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.
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   

2.
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   

3.
—The effects of the amino acids (phenylalanine, valine, leucine and isoleucine) which accumulate in phenylketonuria (PKU) and maple syrup urine disease (MSUD), and their analogue α-keto acids (phenylpyruvate, α-keto isovalerate, α-keto isocaproate, α-keto-β-Me valerate) have been studied on rat brain mitochondrial respiration. Both phenylpyruvate and α-keto isocaproate specifically inhibited the oxidation of pyruvate plus malate and β-hydroxybutyrate plus malate by rat brain mitochondria in the presence of ADP. However, no inhibitory effects of similar concentrations of phenylpyruvate or α-keto isocaproate were observed on the isolated semipurified pyruvate or β-hydroxybutyrate dehydrogenases from rat brain mitochondria. The transport of pyruvate and β-hydroxybutyrate across the brain mitochondrial membrane was studied by both uptake and exchange of radioactively labelled substrates. Both these processes were inhibited by phenylpyruvate and α-ketoisocaproate. The results are interpreted as providing evidence for both pyruvate and β-hydroxybutyrate translocases across the brain mitochondrial membrane, and that the inhibition of these systems by phenylpyruvate and α-keto isocaproate may be important lesions in phenylketonuria and maple syrup urine disease respectively.  相似文献   

4.
Seedlings of mung bean (Phaseolus aureus) contain a soluble enzyme capable of converting l-tryptophan to indole-3-pyruvic acid by transamination. The concentration of the enzyme is highest in the stem meristem and primary leaves and lowest in the roots. The enzyme was purified 28.6 fold by ammonium sulphate precipitation, Sephadex G-200 filtration, and electrophoresis. The isoelectric point of the enzyme protein was pH 6.6. The optimum pH and temperature for the catalytic conversion were ca. 8.5 and 53°C respectively. Using l -tryptophan and α-ketoglutarate as substrates Km was found to be 3.3 × 10?4 M and the activation energy 18,270 cal per mole. The enzyme converted only the l -form of tryptophan, phenylalanine, tyrosine, and histidine. Out of 13 other l -amino acids tested 8 could be transaminated. Eight α-keto acids tested could all be used as substrates. High efficiency of an α-keto acid as an amino group acceptor agreed usually with high efficiency of the corresponding amino acid as a donor. The pari ß-methyl-α-ketoisovaleric acid and isoleucine was an exception to that rule. Addition of pyridoxalphosphate to the reaction mixture was not needed. The indole-3-pyruvic acid formed in the reaction was trapped and partly stabilized as its borate complex and measured spectrophotometrically at 327 nm. The keto acid formed was further identified by chromatography of its 2,4-dinitrophenylhydrazone in 4 solvent systems. When using α-keto-glutaric acid as a substrate, the glutamic acid produced was determined by the glutamate dehydrogenase method. The sensitivity of the assay permits enzyme determinations in extracts from 5 mg leaves or 100 mg roots.  相似文献   

5.
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   

6.
Abstract

Branched-chain amino acids (BCAAs) exhibit many physiological functions. However, the potential link and mechanism between BCAA and skin function are unknown. We examined the effects of deletion of branched-chain α-keto acid dehydrogenase kinase (BDK), a key enzyme in BCAA catabolism, on type I and III tropocollagen syntheses in mice. Leucine and isoleucine levels were significantly lower in the skin of BDK-KO mice compared with wild-type mice. No changes in valine concentrations were observed. The levels of type I and III tropocollagen proteins and mRNAs (COL1A1 and COL3A1) were significantly lower in the skin of BDK-KO mice compared with wild-type mice. The phosphorylation of p70 S6 kinase, which indicates mammalian target of rapamycin (mTOR) activation, was reduced in the skin of BDK-KO mice compared with wild-type mice. These findings suggest that deficiencies of leucine and isoleucine reduce type I and III tropocollagen syntheses in skin by suppressing the action of mTOR.  相似文献   

7.
3-Mercaptopyruvate sulfurtransferase catalyzes the transfer of sulfur from 3-mercaptopyruvate to several possible acceptor molecules, one of which is cyanide. Because the transsulfuration of cyanide is the primary in vivo mechanism of detoxification, 3-mercaptopyruvate sulfurtransferase may function in the enzymatic detoxification of cyanide in vivo. Three α-keto acids (α-ketobutyrate, α-ketoglutarate, and pyruvate) have previously been demonstrated to be cyanide antidotes in vivo, and it has been suggested that this is due to the nonenzymatic binding of cyanide by the α-keto acid. However, it has also been proposed that α-keto acids may increase the activity of enzymes involved in the transsulfuration of cyanide. Thus, the effect of these three α-keto acids on the enzyme 3-mercaptopyruvate sulfurtransferase was examined. All three α-keto acids inhibited 3-mercaptopyruvate sulfurtransferase in a concentration-dependent manner and were determined to be uncompetitive inhibitors of MST with respect to 3-mercaptopyruvate. The inhibitor constant Ki was estimated by two methods for each inhibitor and ranged from 4.3 to 6.3 mM. The I50, which is the inhibitor concentration that produces 50% inhibition, was calculated for all three α-keto acids and ranged between 9.5 and 13.7 mM. These observations add further support to the hypothesis that the mechanisms of the α-keto acid antidotes is the nonenzymatic binding of cyanide, not stimulation of enzymes involved in the transsulfuration of cyanide to thiocyanate. © 1996 John Wiley & Sons, Inc.  相似文献   

8.
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   

9.
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   

10.
By use of a relatively new mixed stationary phase, complete separation of the branched-chain α-keto acids as O-trimethylsilyl-quinoxalinol derivatives is achieved within 10 min by packed column gas chromatography. Precise quantification of less than 5 nmol of α-keto acids in biological samples is possible. In small aqueous samples the α-keto acids are directly derivatized without prior purification. Plasma need only be deproteinized by perchlorate and neutralized before derivatization. Average relative precision for determination of the three main branched-chain α-keto acids is ± 5.8%.  相似文献   

11.
Maple syrup urine disease (MSUD) is a metabolic disorder caused by the deficiency of the activity of the mitochondrial enzyme complex branched-chain l-2-keto acid dehydrogenase. The metabolic block results in tissue and body fluid accumulation of the branched-chain amino acids leucine (Leu), isoleucine and valine, as well as of their respective α-keto acids. Neurological sequelae are usually present in MSUD, but the pathophysiologic mechanisms of neurotoxicity are still poorly known. It was previously demonstrated that Leu elicits oxidative stress in rat brain. In the present study we investigated the possible mechanisms involved in Leu-induced oxidative damage. We observed a significant attenuation of Leu-elicited increase of thiobarbituric acid-reactive substances (TBA-RS) measurement when cortical homogenates were incubated in the presence of the free radical scavengers ascorbic acid plus trolox, dithiothreitol, glutathione, and superoxide dismutase, suggesting a probable involvement of superoxide and hydroxyl radicals in this effect. In contrast, the use of Nω-nitro-l-arginine methyl ester or catalase (CAT) did not affect TBA-RS values. We also demonstrated an inhibitory effect of Leu on the activities of the antioxidant enzymes CAT and gluthathione peroxidase, as well as a significant reduction in the membrane-protein thiol content from mitochondrial enriched preparations. Furthermore, dichlorofluorescein levels were increased although not significantly by Leu. Taken together, our present data indicate that an unbalance between free radical formation and inhibition of critical enzyme activities may explain the mechanisms involved in the Leu-induced oxidative damage.  相似文献   

12.
Various α-keto acids were separated as their 2,4-dinitrophenylhydrazine derivatives by ion-pair, reverse-phase, high-pressure liquid chromatography. Excellent baseline resolution was obtained for a seven-component homologous series of α-keto acid dinitrophenylhydrazones at increasing carbon-chain length. Branched-chain keto acids were also separated. Resolution of syn and anti isomers of the α-keto acid derivatives was possible. Pyruvate from biological material was located and identity confirmed by an enzymic peak shift technique. Monitoring at 366 nm permits low-level (nanogram) amounts of keto acids to be detected. Ion pair versus ion exchange is discussed with regard to the mechanism of chromatographic separation.  相似文献   

13.
Electro-olfactograms were used to determine sensitivity and specificity of olfactory organs of female sea lampreys (Petromyzon marinus) to four bile acids: 3-keto petromyzonol sulfate and 3-keto allocholic acid from spermiating males and petromyzonol sulfate and allocholic acid from larvae. Spermiating male bile acids are thought to function as a mating pheromone and larval bile acids as a migratory pheromone. The response threshold was 10–12 mol l–1 for 3-keto petromyzonol sulfate and 10–10 mol l–1 for the other bile acids. At concentrations above 10–9 mol l–1, the sulfated bile acids showed almost identical potency, as did the non-sulfated bile acids. The two sulfated bile acids were more potent than the two non-sulfated ones. In addition, 3-keto petromyzonol sulfate and water conditioned with spermiating males induced similar concentration-response curves and response thresholds. Cross-adaptation experiments demonstrated that the sulfated and non-sulfated bile acids represent different odors to the olfactory epithelium of females. Further exploration revealed that 3-keto petromyzonol sulfate represents a different odor than petromyzonol sulfate, while 3-keto allocholic acid and allocholic acid represent the same odor. Results indicate that male-specific bile acids are potent and specific stimulants to the female olfactory organ, supporting the previous hypothesis that these bile acids function as a pheromone.Abbreviations 3kACA 3-keto allocholic acid - 3kPZS 3-keto petromyzonol sulfate - ACA allocholic acid - ANOVA analysis of variance - ELISA enzyme-linked immunosorbent assay - EOG electro-olfactogram - PIR percent initial response - PZS petromyzonol sulfate - SMW spermiating male washings  相似文献   

14.
We report a highly enantioselective oligomerization of isoleucine stereomers in the salt-induced peptide formation reaction under plausibly prebiotic earth conditions. Up to 6.5-fold superiority in reactivity of l-isoleucine was observed, compared to its d-enantiomer, after 14 evaporation cycles in the presence of Cu2+ and NaCl. Since isoleucine is among the proteinogenic amino acids that were found enantioenriched in meteorites, this present work may further correlate the extraterrestrial delivery and endogenous production of biological homochirality by virtue of a protein constituent rather than the rarely occurring α-methylated amino acids.  相似文献   

15.
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   

16.
A high-performance liquid chromatographic method has been developed for the determination of α-keto acids in human urine and plasma. These acids were prepurified using a column of hydrazide gel and derivatized with o-phenylenediamine into 2-quinoxalinol derivatives, which were extracted into ethyl acetate. The 2-quinoxialinol derivatives were separated by reversed-phase paired-ion chromatography using a 250 × 4 mm-i.d. column packed with LiChrosorb RP-8 (5 μm). This method is sensitive, selective, and reproducible. The α-keto acids in urine and plasma from normal individuals were determined.  相似文献   

17.
A method is described for the quantitative determination of aliphatic α-keto acids in urine after derivatization with o-phenylenediamine and bis(trimethylsilyl)trifluoroacetamide, α-Ketovaleric acid and α-ketocaprylic acid are used as internal standards.The chemical yield is 80–100%. At physiological concentrations, the coefficient of variation after repeated derivatizations is 4% for pyruvic acid and 14% for α-ketoglutaric acid.With mass spectrometric single-ion detection at mc = 217, 232 and 245, the biologically interesting aliphatic α-keto acids can be determined at very low levels in biological fluids.  相似文献   

18.
As an extension of earlier work on aliphatic α-keto acids, a method is described for the quantitative gas chromatographic determination of urinary aromatic α-keto acids. The keto acids are derivatized with o-phenylenediamine to yield the quinoxalinols. These compounds are chromatographed after trimethylsilyation.The aromatic keto acids are stabilized by sodium dithionite (4 mg/ml urine) and storage below 0°. The final derivatives are stable for weeks at room temperature.Low resolution mass spectra are reported. The fragmentation mechanims are elucidated by analysis of O-trimethylsilyl-(TMS)-quinoxalinols, O-(TMS-d9)-quinoxalinols and O-TMS-6(7)-chloroquinoxalinols.  相似文献   

19.
R Blomstrand  L Kager  O Lantto 《Life sciences》1973,13(8):1131-1141
In order to elucidate the mechanisms involved in the acute ethanol-induced liver triglyceride accumulation, the oxidation, esterification and β-keto acid formation have been studied in rat and human liver slices after incubation with albumin bound, long chain fatty acids (palmitic. oleic and linoleic acids).The addition of alcohol to rat and human liver slices depressed the formation of 14CO2 from palmitic acid-1-14C, oleic acid-1-14C and linoleic acid-1-14C. The esterification to triglycerides and phospholipids was increased and the formation of β-keto acids was decreased by alcohol.Addition of 4-methylpyrazole, an inhibitor of liver alcohol dehydrogenase, almost prevented the alcohol effect on the lipid metabolism of the liver slices. The oxidation of alcohol is thus obligatory for the decreased β-oxidation of fatty acids, the increased esterification and for the decreased formation of β-keto acids. The results suggest that ethanol oxidation inhibits β-oxidation of fatty acids and that this primary effect leads to accumulation of liver triglycerides by increased esterification of plasma free fatty acids.  相似文献   

20.
《Inorganica chimica acta》1988,152(3):181-188
The rates of the transamination reactions of α-amino acids and α-keto acids were followed by measurement of the 200 MHz proton NMR spectra of solution species as a function of time. Reaction systems measured in D2O at 10 °C consisted of 1:1:1 molar ratios of pyridoxal:α-amino acid:Al(III) or pyridoxamine:α-keto acid:Al(III). Amino and keto acids employed are alanine, α-aminoisobutyric acid, valine, phenylglycine, pyruvic acid, and α-ketobutyric acid. A negatively charged deprotonated Schiff base coordinated to Al(III) was detected in all systems that undergo transamination (i.e., except α-aminoisobutyric acid). The intermediate resembles the aldimine Al(III) chelate with NMR resonances shifted upfield in accordance with its greater negative charge. Its equilibrium concentration is reached in the time required to reach transamination equilibrium and is maintained in solution at a ca. 10–20% of the aldimine Schiff base concentration.  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