboxylation of pyruvate   总被引:2,自引:0,他引:2  
Triton X-100, retinol, retinoic acid, retinal, hexane, dithiothreitol, mercaptoethanol, and some other commercially available chemicals caused nonenzymatic decarboxylation of pyruvate and alpha-ketoglutarate. "Lipids" obtained from human or pigeon liver homogenates using isopropanol/hexane also had very high nonenzymatic decarboxylating activity on these two alpha-ketoacids; most of this activity could be traced to the hexane (Eastman) used in the extraction. Optimum pH of the reaction with dithiothreitol and mercaptoethanol was 7-8 and with the other chemicals around 10, but considerable activity was present at pH 7-8. Liver homogenates had a scavenger effect on the decarboxylating activity of Triton X-100 and of dithiothreitol. Dithiothreitol and mercaptoethanol at high concentrations (greater than 1 mM) also had a scavenger effect on the decarboxylating activity of the "lipids." Pretreatment of Triton X-100, dithiothreitol, retinol, and the "lipids" with catalase markedly decreased the decarboxylating activity, while treatment with boiled catalase failed to do so. The results suggest that these compounds contain oxidizing contaminants, perhaps peroxide derivatives. Powerful oxidizing impurities have been reported in Triton X-100 from various sources by Y. Ashani and G. N. Catravas (1980, Anal. Biochem 109, 55-62). Such peroxide derivatives may cause nonenzymatic decarboxylation of pyruvate and alpha-ketoglutarate, presumably by a mechanism similar to the well-known nonenzymatic decarboxylation of alpha-ketoacids by hydrogen peroxide. In the absence of catalase and/or other protective agents against reactive oxygen derivatives, these chemicals would interfere in the assays of pyruvate dehydrogenase, pyruvate dehydrogenase complex, and alpha-ketoglutarate dehydrogenase complex which depend on the release of 14CO2 from alpha[1-14C]ketoacids.  相似文献   

13.
We have applied a model that permits the estimation of the sensitivity of flux through branch point enzymes (D. C. LaPorte, K. Walsh, and D. E. Koshland, J. Biol. Chem. 259:14068-14075, 1984) in order to analyze the control of flux through the lactate-acetate branch point of Selenomonas ruminantium grown in glucose-limited continuous culture. At this branch point, pyruvate is the substrate of both the NAD-dependent L-(+)-lactate dehydrogenase (LDH) and the pyruvate:ferredoxin oxidoreductase (PFOR). The LDH was purified, and it exhibited positive cooperativity for the binding of pyruvate. The LDH had an [S].5 for pyruvate of 0.43 mM, a Hill coefficient of 2.4, and a K' equal to 0.13 mM. The PFOR, assayed in cell extracts, exhibited Michaelis-Menten kinetics for pyruvate, with a Km of 0.49 mM. Carbon flux through the LDH and the PFOR increased 80-fold and 3-fold, respectively, as the dilution rate was increased from 0.07 to 0.52 h-1 in glucose-limited continuous culture. There was nearly a twofold increase, from 6.5 to 11.2 mumol min-1 mg of protein-1 in the specific activity (i.e., maximum velocity) of the LDH at dilution rates of 0.11 and 0.52 h-1, respectively. A flux equation was used to calculate the intracellular concentration of pyruvate; a fourfold increase in pyruvate, from 0.023 to 0.093 mM, was thereby predicted as the dilution rate was increased from 0.07 to 0.52 h-1. When these calculated values of intracellular pyruvate concentration were inserted into the flux equation, the predicted values of flux through the LDH and the PFOR were found to match closely the flux actually measured in the chemostat-grown cells. Thus, the 80-fold increase in flux through the LDH was due to a twofold increase in the maximum velocity of the LDH and a fourfold increase in the intracellular pyruvate concentration. In addition, the flux through the LDH exhibited ultrasensitivity to changes in both the maximum velocity of the LDH and the intracellular concentration of pyruvate. The flux through the PFOR exhibited ultrasensitivity to changes in the maximum velocity of the LDH and hyperbolic sensitivity to changes in the intracellular concentration of pyruvate.  相似文献   

14.
15.
It is shown that the relative amount of the holoenzyme in the highly purified pyruvate dehydrogenase complex from the bovine brain is higher when the enzyme activity is assayed in the reaction of nonoxidative formation of acetaldehyde as compared to the pyruvate: NAD+ reductase reaction. The S0.5 values for thiamine pyrophosphate are as following: (TPP) (0.314 +/- 0.22) x 10(-7) M with reaction of nonoxidative formation of acetaldehyde, (0.188 +/- 0.08) x 10(-6) M and (1.65 +/- 1.16) x 10(-6) M in case of the pyruvate: NAD+ reductase reaction. TPP in the concentration of (0.5-6.0) x 10(-7) M completely protects the sites of nonoxidative formation of acetaldehyde from modification by the coenzyme analogs, 4'-oxythiamine pyrophosphate and tetrahydrothiamine pyrophosphate. However, the pyruvate: NAD+ reductase activity of the pyruvate dehydrogenase complex is inhibited in this case by 30-34%. The data obtained suggest that in contrast to the pyruvate: NAD+ reductase reaction the conversion of pyruvate to acetaldehyde occurs by the sites which tightly bound TPP.  相似文献   

16.
Pyruvate was produced from glucose by Escherichia coli BW25113 that contained formate dehydrogenase (FDH) from Mycobacterium vaccae. In aerobic shake-flask culture (K (L) a?=?4.9?min(-1)), the recombinant strain produced 6.7?g pyruvate?l(-1) after 24?h with 4?g sodium formate?l(-1) and a yield of 0.34?g pyruvate?g?glucose(-1). These values were higher than those of the original strain (0.2?g?l(-1) pyruvate and 0.02?g pyruvate?g?glucose(-1)). Based on the reaction mechanism of FDH, the introduction of FDH into E. coli enhances the accumulation of pyruvate by the regeneration of NADH from NAD(+) since NAD(+) is a shared cosubstrate with the pyruvate dehydrogenase complex, which decarboxylates pyruvate to acetyl-CoA and CO(2). The oxygenation level was enough high to inactivate lactate dehydrogenase, which was of benefit to pyruvate accumulation without lactate as a by-product.  相似文献   

17.
A comparative study of the pyruvate dehydrogenase complex and its pyruvate dehydrogenase component was carried out by using the circular dichroism method. It was found that the spectral properties of the pyruvate dehydrogenase complex are determined by those of its first component: i) the spectrum of the thiamine pyrophosphate-free pyruvate dehydrogenase complex displayed the main characteristics of the pyruvate dehydrogenase component; ii) the appearance of the charge transfer complex band during thiamine pyrophosphate saturation was revealed for the both proteins; iii) in both cases the charge transfer complex band disappeared after the interaction of the holoform with pyruvate and reappeared after the addition of dithiothreitol used as a deacetylating reagent. Coenzyme A in the same reaction selectively deacetylated the pyruvate dehydrogenase complex (but not its pyruvate dehydrogenase component). The spectral dynamics of pyruvate dehydrogenase reflects the functional changes in the enzyme active centers during the catalytic act. The similarity of the spectral behaviour of pyruvate dehydrogenase within the complex structure and in the isolated state provides support for the earlier proposed mechanism of the pyruvate dehydrogenase action and ensures a methodological basis for its direct investigation within the complex structure.  相似文献   

18.
A monospecific antibody recognizing two membrane proteins in Acholeplasma laidlawii identified a plasmid clone from a genomic library. The nucleotide sequence of the 4.6-kbp insert contained four sequential genes coding for proteins of 39 kDa (E1 alpha, N terminus not cloned), 36 kDa (E1 beta), 57 kDa (E2), and 36 kDa (E3; C terminus not cloned). The N termini of the cloned E2, E1 beta, and native A. laidlawii E2 proteins were verified by amino acid sequencing. Computer-aided searches showed that the translated DNA sequences were homologous to the four subenzymes of the pyruvate dehydrogenase complexes from gram-positive bacteria and humans. The plasmid-encoded 57-kDa (E2) protein was recognized by antibodies against the E2 subenzymes of the pyruvate and oxoglutarate dehydrogenase complexes from Bacillus subtilis. A substantial fraction of the E2 protein as well as part of the pyruvate dehydrogenase enzymatic activity was associated with the cytoplasmic membrane in A. laidlawii. In vivo complementation with three different Escherichia coli pyruvate dehydrogenase-defective mutants showed that the four plasmid-encoded proteins were able to restore pyruvate dehydrogenase enzyme activity in E. coli. Since A. laidlawii lacks oxoglutarate dehydrogenase and most likely branched-chain dehydrogenase enzyme complex activities, these results strongly suggest that the sequenced genes code for the pyruvate dehydrogenase complex.  相似文献   

19.
The enzyme-[14C] carboxybiotin complex of sheep liver pyruvate carboxylase was isolated and the reaction between this and pyruvate was studied by using the quenched-flow rapid-reaction technique. At 0.5 degrees C the reaction was 80% complete within 180 ms. The reaction was monophasic and obeyed pseudo-first-order kinetics. Increasing concentrations of Mg2+ caused a decrease in the magnitude of the observed pseudo-first-order rate constant. Throughout the carboxylation of pyruvate, the rate-limiting step of the reaction occurred after the dissociation of carboxybiotin from the first sub-site, whereas in the slow phase of the reaction with 2-oxobutyrate this dissociation is the rate-limiting step. It is possible, from the reaction scheme proposed, that the inhibition of overall enzymic activity by high concentrations of Mg2+ could be caused by the transfer of the carboxy group from biotin to pyruvate becoming rate-limiting. The efficacy of a substrate as a signal for the movement of carboxybiotin from the first sub-site is reflected by the amount that the effective affinity of the enzyme- carboxybiotin complex for Mg2+ is lowered. In the presence of the substrates tested, the affinities of the carboxybiotin complex can be arranged in order of increasing magnitude, i.e.: (formula; see text). The kinetics of the decay of the enzyme-[14C] carboxybiotin complex at 0 degree C in the absence of substrates are similar to the reaction with pyruvate except that the carboxybiotin is also unstable in the first sub-site, to some degree. This similarity allows for the proposal of a general scheme for the decarboxylation of the enzyme- carboxybiotin complex in the presence or in the absence of substrates.  相似文献   

20.
Cooperative interaction of pyruvate with the pyruvate dehydrogenase (PD) complex from pigeon breast muscle was shown. The sigmoidal dependence of the reaction rate on pyruvate concentration was observed for the PD complex. The Hill coefficient is equal to 1,5; no inhibition by the substrate (up to 2.2.10(-3) M) was found. The kinetic behaviour of the isolated pyruvate dehydrogenase component (PDH) analyzed under similar conditions, is more complex; this may be probably due to the presence of oligomeric forms with different molecular weights and specific activities. The competitive inhibitor of the PD complex--an amide of pyruvic acid (PA) (Ki=6.3-10(-6) M) activates the enzyme at low concentrations (less than 2,10(-6) M). When PA is present, the dependence of the reaction rate on pyruvate concentration gives a usual hyperbolic curve, v of [S]o. It is concluded that pyruvate may have a regulatory effect on the activity of muscle PD complex.  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