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.
We purified branched-chain keto acid dehydrogenase to a specific activity of 10 mumol/min per mg of protein from Pseudomonas putida grown on valine. The purified enzyme was active with 2-ketoisovalerate, 2-ketoisocaproate, and 2-keto-3-methylvalerate in a ratio of 1.0:0.8:0.7 but showed no activity with either pyruvate or 2-ketoglutarate. There were four polypeptides in the purified enzyme (molecular weights, 49,000, 46,000, 39,000, and 37,000). The purified enzyme was deficient in the specific lipoamide dehydrogenase produced during growth on valine (molecular weight, 49,000). Branched-chain keto acid dehydrogenase required L-valine, oxidized nicotinamide adenine dinucleotide, coenzyme A, thiamine pyrophosphate, and magnesium chloride. A partially purified preparation catalyzed the oxidation of 2-keto-[1-14C]isovalerate to [14C]carbon dioxide, isobutyryl-coenzyme A, and reduced nicotinamide adenine dinucleotide in equimolar amounts. Both the Km and the Vmax for 2-ketoisovalerate were affected by the addition of L-valine to the assay mixture. However, only the Vmax values for oxidized nicotinamide adenine dinucleotide and coenzyme A were affected when L-valine was present. This suggested that valine acted by affecting the binding of branched-chain keto acids to subunit E1 of the complex.  相似文献   

2.
A reliable and reproducible assay was developed for measuring mitochondrial α-keto acid decarboxylase activity using ferricyanide as the electron acceptor. This method permitted the functional isolation and investigation of the decarboxylase step of the branched-chain α-keto acid dehydrogenases in rat liver mitochondria. Pyruvate and α-ketoglutarate decarboxylases are known to be separate and distinct enzymes from the branched-chain α-keto acid decarboxylases and were studied as controls. The relative specific activities of rat liver mitochondrial decarboxylases as measured by the ferricyanide assay showed that pyruvate and α-ketoglutarate were decarboxylated twice as rapidly as α-ketoisovalerate and four to ten times as fast as α-keto-β-methylvalerate and α-ketoisocaproate. The three branched-chain α-keto acids individually inhibit pyruvate and α-ketoglutarate decarboxylases. Inactivation of mitochondrial branched-chain α-keto acid decarboxylase activity by freezing and thawing and by prolonged storage resulted in a proportional decrease in decarboxylase activity toward each of the three branched-chain α-keto acids. However, hypophysectomy was found to increase decarboxylase activity with α-keto-β-methylvalerate to four times normal and with α-ketoisovalerate to three times normal, but the activity with α-ketoisocaproate was not changed. Hypophysectomy did not alter mitochondrial decarboxylase activity with pyruvate, α-ketoglutarate, or α-ketovalerate. The finding that hypophysectomy differentially alters the mitochondrial decarboxylase activity with the three branched-chain α-keto acids suggests either that there is more than one substrate-specific enzyme with branched-chain α-keto acid decarboxylase activity or that there is a modification of one enzyme such that the catalytic activity is selectively altered toward the three substrates.  相似文献   

3.
2-Ketoisovalerate is used as a therapeutic agent, and a 2-ketoisovalerate-producing organism may serve as a platform for products deriving from this 2-keto acid. We engineered the wild type of Corynebacterium glutamicum for the growth-decoupled production of 2-ketoisovalerate from glucose by deletion of the aceE gene encoding the E1p subunit of the pyruvate dehydrogenase complex, deletion of the transaminase B gene ilvE, and additional overexpression of the ilvBNCD genes, encoding the l-valine biosynthetic enzymes acetohydroxyacid synthase (AHAS), acetohydroxyacid isomeroreductase, and dihydroxyacid dehydratase. 2-Ketoisovalerate production was further improved by deletion of the pyruvate:quinone oxidoreductase gene pqo. In fed-batch fermentations at high cell densities, the newly constructed strains produced up to 188 ± 28 mM (21.8 ± 3.2 g liter(-1)) 2-ketoisovalerate and showed a product yield of about 0.47 ± 0.05 mol per mol (0.3 ± 0.03 g per g) of glucose and a volumetric productivity of about 4.6 ± 0.6 mM (0.53 ± 0.07 g liter(-1)) 2-ketoisovalerate per h in the overall production phase. In studying the influence of the three branched-chain 2-keto acids 2-ketoisovalerate, 2-ketoisocaproate, and 2-keto-3-methylvalerate on the AHAS activity, we observed a competitive inhibition of the AHAS enzyme by 2-ketoisovalerate.  相似文献   

4.
The various neutral amino acids and aliphatic 2-keto acids exhibit differential effects on insulin secretion. The common denominator for all these effects is the 2-ketoglutarate generation in the pancreatic B-cell mitochondria. The neutral amino acidsl-leucine andl-norvaline and the aliphatic ketomonocarboxylic acids 2-ketoisocaproate, 2-ketocaproate, 2-ketovalerate, and 2-keto-3-methylvalerate all stimulate insulin secretion and increase 2-ketoglutarate generation in pancreatic B-cell mitochondria through activation of glutamate dehydrogenase and transamination withl-glutamate andl-glutamine, respectively. The neutral amino acidsl-valine,l-norleucine, andl-alanine and the aliphatic 2-keto acids 2-ketoisovalerate and pyruvate do not stimulate insulin secretion and do not increase 2-ketoglutarate generation in pancreatic B-cell mitochondria. Inhibition of 2-keto acid induced insulin secretion byl-valine andl-isoleucine is accompanied by reduced 2-ketoglutarate generation in pancreatic B-cell mitochondria. Thus intramitochondrial 2-ketoglutarate generation in pancreatic B-cells may regulate the insulin secretory potency of amino acids and 2-keto acids.  相似文献   

5.
The effect of various amino acid metabolites on glycine oxidation by rat liver homogenate was investigated. Three compounds, α-ketoisovaleric acid, α-ketoisocaproic acid, and α-keto-β-methylvaleric acid, were found to inhibit glycine oxidation by 40–60%. In addition, these compounds also inhibited the glycine-CO2 exchange reaction, a partial reaction of glycine synthase. The reverse reaction, glycine synthesis, was stimulated 4-fold by these α-keto acids. Pyruvate and α-ketoglutarate had no effect on any of these reactions. The parent amino acids, valine, isoleucine, and leucine, also had no effect on the reactions nor did any of their other metabolites with the exception of the branched-chain α-keto acids. The concentration dependence of the inhibition of glycine oxidation and stimulation of glycine synthesis by these branched-chain α-keto acids suggested that the inhibition of glycine oxidation by these compounds was the result of their further oxidation by branched-chain α-keto acid dehydrogenase. However, the products of the branched-chain α-keto acid dehydrogenase, isobutyryl CoA, isovaleryl CoA, or α-methylbutyryl CoA had no effect on glycine oxidation. Thus, it appeared that either the branched-chain α-keto acids altered glycine oxidation by direct binding to glycine synthase or that electrons derived from the oxidation of branched-chain α-keto acids were transferred to the glycine synthase system. It is proposed that glycine synthase and branched-chain α-keto acid dehydrogenase either share a common subunit, possibly lipoamide dehydrogenase, or are so arranged on the mitochondrial membrane that electron transfer between these two enzymes occurs.  相似文献   

6.
Addition of NADH to crude but not to pure branched-chain α-keto acid decarboxylase decreased the CO2 production from α-keto-β-methylvalerate (KMV) suggesting the presence of an NADH dependent inhibitor in the crude enzyme from Bacillus subtilis. This NADH-dependent decarboxylase inhibitor was purified to homogeneity by a fast protein liquid chromatography system.

The purified inhibitor was identical with leucine dehydrogenase as to N-terminal amino acid squence (35 residues) and molecular weight, and catalyzed the oxidative deamination of three branched chain amino acids (BCAAs), valine, leucine, and isoleucine. The decarboxylase inhibitor was therefore identified as leucine dehydrogenase. A decreased substrate availability caused by leucine dehydrogenase thus reasonably accounted for the NADH dependent inhibition of the decarboxylation. In turn, the observation that leucine dehydrogenase competes with the decarboxylase for branched-chain α-keto acid (BCKA) suggested an involvement of this enzyme in the branched chain fatty acid (BCFA) biosynthesis. This view was supported by the observation that addition of NAD to crude fatty acid synthetase increased the incorporation of isoleucine into BCFAs. Pyridoxal-5′-phosphate and α-ketoglutarate, cofactors for BCAA transaminase, modulated BCFA biosynthesis from isoleucine in vitro, suggesting also the involvement of transaminase reaction in BCFA biosynthesis.  相似文献   

7.
Two types of Pseudomonas putida PpG2 mutants which were unable to degrade branched-chain amino acids were isolated after mutagenesis and selection for ability to grow on succinate, but not valine, as a sole source of carbon. These isolates were characterized by growth on the three branched-chain amino acids (valine, isoleucine, and leucine), on the corresponding branched-chain keto acids (2-ketoisovalerate, 2-keto-3-methylvalerate, and 2-ketoisocaproate), and on other selected intermediates as carbon sources, and by their enzymatic composition. One group of mutants lost 2-ketoisovalerate-inducible branched-chain keto acid dehydrogenase that was active on all three keto acids. There was also a concomitant loss of ability to grow on all three branched-chain amino acids as well as on all three corresponding keto acids, but there was retention of ability to use subsequent intermediates in the catabolism of branched-chain amino acids. Another type of mutant showed a marked reduction in branched-chain amino acid transaminase activity and grew poorly at the expense of all three amino acids, but it utilized subsequent intermediates as carbon sources. Both the transaminase and branched-chain keto acid dehydrogenase mutants retained the ability to degrade camphor. These findings are consistent with the view that branched-chain amino acid transaminase and branched-chain keto acid dehydrogenase are common enzymes in the catabolism of valine, isoleucine, and leucine.  相似文献   

8.
Regulation of valine catabolism in Pseudomonas putida   总被引:2,自引:10,他引:2       下载免费PDF全文
The activities of six enzymes which take part in the oxidation of valine by Pseudomonas putida were measured under various conditions of growth. The formation of four of the six enzymes was induced by growth on d- or l-valine: d-amino acid dehydrogenase, branched-chain keto acid dehydrogenase, 3-hydroxyisobutyrate dehydrogenase, and methylmalonate semialdehyde dehydrogenase. Branched-chain amino acid transaminase and isobutyryl-CoA dehydrogenase were synthesized constitutively. d-Amino acid dehydrogenase and branched-chain keto acid dehydrogenase were induced during growth on valine, leucine, and isoleucine, and these enzymes were assumed to be common to the metabolism of all three branched-chain amino acids. The segment of the pathway required for oxidation of isobutyrate was induced by growth on isobutyrate or 3-hydroxyisobutyrate without formation of the preceding enzymes. d-Amino acid dehydrogenase was induced by growth on l-alanine without formation of other enzymes required for the catabolism of valine. d-Valine was a more effective inducer of d-amino acid dehydrogenase than was l-valine. Therefore, the valine catabolic pathway was induced in three separate segments: (i) d-amino acid dehydrogenase, (ii) branched-chain keto acid dehydrogenase, and (iii) 3-hydroxyisobutyrate dehydrogenase plus methylmalonate semialdehyde dehydrogenase. In a study of the kinetics of formation of the inducible enzymes, it was found that 3-hydroxyisobutyrate and methylmalonate semialdehyde dehydrogenases were coordinately induced. Induction of enzymes of the valine catabolic pathway was studied in a mutant that had lost the ability to grow on all three branched-chain amino acids. Strain PpM2106 had lowered levels of branched-chain amino acid transaminase and completely lacked branched-chain keto acid dehydrogenase when grown in medium which contained valine. Addition of 2-ketoisovalerate, 2-ketoisocaproate, or 2-keto-3-methylvalerate to the growth medium of strain PpM2106 resulted in induction of normal levels of branched-chain keto acid dehydrogenase; therefore, the branched-chain keto acids were the actual inducers of branched-chain keto acid dehydrogenase.  相似文献   

9.
Pseudomonas putida oxidized isoleucine to acetyl-coenzyme A (CoA) and propionyl-CoA by a pathway which involved deamination of d-isoleucine by oxidation and l-isoleucine by transamination, oxidative decarboxylation, and beta oxidation at the ethyl side chain. At least three separate inductive events were required to form all of the enzymes of the pathway: d-amino acid dehydrogenase was induced during growth in the presence of d-isoleucine; branched-chain keto dehydrogenase was induced during growth on 2-keto-3-methylvalerate and enzymes specific for isoleucine metabolism; tiglyl-CoA hydrase and 2-methyl-3-hydroxybutyryl-CoA dehydrogenase were induced by growth on isoleucine, 2-keto-3-methylvalerate, 2-methylbutyrate, or tiglate. Tiglyl-CoA hydrase and 2-methyl-3-hydroxybutyryl-CoA dehydrogenase were purified simultaneously by several enzyme concentration procedures, but were separated by isoelectric focusing. Isoelectric points, pH optima, substrate specificity, and requirements for enzyme action were determined for both enzymes. Evidence was obtained that the dehydrogenase catalyzed the oxidation of 2-methyl-3-hydroxybutyryl-CoA to 2-methylacetoacetyl-CoA. 2-Methyl-3-hydroxybutyryl-CoA dehydrogenase catalyzed the oxidation of 3-hydroxybutyryl-CoA, but l-3-hydroxyacyl-CoA dehydrogenase from pig heart did not catalyze the oxidation of 2-methyl-3-hydroxybutyryl-CoA; therefore, they appeared to be different dehydrogenases. Furthermore, growth on tiglate resulted in the induction of tiglyl-CoA hydrase and 2-methyl-3-hydroxybutyryl-CoA dehydrogenase, but these two enzymes were not induced during growth on crotonate or 3-hydroxybutyrate.  相似文献   

10.
Hypophysectomy caused a marked but transient increase in branched-chain α-keto acid decarboxylase activities in rat liver mitochondria, peaking at about nine days post-surgery. The magnitude of increase is different for each of the three branched-chain α-keto acids. The activities then fall to a new steady state in three weeks with α-ketoisovalerate decarboxylase activity within the normal range, α-keto-β-methylvalerate decarboxylase activity at twice normal, and α-ketoisocaproate decarboxylase activity decreased to a level too low for accurate measurements.  相似文献   

11.
High aminotransferase activities catalyzing the reactions between L-glutamate and L-glutamine and the aliphatic ketomonocarboxylic acids 2-ketoisocaproate, 2-ketocaproate, and 2-ketoisovalerate were observed in pancreatic B-cell mitochondria. While maximal rates of transamination with L-glutamate were observed in the presence of micromolar concentrations of keto acid, maximal rates of transamination with L-glutamine were recorded only in the presence of millimolar concentrations of keto acid. The insulin secretagogue 2-ketoisocaproate was the most effective transamination partner for L-glutamate, while the insulin secretagogue 2-ketocaproate was the most effective transamination partner for L-glutamine. Since B-cell mitochondria are well supplied with L-glutamate and L-glutamine, 2-ketoglutarate generation in the presence of these two neutral 2-keto acids may be an important prerequisite for their insulin secretory potency. High rates of transamination of 2-ketoglutarate were observed in the pancreatic B-cell mitochondria with the branched-chain amino acids L-leucine and L-valine, but not with L-norleucine. In connection with the ability of L-leucine to activate glutamate dehydrogenase, this high activity of the branched-chain amino acid aminotransferase in pancreatic B-cell mitochondria may provide an explanation for the insulin secretory potency of this amino acid.  相似文献   

12.
The effect of octanoate on the oxidative decarboxylation of 14C-labeled amino acids has been studied in perfused hindquarter and liver of rat. Regulation of the branched-chain α-keto acid dehydrogenase has been further studied with α-[14C-1]ketoisovalerate in isolated rat muscle and liver mitochondria. (1) Octanoate has a stimulatory effect on the oxidation of branched-chain amino acids in perfused hindquarter. The oxidative decarboxylation of other amino acids are inhibited. Octanoate inhibits the oxidative decarboxylation of all amino acids in perfused liver. (2) The oxidation of valine is stimulated by octanoate and hexanoate also in isolated muscle mitochondria. The stimulatory effect is probably related to activation of the fatty acids since acyl-carnitines inhibit the oxidation. (3) The oxidation of α-ketoisovalerate in mitochondria is inhibited by competing substrates (pyruvate, α-ketoglutarate and succinate). This inhibition is counteracted by octanoate and ADP. (4) Low concentrations (1–5 μM) of 2,4-dinitrophenol (DNP) activates wheras higher concentrations inactivates the branched-chain α-keto acid dehydrogenase in intact but not in solubilized muscle mitochondria. The inactivation is counteracted by ATP, but is increased by octanoate. (5) The observations seem to suggest that the activation (like the inactivation) of branched-chain α-keto acid dehydrogenase in skeletal muscle is dependent on the mitochondrial energy state which therefore may regulate both activation and inactivation of the dehydrogenase.  相似文献   

13.
Branched-chain keto acid dehydrogenase, an enzyme in the common pathway of branched-chain amino acid catabolism of Pseudomonas putida, is a multienzyme complex which catalyzes the oxidative decarboxylation of branched-chain keto acids. The objective of the present study was to isolate strains with mutations of this and other keto acid dehydrogenases and to map the location of the mutations on the chromosome of P. putida. Several strains with mutations of branched-chain keto acid dehydrogenase, two pyruvate and two 2-ketoglutarate dehydrogenase, were isolated, and the defective subunits were identified by biochemical analysis. By using a recombinant XYL-K plasmid to mediate conjugation, these mutations were mapped in relation to a series of auxotrophic and other catabolic mutations. The last time of entry recorded was at approximately 35 min, and the data were consistent with a single point of entry. Branched-chain keto acid dehydrogenase mutations affecting E1, E1 plus E2, and E3 subunits mapped at approximately 35 min. One other strain affected in the common pathway was deficient in branched-chain amino acid transaminase, and the mutation was mapped at 16 min. The mutations in the two pyruvate dehydrogenase mutants, one deficient in E1 and the other deficient in E1 plus E2, mapped at 22 minutes. The 2-ketoglutarate dehydrogenase mutation affecting the E1 subunit mapped at 12 minutes. A 2-ketoglutarate dehydrogenase mutant deficient in E3 was isolated, but the mutation proved too leaky to map.  相似文献   

14.
Isopycnic sucrose gradient separation of rat liver organelles revealed the presence of two distinct branched-chain α-keto acid decarboxylase activities; a mitochondrial activity, which decarboxylates the three branched-chain α-keto acids and requires CoA and NAD+ and a cytosolic activity, which decarboxylates α-ketoisocaproate, but not α-ketoisovalerate, or α-keto-β-methylvalerate. The latter enzyme does not require added CoA or NAD+. Assay conditions for the cytosolic α-ketoisocaproate decarboxylase activity were optimized and this activity was partially characterized. In rat liver cytosol preparations this activity has a pH optimum of 6.5 and is activated by 1.5 m ammonium sulfate. The decarboxylase activity has an apparent Km of 0.03 mm for α-ketoisocaproate when optimized assay conditions are employed. Phenylpyruvate is a very potent inhibitor. α-Ketoisovalerate, α-keto-β-methylvalerate, α-ketobutyrate, and α-ketononanoate also inhibit the α-ketoisocaproate decarboxylase activity. The data indicate that the soluble α-ketoisocaproate decarboxylase is an oxidase. Rat liver cytosol preparations consumed oxygen when either α-ketoisocaproate or α-keto-γ-methiolbutyrate were added. None of the other α-keto acids tested stimulated oxygen consumption. 1-14C-Labeled α-keto-γ-methiolbutyrate is also decarboxylated by cytosol preparations. The α-ketoisocaproate oxidase was purified 20-fold from a 70,000g supernatant fraction of a rat liver homogenate. In these preparations the activity was increased 4-fold by the addition of dithiothreitol, ferrous iron, and ascorbate. The major product of this enzyme activity is β-hydroxyisovalerate. Isovalerate is not a free intermediate in the reaction. The data indicate an alternative pathway for metabolism of α-ketoisocaproate which produces β-hydroxyisovalerate.  相似文献   

15.
Treatment of pure 2-keto-4-hydroxyglutarate aldolase from Escherichia coli, a "lysine-type," Schiff-base mechanism enzyme, with the substrate analog bromopyruvate results in a time- and concentration-dependent loss of enzymatic activity. Whereas the substrates pyruvate and 2-keto-4-hydroxyglutarate provide greater than 90% protection against inactivation by bromopyruvate, no protective effect is seen with glycolaldehyde, an analog of glyoxylate. Inactivation studies with [14C] bromopyruvate show the incorporation of 1.1 mol of 14C-labeled compound/enzyme subunit; isolation of a radioactive peptide and determination of its amino acid sequence indicate that the radioactivity is associated with glutamate 45. Incubation of the enzyme with excess [14C]bromopyruvate followed by denaturation with guanidine.HCl allow for the incorporation of carbon-14 at cysteines 159 and 180 as well. Whereas the presence of pyruvate protects Glu-45 from being esterified, it does not prevent the alkylation of these 2 cysteine residues. The results indicate that Glu-45 of E. coli 2-keto-4-hydroxyglutarate aldolase is essential for catalytic activity, most likely acting as the amphoteric proton donor/acceptor that is required as a participant in the overall mechanism of the reaction catalyzed.  相似文献   

16.
A Saha  Y C Huang  R F Colman 《Biochemistry》1989,28(21):8425-8431
The substrate affinity label 3-bromo-2-ketoglutarate (BrKG) reacts covalently with pig heart NAD+-specific isocitrate dehydrogenase with complete inactivation and incorporation of about 0.8 mol of reagent/mol of average enzyme subunit [Bednar, R.A., Hartman, F.C., & Colman, R.F. (1982) Biochemistry 21, 3681-3689]. Protection against inactivation is provided by isocitrate and Mn2+. We have now identified a critical modified peptide by comparison of the peptides labeled by BrKG at pH 6.1 in the absence and presence of isocitrate and Mn2+. Modified enzyme, isolated from unreacted BrKG, was incubated with [3H]NaBH4 to reduce the keto group of protein-bound 2-ketoglutarate and thereby introduce a radioactive tracer into the modified amino acid. Following carboxymethylation and digestion with trypsin, the specific modified peptide was isolated by reverse-phase HPLC, first in 0.1% trifluoroacetic acid with a gradient in acetonitrile and then in 20 mM ammonium acetate, pH 5.8, with an acetonitrile gradient. Gas-phase sequencing gave the modified peptide: Ser-Ala-X-Val-Pro-Val-Asp-Phe-Glu-Glu-Val-Val-Val-Ser-Ser-Asn-Ala-Asp-Gl u-Glu- Asp-Ile-Arg. The corresponding tryptic peptide that was isolated from unmodified enzyme yielded the same sequence except for (carboxymethyl)cysteine at position 3, suggesting that cysteine is the target of 3-bromo-2-ketoglutarate. Pig heart NAD+-dependent isocitrate dehydrogenase is composed of three distinct subunits (alpha, beta, and gamma) that can be separated by chromatofocusing in urea and identified by analytical gel isoelectric focusing. The peptide modified by 3-bromo-2-ketoglutarate, which is in or near the substrate site, is derived only from the separated gamma subunit.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
In the absence of any other oxidizable substrate, the perfused rat heart oxidizes [1-14C]leucine to 14CO2 at a rapid rate and releases only small amounts of α-[1-14C]ketoisocaproate into the perfusion medium. The branched-chain α-keto acid dehydrogenase complex, assayed in extracts of mitochondria prepared from such perfused hearts, is very active. Under such perfusion conditions, dichloroacetate has almost no effect on [1-14C]leucine oxidation, α-[1-14C]ketoisocaproate release, or branched-chain α-keto acid dehydrogenase activity. Perfusion of the heart with some other oxidizable substrate, e.g., glucose, pyruvate, ketone bodies, or palmitate, results in an inhibition of [1-14C]leucine oxidation to 14CO2 and the release of large amounts of α-[1-14C]ketoisocaproate into the perfusion medium. The branched-chain α-keto acid dehydrogenase complex, assayed in extracts of mitochondria prepared from such hearts, is almost completely inactivated. The enzyme can be reactivated, however, by incubating the mitochondria at 30 °C without an oxidizable substrate. With hearts perfused with glucose or ketone bodies, dichloroacetate greatly increases [1-14C]leucine oxidation, decreases α-[1-14C]ketoisocaproate release into the perfusion medium, and activates the branched-chain α-keto acid dehydrogenase complex. Pyruvate may block dichloroacetate uptake because dichloroacetate neither stimulates [1-14C]leucine oxidation nor activates the branched-chain α-keto acid dehydrogenase complex of pyruvate-perfused hearts. It is suggested that leucine oxidation by heart is regulated by the activity of the branched-chain α-keto acid dehydrogenase complex which is subject to interconversion between active and inactive forms. Oxidizable substrates establish conditions which inactivate the enzyme. Dichloroacetate, known to activate the pyruvate dehydrogenase complex by inhibition of pyruvate dehydrogenase kinase, causes activation of the branched-chain α-keto acid dehydrogenase complex, suggesting the existence of a kinase for this complex.  相似文献   

18.
Polyclonal antibodies directed against the dihydrolipoyl transacylase (E2) and alpha subunit of branched-chain alpha-keto acid decarboxylase (E1 alpha) components of the bovine branched-chain keto acid dehydrogenase complex were shown to cross-react with the E2 and E1 alpha polypeptides of the enzyme complex of different rat tissues. Phosphorylation of the branched-chain keto acid dehydrogenase complex resulted in inhibition of enzyme activity concomitant with phosphate incorporation into the E1 alpha polypeptide. Phosphorylation of E1 alpha slowed its rate of migration through sodium dodecyl sulfate-polyacrylamide gels. This permitted resolution of the phosphorylated and unphosphorylated forms of E1 alpha on immunoblots. Liver and skeletal muscle mitochondria were prepared from rats consuming 6, 20, or 50% casein diets. The enzyme complex in mitochondria was measured by radioisotopic enzyme assay and immunoassay. Liver branched-chain keto acid dehydrogenase was 25% active in rats consuming 6% casein diets; whereas in rats consuming 20 or 50% casein diets, the liver enzyme was 82 or 100% active, respectively. Branched-chain keto acid dehydrogenase of muscle was 10, 13, and 22% active, respectively, in rats consuming 6, 20, and 50% casein diets. The amount of protein consumed by rats did not affect the total amount of the enzyme complex per unit of mitochondrial protein as measured by either the radioisotopic assay (enzyme activity) or the immunoassay. However, the protein intake of rats did affect activity of the enzyme kinase in liver. Liver branched-chain keto acid dehydrogenase kinase was more active in rats consuming 6% casein than in those fed chow or 50% casein diets. The amount of protein consumed by rats thus influences the enzyme activity in liver and muscle by affecting the reversible phosphorylation mechanism and not by induction of branched-chain keto acid dehydrogenase.  相似文献   

19.
A whole cell biotransformation system for reductive amination has been studied in recombinant Escherichia coli cells. Reductive amination of 2-keto-3-methylvalerate to l-isoleucine by a two-enzyme-cascade was achieved by overproduction of endogenous l-alanine dependent transaminase AvtA and heterologous l-alanine dehydrogenase from Bacillus subtilis in recombinant E. coli. Up to 100 mM l-isoleucine were produced from 100 mM 2-keto-3-methylvalerate and 100 mM ammonium sulfate. Regeneration of NADH as cofactor in the whole cell system was driven by glucose catabolism. The effects of defined gene deletions in the central carbon metabolism on biotransformation were tested. Strains lacking the NuoG subunit of NADH:ubiquinone oxidoreductase (complex I) or aceA encoding the glyoxylate cycle enzyme isocitrate lyase exhibited increased biotransformation rates.  相似文献   

20.
X Mai  M W Adams 《Journal of bacteriology》1996,178(20):5890-5896
Thermococcus litoralis is a strictly anaerobic archaeon (archaebacterium) that grows at temperatures up to 98 degrees C by fermenting peptides. It is known to contain three distinct ferredoxin-dependent, 2-keto acid oxidoreductases, which use pyruvate, aromatic 2-keto acids such as indolepyruvate, or branched-chain 2-keto acids such as 2-ketoisovalerate, as their primary substrates. We show here that T. litoralis also contains a fourth member of this family of enzymes, 2-ketoglutarate ferredoxin oxidoreductase (KGOR). In the presence of coenzyme A, KGOR catalyzes the oxidative decarboxylation of 2-ketoglutarate to succinyl coenzyme A and CO2 and reduces T. litoralis ferredoxin. The enzyme was oxygen sensitive (half-life of approximately 5 min) and was purified under anaerobic conditions. It had an M(r) of approximately 210,000 and appeared to be an octomeric enzyme (alpha2beta2gamma2delta2) with four different subunits with M(r)s of 43,000 (alpha), 29,000 (beta), 23,000 (gamma), and 10,000 (delta). The enzyme contained 0.9 mol of thiamine PPi and at least four [4Fe-4S] clusters per mol of holoenzyme as determined by metal analyses and electron paramagnetic resonance spectroscopy. Significant amounts of other metals (Cu, Zn, Mo, W, and Ni) were not present (<0.1 mol/mol of holoenzyme). Pure KGOR did not utilize other 2-keto acids, such as pyruvate, indolepyruvate, or 2-ketoisovalerate, as substrates, and the apparent Km values for 2-ketoglutarate, coenzyme A, T. litoralis ferredoxin, and thiamine PPi were approximately 250, 40, 8, and 9 microM, respectively. The enzyme was virtually inactive at 25 degrees C and exhibited optimal activity above 90 degrees C (at pH 8.0) and at pH 8.0 (at 80 degrees C). KGOR was quite thermostable, with a half-life at 80 degrees C (under anaerobic conditions) of about 2 days. An enzyme analogous to KGOR has been previously purified from a mesophilic archaeon, but the molecular properties of T. litoralis KGOR more closely resemble those of the other oxidoreductases from hyperthermophiles. In contrast to these enzymes, however, KGOR appears to have a biosynthetic function rather than a role in energy conservation.  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