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.
Branched chain alpha-ketoacid dehydrogenase (EC 1.2.4.4) complex, the rate-limiting enzyme of branched chain amino acid catabolism in most tissues, is subject to regulation by covalent modification, with phosphorylation inactivating and dephosphorylation activating the complex. The enzyme complex from liver of chow-fed rats is mainly in the active form but that from liver of rats fed a low-protein diet is mainly in the inactive form. Isolated hepatocytes were used to identify factors that affect interconversion of branched chain alpha-ketoacid dehydrogenase. The enzyme present in hepatocytes of rats fed a low-protein diet appears much more responsive to regulation by covalent modification than the branched chain alpha-ketoacid dehydrogenase present in hepatocytes of normal chow-fed rats. alpha-Chloroisocaproate, a specific inhibitor of the kinase responsible for phosphorylation and inactivation of the complex, greatly stimulates oxidation of alpha-keto[1-14C]isovalerate by hepatocytes prepared from rats fed a low-protein diet but not from normal chow-fed rats. Oxidizable substrates are also much more effective inhibitors of branched chain alpha-ketoacid oxidation with hepatocytes from rats fed a low-protein diet than from normal chow-fed rats. Activity measurements with cell-free extracts suggest that changes in flux through the dehydrogenase with intact hepatocytes prepared from rats fed a low-protein diet are explained in large part by changes in the proportion of the enzyme in the active, dephosphorylated form. Regulation of liver branched chain alpha-ketoacid dehydrogenase by covalent modification functions to conserve branched chain amino acids for protein synthesis during periods of restricted dietary protein intake.  相似文献   

2.
The effects of branched-chain alpha-ketoacids on flux through and activity state of the branched-chain alpha-ketoacid dehydrogenase complex were studied in hepatocytes prepared from chow-fed, starved, and low-protein-diet-fed rats. Very low concentrations of alpha-ketoisocaproate caused a dramatic stimulation (50% activation at 20 microM) of alpha-ketoisovalerate decarboxylation in hepatocytes from low-protein-fed rats. alpha-Keto-beta-methylvalerate was also effective, but less so than alpha-ketoisocaproate. alpha-Ketoisocaproate did not stimulate alpha-ketoisovalerate decarboxylation by hepatocytes from chow-fed or starved rats. To a smaller degree, alpha-keto-beta-methylvalerate and alpha-ketoisovalerate stimulated alpha-ketoisocaproate decarboxylation by hepatocytes from low-protein-fed rats. The implied order of potency of stimulation of flux through branched-chain alpha-ketoacid dehydrogenase was alpha-ketoisocaproate greater than alpha-keto-beta-methylvalerate greater than alpha-ketoisovalerate, i.e., the same order of potency of these compounds as branched-chain alpha-ketoacid dehydrogenase kinase inhibitors. Fluoride, known to inhibit branched-chain alpha-ketoacid dehydrogenase phosphatase, largely prevented alpha-ketoisocaproate and alpha-chloroisocaproate activation of flux through the branched-chain alpha-ketoacid dehydrogenase. Assay of the branched-chain alpha-ketoacid complex in cell-free extracts of hepatocytes isolated from low-protein-diet-fed rats confirmed that alpha-ketoacids affected the activity state of the complex. Branched-chain alpha-ketoacids failed to activate flux in hepatocytes prepared from chow-fed and starved rats because essentially all of the complex was already in the dephosphorylated, active state. These findings indicate that inhibition of branched-chain alpha-ketoacid dehydrogenase kinase activity by branched-chain alpha-ketoacids is important for regulation of the activity state of hepatic branched-chain alpha-ketoacid dehydrogenase.  相似文献   

3.
Human skin fibroblasts in suspension are able to degrade [1-14C]-labeled alpha- and gamma-methyl branched chain fatty acids such as pristanic and homophytanic acid. Pristanic acid was converted to propionyl-CoA, whereas homophytanic acid was beta-oxidized to acetyl-CoA. Incubation of skin fibroblasts with [1-14C]-labeled fatty acids for longer periods produced radiolabeled carbon dioxide, presumably by further degradation of acetyl-CoA or propionyl-CoA generated by beta-oxidation. Under the same conditions similar products were produced from very long chain fatty acids, such as lignoceric acid. Inclusion of digitonin (> 10 micrograms/ml) in the incubations strongly inhibited carbon dioxide production but stimulated acetyl-CoA or propionyl-CoA production from fatty acids. ATP, Mg2+, coenzyme A, NAD+ and L-carnitine stimulated acetyl-CoA or propionyl-CoA production from [1-14C]-labeled fatty acids in skin fibroblast suspensions. Branched chain fatty acid beta-oxidation was reduced in peroxisome-deficient cells (Zellweger syndrome and infantile Refsum's disease) but they were beta-oxidized normally in cells from patients with X-linked adrenoleukodystrophy (ALD). Under the same conditions, lignoceric acid beta-oxidation was impaired in the above three peroxisomal disease states. These results provide evidence that branched chain fatty acid, as well as very long chain fatty acid, beta-oxidation occurs only in peroxisomes. As the defect in X-linked ALD is in a peroxisomal fatty acyl-CoA synthetase, which is believed to be specific for very long chain fatty acids, we postulate that different synthetases are involved in the activation of branched chain and very long chain fatty acids in peroxisomes.  相似文献   

4.
The lipoic acid bearing domain (hbLBD) of human mitochondrial branched chain alpha-ketoacid dehydrogenase (BCKD) plays important role of substrate channeling in oxidative decarboxylation of the branched chain alpha-ketoacids. Recently hbLBD has been found to follow two-step folding mechanism without detectable presence of stable or kinetic intermediates. The present study describes the conformational stability underlying the folding of this small beta-barrel domain. Thermal denaturation in presence of urea and isothermal urea denaturation titrations are used to evaluate various thermodynamic parameters defining the equilibrium unfolding. The linear extrapolation model successfully describes the two-step; native state <-->denatured state unfolding transition of hbLBD. The average temperature of maximum stability of hbLBD is estimated as 295.6 +/- 0.9 K. Cold denaturation of hbLBD is also predicted and discussed.  相似文献   

5.
Branched chain alpha-ketoacid dehydrogenase is a heteroprotein complex of mitochondria and commits the branched chain alpha-ketoacids to their catabolic fate. Inherited nuclear mutations in humans decrease the activity of this complex and result in maple syrup urine disease. Here we demonstrate the restoration of branched chain alpha-ketoacid dehydrogenase activity to fibroblasts from a child with this disorder by transfection with a cDNA for the prebranched chain acyltransferase. Prior to transfection these fibroblasts contained the prebranched chain acyltransferase gene but failed to transcribe the gene and thus lacked the protein. Regulation of the restored complex by phosphorylation mechanisms resembles that of wild-type cells. These results describe a human cell modeling system for testing engineered proteins and support the possibility of gene replacement therapy for this human disorder.  相似文献   

6.
Nucleotide sequence was determined for a 1.6-kilobase human cDNA putative for the branched chain acyltransferase protein of the branched chain alpha-ketoacid dehydrogenase complex. Translation of the sequence reveals an open reading frame encoding a 315-amino acid protein of molecular weight 35,759 followed by 560 bases of 3'-untranslated sequence. Three repeats of the polyadenylation signal hexamer ATTAAA are present prior to the polyadenylate tail. Within the open reading frame is a 10-amino acid fragment which matches exactly the amino acid sequence around the lipoate-lysine residue in bovine kidney branched chain acyltransferase, thus confirming the identity of the cDNA. Analysis of the deduced protein structure for the human branched chain acyltransferase revealed an organization into domains similar to that reported for the acyltransferase proteins of the pyruvate and alpha-ketoglutarate dehydrogenase complexes. This similarity in organization suggests that a more detailed analysis of the proteins will be required to explain the individual substrate and multienzyme complex specificity shown by these acyltransferases.  相似文献   

7.
The amount of pyruvate dehydrogenase in the active form (PDHa) was increased 1.7-fold compared with controls in heart muscle of mice 1 week after induction of obesity with a single injection of gold-thioglucose. At 4 weeks post injection, the amount of PDHa was decreased to 32% of control, a value which was observed in later stages of the obesity syndrome. In contrast, liver PDHa was increased and remained at an increased activity during the development of obesity. Despite normal post-prandial serum insulin contents, liver membrane insulin-receptor numbers were decreased 1 week after gold-thioglucose injection, and there was no change in receptor affinity. The decrease in heart PDHa in the obese animals was reversed by a single dose of 2-tetradecylglycidic acid, but this inhibitor of mitochondrial fatty acid oxidation did not affect liver PDHa in these animals. These early and diverse changes in PDHa argue for a multifactorial aetiology in the development of the whole-body insulin resistance seen in older gold-thioglucose-treated obese animals.  相似文献   

8.
Short/branched chain acyl-CoA dehydrogenase (SBCAD), isovaleryl-CoA dehydrogenase (IVD), and isobutyryl-CoA dehydrogenase (IBD) are involved in metabolism of isoleucine, leucine, and valine, respectively. These three enzymes all belong to acyl-CoA dehydrogenase (ACD) family, and catalyze the dehydrogenation of monomethyl branched-chain fatty acid (mmBCFA) thioester derivatives. In the present work, the catalytic properties of rat SBCAD, IVD, and IBD, including their substrate specificity, isomerase activity, and enzyme inhibition, were comparatively studied. Our results indicated that SBCAD has its catalytic properties relatively similar to those of straight-chain acyl-CoA dehydrogenases in terms of their isomerase activity and enzyme inhibition, while IVD and IBD are different. IVD has relatively broader substrate specificity than those of the other two enzymes in accommodating various substrate analogs. The present study increased our understanding for the metabolism of monomethyl branched-chain fatty acids (mmBCFAs) and branched-chain amino acids (BCAAs), which should also be useful for selective control of a particular reaction through the design of specific inhibitors.  相似文献   

9.
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   

10.
The mechanism of inhibition of pyruvate carboxylase, pyruvate dehydrogenase, and carbamyl phosphate synthetase induced by alpha-ketoisovalerate metabolism has been investigated in isolated rat hepatocytes incubated with lactate, pyruvate, ammonia, and ornithine as substrates. Half-maximum inhibitions of flux through each of these enzyme steps were obtained with 0.3 mM alpha-ketoisovalerate. The inhibition of pyruvate carboxylase flux by alpha-ketoisovalerate was largely reversed by oleate addition, but pyruvate dehydrogenase flux was inhibited further. Inhibition of flux through pyruvate carboxylase could be attributed mainly to the fall of its allosteric activator, acetyl-CoA, with some additional effect due to inhibition by methylmalonyl-CoA. Tissue acetyl-CoA levels decrease as a result of an inhibition of the active form of pyruvate dehydrogenase. Kinetic studies with the purified pig heart pyruvate dehydrogenase complex showed that methyl-malonyl-CoA, propionyl-CoA, and isobutyryl-CoA were inhibitory, the latter noncompetitive with CoASH with an apparent Ki of 90 microM. The observed inhibition of pyruvate dehydrogenase flux correlated with increases of the acetyl-CoA/CoASH and propionyl-CoA/CoASH ratios and isobutyryl-CoA levels, while increases of the mitochondrial NADH/NAD+ ratio explained differences between the effects of alpha-ketoisovalerate and propionate. Carbamyl phosphate synthetase I purified from rat liver was shown to be inhibited directly by methylmalonyl-CoA (apparent Ki of 5 mM). Inhibition of flux through carbamyl phosphate synthetase during alpha-ketoisovalerate metabolism could be attributed both to a direct inhibitory effect of methyl-malonyl-CoA and to a diminished activation by N-acetylglutamate. Direct effects of various acyl-CoA metabolites on these key enzymes may explain symptoms of hypoglycemia and hyperammonemia observed in patients with inherited disorders of organic acid metabolism.  相似文献   

11.
The metabolism of isoleucine to active amyl alcohol (2-methylbutanol) in yeast was examined by the use of (13)C nuclear magnetic resonance spectroscopy, combined gas chromatography-mass spectrometry, and a variety of mutants. From the identified metabolites a number of routes between isoleucine and active amyl alcohol seemed possible. All involved the initial decarboxylation of isoleucine to alpha-keto-beta-methylvalerate. The first, via branched chain alpha-ketoacid dehydrogenase to alpha-methylbutyryl-CoA, was eliminated because abolition of branched-chain alpha-ketoacid dehydrogenase in an lpd1 disruption mutant did not prevent the formation of active amyl alcohol. However, the lpd1 mutant still produced large amounts of alpha-methylbutyrate which initially seemed contradictory because it had been assumed that alpha-methylbutyrate was derived from alpha-methylbutyryl-CoA via acyl-CoA hydrolase. Subsequently it was observed that alpha-methylbutyrate arises from the non-enzymic oxidation of alpha-methylbutyraldehyde (the immediate decarboxylation product of alpha-keto-beta-methylvalerate). Mutant studies showed that one of the decarboxylases encoded by PDC1, PDC5, PDC6, YDL080c, or YDR380w must be present to allow yeast to utilize alpha-keto-beta-methylvalerate. Apparently, any one of this family of decarboxylases is sufficient to allow the catabolism of isoleucine to active amyl alcohol. This is the first demonstration of a role for the gene product of YDR380w, and it also shows that the decarboxylation steps for each alpha-keto acid in the catabolic pathways of leucine, valine, and isoleucine are accomplished in subtly different ways. In leucine catabolism, the enzyme encoded by YDL080c is solely responsible for the decarboxylation of alpha-ketoisocaproate, whereas in valine catabolism any one of the isozymes of pyruvate decarboxylase will decarboxylate alpha-ketoisovalerate.  相似文献   

12.
Branched chain alpha-ketoacid dehydrogenase (EC 1.2.4.3(4)) was solubilized and purified from bovine liver mitochondria for the first time. Decarboxylation of alpha-ketoisovalerate, alpha-keto-beta-methylvalerate, and alpha-ketoisocaproate was catalyzed by this multienzyme complex and this activity was co-purified for each substrate. Three enzymatic functions were contained in the complex including decarboxylation of the above ketoacids, transacylation of their simple acid derivatives, and reduction of NAD+ as an overall reaction. Product stoichiometry of these three reactions was 1 CO2:1 acyl-CoA:1 NADH. Activity depended upon the addition of thiamin pyrophosphate, CoASH, and NAD+ which were dissociable cofactors. Physically, two active forms of the enzyme complex were found: a 275,000-dalton unit and a 2 x 10(6)-dalton component. Both showed a characteristic flavin spectra and catalyzed all functions of the complex, implying that 10 small units aggregated into the larger unit. The soluble complex as visualized by electron microscopy had a diameter ranging from 12 to 24 nm corresponding to a molecular weight of 2 x 10(6). The size of the native membrane-bound component remains to be determined.  相似文献   

13.
The contribution of peroxisomal fatty acid beta-oxidation to ethanol metabolism was examined in deermice hepatocytes. Addition of 1 mM oleate to hepatocytes isolated from fasted alcohol dehydrogenase (ADH)-positive deermice in the presence of 4-methylpyrazole or to hepatocytes from fasted or fed ADH-negative deermice produced only a slight and statistically not significant increase in ethanol oxidation. Lactate (10 mM), which is not a peroxisomal substrate, showed a greater effect on ethanol oxidation. There was also a lack of oleate effect on the oxidation of ethanol by hepatocytes of ADH-positive deermice. Furthermore, in ADH-negative deermice, the catalase inhibitor azide (0.1 mM) did not inhibit the increase in ethanol oxidation by oleate and lactate. The rate of oleate oxidation by hepatocytes from fasted ADH-negative deermice was much lower than that of ethanol. These results indicate that in deermice hepatocytes, peroxisomal fatty acid oxidation does not play major role in ethanol metabolism.  相似文献   

14.
The availability of genetically modified mice requires the development of methods to assess heart function and metabolism in the intact beating organ. With the use of radioactive substrates and ex vivo perfusion of the mouse heart in the working mode, previous studies have documented glucose and fatty acid oxidation pathways. This study was aimed at characterizing the metabolism of other potentially important exogenous carbohydrate sources, namely, lactate and pyruvate. This was achieved by using (13)C-labeling methods. The mouse heart perfusion setup and buffer composition were optimized to reproduce conditions close to the in vivo milieu in terms of workload, cardiac functions, and substrate-hormone supply to the heart (11 mM glucose, 0.8 nM insulin, 50 microM carnitine, 1.5 mM lactate, 0.2 mM pyruvate, 5 nM epinephrine, 0.7 mM oleate, and 3% albumin). The use of three differentially (13)C-labeled carbohydrates and a (13)C-labeled long-chain fatty acid allowed the quantitative assessment of the metabolic origin and fate of tissue pyruvate as well as the relative contribution of substrates feeding acetyl-CoA (pyruvate and fatty acids) and oxaloacetate (pyruvate) for mitochondrial citrate synthesis. Beyond concurring with the notion that the mouse heart preferentially uses fatty acids for energy production (63.5 +/- 3.9%) and regulates its fuel selection according to the Randle cycle, our study reports for the first time in the mouse heart the following findings. First, exogenous lactate is the major carbohydrate contributing to pyruvate formation (42.0 +/- 2.3%). Second, lactate and pyruvate are constantly being taken up and released by the heart, supporting the concept of compartmentation of lactate and glucose metabolism. Finally, mitochondrial anaplerotic pyruvate carboxylation and citrate efflux represent 4.9 +/- 1.8 and 0.8 +/- 0.1%, respectively, of the citric acid cycle flux and are modulated by substrate supply. The described (13)C-labeling strategy combined with an experimental setup that enables continuous monitoring of physiological parameters offers a unique model to clarify the link between metabolic alterations, cardiac dysfunction, and disease development.  相似文献   

15.
Dichloroacetate has effects upon hepatic metabolism which are profoundly different from its effects on heart, skeletal muscle, and adipose tissue metabolism. With hepatocytes prepared from meal-fed rats, dichloroacetate was found to activate pyruvate dehydrogenase, to increase the utilization of lactate and pyruvate without effecting an increase in the net utilization of glucose, to increase the rate of fatty acid synthesis, and to decrease slightly [1-14C]oleate oxidation to 14CO2 without decreasing ketone body formation. With hepatocytes isolated from 48-h-starved rats, dichloroacetate was found to activate pyruvate dehydrogenase, to have no influence on net glucose utilization, to inhibit gluconeogenesis slightly with lactate as substrate, and to stimulate gluconeogenesis significantly with alanine as substrate. The stimulation of fatty acid synthesis by dichloroacetate suggests that the activity of pyruvate dehydrogenase can be rate determining for fatty acid synthesis in isolated liver cells. The minor effects of dichloroacetate on gluconeogenesis suggest that the regulation of pyruvate dehydrogenase is only of marginal importance in the control of gluconeogenesis.  相似文献   

16.
The activity of branched chain alpha-keto acid dehydrogenase in extracts of adipose tissue was elevated after homogenization of tissue segments which had been incubated in buffer containing 0.3 mM leucine. A maximum increase (4-fold) was observed in extracts of tissues incubated in buffer containing 2.5 mM leucine, alpha-Ketoisocaproate and leucine caused maximum increases which were of similar magnitude and which required the same length of incubation of the tissue segments (5 to 15 min). The effect of leucine on branched chain alpha-keto acid dehydrogenase activity was observed both in the presence and absence of insulin, which also increased the activity of the enzyme in tissue extracts. Intact adipose tissue segments oxidized [I-14C]leucine at a maximum rate approximately 4 times that of [1-(14)C]valine. The rate of valine oxidation by intact tissue segments was doubled by addition of 0.2 to 0.5 mM unlabeled leucine, but not isoleucine, to medium containing 2 mM [1-(14)C]valine. Leucine, but not valine, also stimulated the rate of oxidation of 2 mM [U-14C]isoleucine by intact tissue segments. These results suggest that branched chain alpha-keto acid dehydrogenase activity, which is thought to limit the rate of branched chain amino acid oxidation in adipose tissue, may be sensitive to changes in the concentration of leucine in rat blood.  相似文献   

17.
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   

18.
Valproate at 0.1 to 5 mM strongly inhibited oxidation of 1-(14C)-palmitate in isolated rat hepatocytes. Valproate at the same concentrations markedly decreased ketogenesis from 1 mM oleate. Valproate in a dose up to 5 mM did not significantly affect cellular concentration of ATP but lowered beta-hydroxybutyrate/acetoacetate and lactate/pyruvate ratios which paralleled its effect on ketogenesis. Moreover concomitant acetyl-CoA levels were drastically decreased by valproate. From this it may be concluded that inhibition of fatty acid oxidation by valproate results in reduced production of two carbons units and a drop of NADH/NAD+ ratio in rat hepatocyte. This suggests that valproate seriously interferes with beta-oxidation of physiological long-chain fatty acids.  相似文献   

19.
The objective of this study was to test the effect of increasing fatty acid concentrations on substrate fluxes through pathways leading to citrate synthesis and release in the heart. This was accomplished using semirecirculating work-performing rat hearts perfused with substrate mixtures mimicking the in situ milieu (5.5 mM glucose, 8 nM insulin, 1 mM lactate, 0.2 mM pyruvate, and 0.4 mM oleate-albumin) and 13C methods. Raising the fatty acid concentration from 0.4 to 1 mM with long-chain oleate or medium-chain octanoate resulted in a lowering ( approximately 20%) of cardiac output and efficiency with unaltered O2 consumption. At the metabolic level, beyond the expected effects of high fatty acid levels on the contribution of pyruvate decarboxylation (reduced >3-fold) and beta-oxidation (enhanced approximately 3-fold) to citrate synthesis, there was also a 2.4-fold lowering of anaplerotic pyruvate carboxylation. Despite the dual inhibitory effect of high fatty acids on pyruvate decarboxylation and carboxylation, tissue citrate levels were twofold higher, but citrate release rates remained unchanged at 11-14 nmol/min, representing <0.5% of citric acid cycle flux. A similar trend was observed for most metabolic parameters after oleate or octanoate addition. Together, these results emphasize a differential modulation of anaplerotic pyruvate carboxylation and citrate release in the heart by fatty acids. We interpret the lack of effects of high fatty acid concentrations on citrate release rates as suggesting that, under physiological conditions, this process is maximal, probably limited by the activity of its mitochondrial or plasma membrane transporter. Limited citrate release at high fatty acid concentrations may have important consequences for the heart's fuel metabolism and function.  相似文献   

20.
1. Adipocytes from rat epididymal fat-pads were incubated for 30 min with 5 mM-glucose and concentrations of lactate, pyruvate and amino acids typical of those found in rat plasma. 2. PDHa (active form of pyruvate dehydrogenase) activity was significantly increased after incubation of the cells with insulin (200 micro-i.u./ml), and decreased by incubation with palmitate (0.5--2 mM). 3. In the presence of insulin, palmitate did not decrease PDHa activity. 4. Dichloroacetate (1 mM) increased PDHa activity in the absence of palmitate to the same extent as did insulin. In the presence of dichloroacetate but the absence of insulin, palmitate decreased PDHa activity. In the presence of dichloroacetate and insulin, palmitate again did not decrease PDHa activity. 5. It is concluded that, in the presence of glucose, insulin has a strong protective action against inactivation of adipocyte PDHa by fatty acids.  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