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.
Acetate and butanoate inhibited and hexanoate and octanoate increased the 14CO2 production from 0.1 mM [1-14C]-labelled 2-oxoisocaproate (KIC) and 2-oxoisovalerate (KIV) in rat hemidiaphragms. Octanoate increased KIC and KIV oxidation in rat soleus muscle, too, inhibited it in human skeletal muscle and had a divergent effect in rat and human heart slices. In rat hemidiaphragms octanoate primarily affected the process of oxidative decarboxylation. No effect was found on transamination rates of branched-chain amino acids and on the CO2 production beyond alpha-decarboxylation. The reverse transamination of branched-chain 2-oxo acids and their incorporation into protein decreased in the presence of octanoate. Octanoate had no effect on KIC and KIV oxidation at higher 2-oxo acid concentrations and in hemidiaphragms from 3-day-starved rats. The observed interactions are discussed and related to regulatory mechanisms, which are known to affect the branched-chain 2-oxo acid dehydrogenase complex.  相似文献   

2.
Oxidative decarboxylation and transamination of 1-14C-branched chain amino and alpha-keto acids were examined in mitochondria isolated from rat heart. Transamination was inhibited by aminooxyacetate, but not by L-cycloserine. At equimolar concentrations of alpha-ketoiso[1-14C]valerate (KIV) and isoleucine, transamination was increased by disrupting the mitochondria with detergent which suggests transport may be one factor affecting the rate of transamination. Next, the subcellular distribution of the aminotransferase(s) was determined. Branched chain aminotransferase activity was measured using two concentrations of isoleucine as amino donor and [1-14C]KIV as amino acceptor. The data show that branched chain aminotransferase activity is located exclusively in the mitochondria in rat heart. Metabolism of extramitochondrial branched chain alpha-keto acids was examined using 20 microM [1-14C]KIV and alpha-ketoiso[1-14C]caproate (KIC). There was rapid uptake and oxidation of labeled branched chain alpha-keto acid, and, regardless of the experimental condition, greater than 90% of the labeled keto acid substrate was metabolized during the 20-min incubation. When a branched chain amino acid (200 microM) or glutamate (5 mM) was present, 30-40% of the labeled keto acid was transaminated while the remainder was oxidized. Provision of an alternate amino acceptor in the form of alpha-keto-glutarate (0.5 mM) decreased transamination of the labeled KIV or KIC and increased oxidation. Metabolism of intramitochondrially generated branched chain alpha-keto acids was studied using [1-14C]leucine and [1-14C]valine. Essentially all of the labeled branched chain alpha-keto acid produced by transamination of [1-14C]leucine or [1-14C]valine with a low concentration of unlabeled branched chain alpha-keto acid (20 microM) was oxidized. Further addition of alpha-ketoglutarate resulted in a significant increase in the rate of labeled leucine or valine transamination, but again most of the labeled keto acid product was oxidized. Thus, catabolism of branched chain amino acids will be favored by a high concentration of mitochondrial alpha-ketoglutarate and low intramitochondrial glutamate.  相似文献   

3.
Human liver BCKADH complex was purified. On SDS-polyacrylamide gel electrophoresis, the purified enzyme complex gave three major bands having molecular weights of 51,000, 46,000, and 36,000, and one minor band with a molecular weight of 55,000. The minor band corresponded in molecular weight to lipoamide oxidoreductase which was purified separately. The purified BCKADH represented only approximately 20% of the maximum activity when assayed without addition of exogenous lipoamide oxidoreductase, indicating that lipoamide oxidoreductase component was readily dissociable from the complex. The BCKADH effectively oxidized all of KIV, KIC, and KMV, yielding apparent Km values in the range of 14-17 microM for those alpha-keto acids. Vmax values obtained were 0.86, 0.61, and 0.51 mumole NADH produced/min/mg of protein for KIV, KIC, and KMV, respectively, in the presence of excess amount of lipoamide oxidoreductase. This ratio of Vmax values was practically identical to those of specific activities obtained with respective branched-chain alpha-keto acids at each purification step. The enzyme complex also oxidized pyruvate and alpha-ketoglutarate to a lesser extent. Kinetic experiments gave Km values of 0.98 and 2.9 mM for pyruvate and alpha-ketoglutarate, respectively, with Vmax of 0.43 and 0.08 mumole NADH produced/min/mg of protein. NAD and CoASH were absolutely required for the reaction. Km values for NAD and CoASH were estimated to be 47 and 25 microM, respectively.  相似文献   

4.
The interaction of various metabolites and agents with the 14CO2 production from 0.1 mM [1-14C]-labelled 2-oxoisocaproate (KIC) and 2-oxoisovalerate (KIV) was studied in rat and human heart and skeletal muscle preparations. Glucose and carnitine had no effect in any of the studied systems; palmitate gave a small increase of KIC oxidation only in soleus muscle. With rat hemidiaphragms a considerable decrease was found in the presence of high concentrations of a competitive branched-chain 2-oxo acid and of pyruvate, and in the presence of ketone bodies. A considerable increase was found in the presence of the branched-chain 2-oxo acid dehydrogenase kinase inhibitor 2-chloroisocaproate and the transminase inhibitor amino-oxyacetate. 2-Oxoglutarate increased and clofibric acid decreased only KIC oxidation. Divergent effects were given by intermediates of the degradation route of KIC and KIV and by monocarboxylate translocator inhibitors. The observed interactions are discussed and related to regulatory mechanisms which are known to affect the branched-chain 2-oxo acid dehydrogenase complex.  相似文献   

5.
Branched-chain alpha-keto acid dehydrogenase (BCKADH) was solubilized as an enzyme complex from rat liver mitochondria by sonic treatment. Dehydrogenase (E1) and dihydrolipoyltransacylase (E2) components of the complex were purified in an associated form and resolved into individual components in the presence of 1 M NaCl, while lipoamide dehydrogenase (E3) component was dissociated from the complex during purification. Analysis by gel electrophoresis in dodecyl sulfate revealed the E1 comprised two different subunits with apparent molecular weights of 36,000 and 45,500, presumably in an equal molar ratio, while E2 consisted of a single subunit with an apparent molecular weight of 51,000. The BCKADH complex was reconstituted by combining E1, E2, and E3, and the formation of the complex was confirmed by analysis by sucrose density gradient centrifugation. The reconstituted enzyme complex oxidized not only alpha-ketoisovalerate (KIV), alpha-ketoisocaproate (KIC), and alpha-keto-beta-methylvalerate (KMV), but also pyruvate and alpha-ketoglutarate. Apparent Km values were 10-12 microM for the branched-chain alpha-keto acids, 2.2 mM for pyruvate, and 2.5 mM for alpha-ketoglutarate.  相似文献   

6.
There was net uptake of branched-chain keto acids by the fetus from the umbilical circulation. Mean fetal uptake of the 3 keto acids 2-keto isovalerate, 2-keto isocaproate and 2-keto methylvalerate was 1.8 mumol/min per kg of fetus. The concentrations in the umbilical vein for these keto acids were 10.9 +/- 3.8 microM (mean +/- SD: 2-keto isovalerate), 19.7 +/- 6.1 microM (2-keto isocaproate) and 14.8 +/- 5.3 microM (2-keto methylvalerate) respectively. The coefficients of extraction for the same keto acids were 17.2%, 16.8% and 11.9% respectively. Fetal uptakes (both mumol/min and mumol/min per kg fetus) were positively correlated with umbilical supply. There were concentration gradients across the placenta, with fetal concentration: maternal concentration ratios of 3.3 +/- 1.5 for 2-keto isovalerate, 2.1 +/- 0.8 for 2-keto isocaproate and 1.3 +/- 0.6 for 2-keto methylvalerate. The net release of 2-keto acids into the umbilical circulation may conserve the carbon skeleton of branched-chain amino acids for fetal metabolism and growth. In the uterine circulation there was not a consistent pattern of release from or uptake by the uteroplacental tissues. It is suggested that branched-chain keto acids may contribute to fetal growth or energy metabolism.  相似文献   

7.
1. The oxidation of the three branched-chain amino acids was regulated in parallel fashion in rat tissues studied in vitro. 2. With 0.1 mM-[1-14C]isoleucine as substrate in the presence of 5.5 mM-glucose, 14CO2 production was 0.6 mumol/2 h per g in the aorta, 0.3 in peripheral nerve, 0.2 in muscle and 0.13 in spinal cord. 3. The ratio 14C oxidized/14C incorporated into proteins with 0.1 mM-[1-14C]leucine was 1.3 in hemidiaphragms, 3.3 in sciatic nerve and 1.0 in nerves undergoing Wallerian degeneration. Leucine oxidation decreased only slightly during degeneration, but protein synthesis doubled. 4. Hemidiaphragms incubated with [1-14C]leucine or 4-methyl-2-oxo[1-14C]pentanoate increased 14CO2 production 7-9-fold as substrate concentration was increased from 0.1 to 0.5 mM; under the same conditions 14CO2 production by nerves increased only 2-3-fold. 5. 2-Oxoglutarate stimulated the oxidation of the branched-chain amino acids by muscles and peripheral nerves and the oxidation of 4-methyl-2-oxopentanoate by hemidiaphragms but not by nerves. 6. Octanoate (0.1-1.0 mM) markedly stimulated the oxidation of branched-chain amino acids and of 4-methyl-2-oxopentanoate in hemidiaphragms, but inhibited oxidation of both by peripheral nerves and spinal cord. In aortas, oxidation of isoleucine (the only substance tested) was inhibited by octanoate. 7. The effects of octanoate and 2-oxoglutarate on leucine oxidation by hemidiaphragms were additive at low concentrations. When maximally stimulating concentrations of either agent were used, addition of the other was ineffective. 8. Pyruvate inhibited the oxidation of branched-chain amino acids and 4-methyl-2-oxopentanoate in all tissues tested. 9. Insulin did not affect the oxidation of 4-methyl-2-oxopentanoate by muscles or nerves. 10. The oxidative decarboxylation of the branched-chain alpha-oxo acids is suggested as a regulatory site of branched-chain amino acid oxidation. Differences in regulation between muscle on the one hand, and nerve and aorta on the other, are discussed.  相似文献   

8.
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   

9.
Parameters of branched-chain amino acids (BCAA; leucine, isoleucine and valine) and protein metabolism were evaluated using L-[1-(14)C]leucine and alpha-keto[1-(14)C]isocaproate (KIC) in the whole body and in isolated perfused liver (IPL) of rats fed ad libitum or starved for 3 days. Starvation caused a significant increase in plasma BCAA levels and a decrease in leucine appearance from proteolysis, leucine incorporation into body proteins, leucine oxidation, leucine-oxidized fraction, and leucine clearance. Protein synthesis decreased significantly in skeletal muscle and the liver. There were no significant differences in leucine and KIC oxidation by IPL. In starved animals, a significant increase in net release of BCAA and tyrosine by IPL was observed, while the effect on other amino acids was non-significant. We conclude that the protein-sparing phase of uncomplicated starvation is associated with decreased whole-body proteolysis, protein synthesis, branched-chain amino acid (BCAA) oxidation, and BCAA clearance. The increase in plasma BCAA levels in starved animals results in part from decreased BCAA catabolism, particularly in heart and skeletal muscles, and from a net release of BCAA by the hepatic tissue.  相似文献   

10.
5 alpha-Cholest-8(14)-en-3 beta-yl-15-one oleate (15-ketosteryl oleate), the oleate ester of a compound with the capacity to lower serum cholesterol, was effectively hydrolyzed by partially purified porcine pancreatic cholesterol esterase with an apparent Km of 0.28 +/- 0.01 mM and a Vmax of 0.62 +/- 0.01 mumol/min per mg protein compared to an apparent Km of 0.19 +/- 0.02 mM and a Vmax of 0.37 +/- 0.02 mumol/min per mg protein for cholesteryl oleate. The 15-ketosteryl oleate was also hydrolyzed by highly purified rat pancreatic cholesterol esterase with an apparent Km of 0.20 +/- 0.01 mM and a Vmax of 86.7 +/- 3.0 mumol/min per mg protein compared to an apparent Km of 0.43 +/- 0.01 mM and a Vmax of 119.8 +/- 2.6 mumol/min per mg protein for cholesteryl oleate. 15-Ketosteryl oleate is, therefore, a good substrate for pancreatic cholesterol esterase from either source. The 15-ketosterol is a weak competitive inhibitor of partially purified porcine pancreatic cholesterol esterase when cholesteryl oleate is the substrate.  相似文献   

11.
Fatty acid-dependent ethanol metabolism   总被引:1,自引:0,他引:1  
Rates of ethanol oxidation by perfused livers from fasted female rats were decreased from 82 +/- 8 to 11 +/- 7 mumol/g/hr by 4-methylpyrazole, an inhibitor of alcohol dehydrogenase. The subsequent addition of fatty acids of various chain lengths in the presence of 4-methylpyrazole increased rates of ethanol uptake markedly. Palmitate (1 mM) increased rates of ethanol oxidation to 95 +/- 8 mumol/g/hr, while octanoate and oleate increased rates to 58 +/- 11 and 68 +/- 15 mumol/g/hr, respectively. Hexanoate, a short-chain fatty acid oxidized predominantly in the mitochondria, had no effect. Addition of oleate also increased the steady-state level of catalase-H2O2. Pretreatment of rats for 1.5 hours with 3-amino-1,2,4-triazole (1.0 g/kg), an inhibitor of catalase, prevented the ethanol-dependent decrease in the steady-state level of catalase-H2O2 completely. Under these conditions, aminotriazole decreased rates of ethanol oxidation by about 50% and blocked the stimulation of ethanol oxidation by fatty acids. Oleate decreased rates of aniline hydroxylation by about 50%, indicating that cytochrome P450 is not involved in the stimulation of ethanol uptake by fatty acids. Furthermore, oleate stimulated ethanol uptake in livers from ADH-negative deermice indicating that fatty acids do not simply displace 4-methylpyrazole from alcohol dehydrogenase. It is concluded that the stimulation of ethanol oxidation by fatty acids is due to increased H2O2 supplied by the peroxisomal beta-oxidation of fatty acids for the catalase-H2O2 peroxidation pathway.  相似文献   

12.
Efflux of branched chain alpha-keto acids from preloaded rat heart mitochondria was slow at low external pH. Efflux was first order, and measured rate constants, kappa efflux, were 0.104 +/- 0.005 and 0.115 +/- 0.006 min-1 for alpha-ketoisovalerate and alpha-ketoisocaproate (KIC), respectively. Efflux was stimulated significantly by branched chain alpha-keto acids and related carboxylates such as alpha-ketocaproate and alpha-ketovalerate, but not by substrates for the pyruvate transporter. KIC was the preferred substrate, and the apparent exchange K0.5 for KIC was 0.14 +/- 0.10 mM. Exchange was 7-8-fold faster than efflux, and the maximal rate of exchange at saturating concentrations of alpha-ketoisovalerate and KIC appeared to be independent of the metabolite used. It is proposed that branched chain alpha-keto acids cross the inner mitochondrial membrane on a specific transporter. Transport occurs with a proton, i.e. by proton symport, and is sensitive to inhibition by cinnamic acid derivatives.  相似文献   

13.
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   

14.
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   

15.
1. A branched-chain 2-oxo acid dehydrogenase was partially purified from ox liver mitochondria. 2. The preparation oxidized 4-methyl-2-oxopentanoate, 3-methyl-2-oxobutyrate and D- and L-3-methyl-2-oxopentanoate. The apparent Km values for the oxo acids and for thiamin pyrophosphate, CoA, NAD+ and Mg2+ were determined. 3. The oxidation of each oxo acid was inhibited by isovaleryl (3-methylbutyryl)-CoA (competitive with CoA) and by NADH (competitive with NAD+); Ki values were determined. 4. The preparation showed substrate inhibition with each 2-oxo acid. The oxidative decarboxylation of 4-methyl-2-oxo[1-14C]pentanoate was inhibited by 3-methyl-2-oxobutyrate and DL-3-methyl-2-oxopentanoate, but not by pyruvate. The Vmax. with 3-methyl-2-oxobutyrate as variable substrate was not increased by the presence of each of the other 2-oxo acids. 5. Ox heart pyruvate dehydrogenase did not oxidize these branched-chain 2-oxo acids and it was not inhibited by isovaleryl-CoA. The branched-chain 2-oxo acid dehydrogenase activity (unlike that of pyruvate dehydrogenase) was not inhibited by acetyl-CoA. 6. It is concluded that the branched-chain 2-oxo acid dehydrogenase activity is distinct from that of pyruvate dehydrogenase, and that a single complex may oxidize all three branched-chain 2-oxo acids.  相似文献   

16.
This paper investigates the relative importances of specific amino acids, and, in particular, branched chain amino acids and their keto derivatives as possible ketogenic precursors in suspensions of liver cells isolated from chick embryos. Addition of the branched chain keto acids stimulated ketogenesis. The order of potency was α-ketoisocaproic acid >α-ketoisovaleric acid >DL- α-keto-β-methyl-n-valeric acid. The relative order of effectiveness for branched chain keto acids was maintained at all comparable concentrations, and in each case maximum rates were observed with concentrations of 1–2 mM. In contrast to the stimulation of ketogenesis by their keto derivatives, branched chain amino acids were ineffective as precursors for ketogenesis. Of the other amino acids (utilised at concentrations present in chick embryo plasma) only Tyr, Lys, Phe and Arg produced significant increases in ketone body formation above the endogenous rate. At these physiological concentrations, the effectiveness of the amino acids were in the order of Tyr > Lys = Phe > Arg. The interactions between three groups of ketogenic precursor (fatty acids, amino acids and keto amino acids, all at physiological concentrations), produced rates of ketogenesis that were purely additive. These results indicate that high concentrations of hydroxybutyrate and acetoacetate found in plasma of developing chick embryos may arise from hepatic metabolism of several distinct precursors. The relative importance of each category of precursor may vary with the precise developmental status of animals.  相似文献   

17.
In thoroughbred horses, red blood cell amino acid transport activity is Na(+)-independent and controlled by three codominant genetic alleles (h, l, s), coding for high-affinity system asc1 (L-alanine apparent Km for influx at 37 degrees C congruent to 0.35 mM), low-affinity system asc2 (L-alanine Km congruent to 14 mM), and transport deficiency, respectively. The present study investigated amino acid transport mechanisms in red cells from four wild species: Przewalski's horse (Equus przewalskii), Hartmann's zebra (Zebra hartmannae), Grevy's zebra (Zebra grevyi), and onager (Equus hemonius). Red blood cell samples from different Przewalski's horses exhibited uniformly high rates of L-alanine uptake, mediated by a high-affinity asc1-type transport system. Mean apparent Km and Vmax values (+/- SE) for L-alanine influx at 37 degrees C in red cells from 10 individual animals were 0.373 +/- 0.068 mM and 2.27 +/- 0.11 mmol (L cells.h), respectively. As in thoroughbreds, the Przewalski's horse transporter interacted with dibasic as well as neutral amino acids. However, the Przewalski asc1 isoform transported L-lysine with a substantially (6.4-fold) higher apparent affinity than its thoroughbred counterpart (Km for influx 1.4 mM at 37 degrees C) and was also less prone to trans-stimulation effects. The novel high apparent affinity of the Przewalski's horse transporter for L-lysine provides additional key evidence of functional and possible structural similarities between asc and the classical Na(+)-dependent system ASC and between these systems and the Na(+)-independent dibasic amino acid transport system y+. Unlike Przewalski's horse, zebra red cells were polymorphic with respect to L-alanine transport activity, showing high-affinity or low-affinity saturable mechanisms of L-alanine uptake. Onager red cells transported this amino acid with intermediate affinity (apparent Km for influx 3.0 mM at 37 degrees C). Radiation inactivation analysis was used to estimate the target size of system asc in red cells from Przewalski's horse. The transporter's in situ apparent molecular weight was 158,000 +/- 2500 (SE).  相似文献   

18.
Maple syrup urine disease (MSUD) is an inherited neurometabolic disorder biochemically characterized by the accumulation of the branched-chain alpha-keto acids (BCKA) alpha-ketoisocaproic (KIC), alpha-keto-beta-methylvaleric (KMV) and alpha-ketoisovaleric (KIV) and their respective branched-chain alpha-amino acids in body fluids and tissues. Affected MSUD patients have predominantly neurological features, including cerebral edema and atrophy whose pathophysiology is not well established. In the present study we investigated the effects of KIC, KMV and KIV on cell morphology, cytoskeleton reorganization, actin immunocontent and on various parameters of oxidative stress, namely total antioxidant reactivity (TAR), glutathione (GSH) and nitric oxide concentrations, and on the activities of catalase (CAT), superoxide dismutase (SOD) and glutathione peroxidase (GPx) in C6 glioma cells. We initially observed that C6 cultivated cells exposed for 3 h to the BCKA (1 and 10 mM) changed their usual rounded morphology to a fusiform or process-bearing cell appearance, while 24 h exposure to these organic acids elicited massive cell death. Rhodamine-labelled phalloidin analysis revealed that these organic acids induced reorganization of the actin cytoskeleton with no modifications on total actin content. It was also observed that 3h cell exposure to low doses of all BCKA (1 mM) resulted in a marked reduction of the non-enzymatic antioxidant defenses, as determined by TAR and GSH measurements. In addition, KIC provoked a reduced activity of SOD and GPx, whereas KMV caused a diminution of SOD activity. In contrast, CAT activity was not modified by the metabolites. Furthermore, nitric oxide production was significantly increased by all BCKA. Finally, we observed that the morphological features caused by BCKA on C6 cells were prevented by the use of the antioxidants GSH (1.0 mM), alpha-tocopherol (trolox; 10 microM) and Nomega-nitro-L-arginine methyl ester (L-NAME; 500 microM). These results strongly indicate that oxidative stress might be involved in the cell morphological alterations and death, as well as in the cytoskeletal reorganization elicited by the BCKA. It is presumed that these findings are possibly implicated in the neuropathological features observed in patients affected by MSUD.  相似文献   

19.
Alpha-ketoisocaproic acid (KIC) is the product of the transamination of the indispensable amino acid leucine, which is the first step in the complete degradation of leucine. To determine the effects of intense exercise on muscle and blood levels of KIC, 7 male volunteers performed cycle exercise to exhaustion. After pedaling at an intensity of 90 W for 3 min, the load was increased by 60 W every 3 min until volitional fatigue. Muscle biopsies were obtained prior to and immediately after exercise and rapidly frozen for later determination of KIC. During exercise, blood lactate levels increased as expected, while plasma KIC levels did not change. Following exercise, plasma KIC levels rose significantly with peak values occurring 15 min after exercise and did not return to pre-exercise values until 60 min after exercise. In contrast, muscle KIC levels increased during exercise from a pre-exercise mean of 49.4 +/- 4.1 mumol X kg-1 wet wt to 78.1 +/- 6.5 mumol X kg-1 after exercise, an average increase of 48% (P less than 0.05). These data indicate that during intense exercise, leucine transamination in muscle may continue at a faster rate than the decarboxylation of KIC. In addition, plasma levels of KIC did not reflect the intracellular accumulation of KIC during exercise, suggesting a delay in the diffusion of KIC from muscle.  相似文献   

20.
The measurement of argininosuccinate lyase (ASase) and arginase, both in liver and erythrocytes, was developed by using a commercial amino acid analyzer. The method is based upon the use of two different substrates, argininosuccinate and arginine for ASase and arginase, respectively, and the measurement of only one final metabolite: ornithine. The use of ornithine as a marker of biological activity of ASase is related to the fact that in the urea cycle, the specific activity of arginase is much higher than that of ASase; thus, during in vitro determinations, arginine, which is the product of ASase, is rapidly converted to ornithine. The sensitivity of the methods is very high since we were able to detect both activities using very diluted rat liver homogenates (0.10 mg protein/ml) or few microliters of human blood. In rat liver the Vmax for ASase and arginase were respectively 0.54 and 140 mumol/h/mg protein; the apparent Km values 1.25 and 13.5 mM. In human erythrocytes the Vmax for the same enzymes were 7.2 and 170 nmol/h/mg Hb and the apparent Km values were 0.66 and 9.5 mM. In 10 healthy volunteers the specific activity of ASase and arginase determined in blood were respectively 8.60 +/- 0.46 and 124.1 +/- 14.5 nmol/h/mg Hb. The results obtained from 2 patients suffering from argininosuccinic aciduria were also reported. In these latter cases while ASase was not detectable in blood, arginase activity was at the lowest end of the confidence limits determined in healthy volunteers.  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