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 62 毫秒
1.
Procedures are described for isolating highly purified porcine liver pyruvate and α-ketoglutarate dehydrogenase complexes. Rabbit serum stabilized these enzyme complexes in mitochondrial extracts, apparently by inhibiting lysosomal proteases. The complexes were purified by a three-step procedure involving fractionation with polyethylene glycol, pelleting through 12.5% sucrose, and a second fractionation under altered conditions with polyethylene glycol. Sedimentation equilibrium studies gave a molecular weight of 7.2 × 106 for the liver pyruvate dehydrogenase complex. Kinetic parameters are presented for the reaction catalyzed by the pyruvate dehydrogenase complex and for the regulatory reactions catalyzed by the pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase. For the overall catalytic reaction, the competitive Ki to Km ratio for NADH versus NAD+ and acetyl CoA versus CoA were 4.7 and 5.2, respectively. Near maximal stimulations of pyruvate dehydrogenase kinase by NADH and acetyl CoA were observed at NADH:NAD+ and acetyl CoA:CoA ratios of 0.15 and 0.5, respectively. The much lower ratios required for enhanced inactivation of the complex by pyruvate dehydrogenase kinase than for product inhibition indicate that the level of activity of the regulatory enzyme is not directly determined by the relative affinity of substrates and products of catalytic sites in the pyruvate dehydrogenase complex. In the pyruvate dehydrogenase kinase reaction, K+ and NH+4 decreased the Km for ATP and the competitive inhibition constants for ADP and (β,γ-methylene)adenosine triphosphate. Thiamine pyrophosphate strongly inhibited kinase activity. A high concentration of ADP did not alter the degree of inhibition by thiamine pyrophosphate nor did it increase the concentration of thiamine pyrophosphate required for half-maximal inhibition.  相似文献   

2.
The pyruvate dehydrogenase complex has been purified 76-fold, to a specific activity of 0.6 μmoles per minute per milligram protein, beginning with isolated pea (Pisum sativum L. var Little Marvel) chloroplasts. Purification was accomplished by rate zonal sedimentation, polyethyleneglycol precipitation, and ethyl-agarose affinity chromatography. Characterization of the substrates as pyruvate, NAD+, and coenzyme-A and the products as NADH, CO2, and acetyl-CoA, in a 1:1:1 stoichiometry unequivocally established that activity was the result of the pyruvate dehydrogenase complex. Immunochemical analysis demonstrated significant differences in structure and organization between the chloroplast pyruvate dehydrogenase complex and the more thoroughly characterized mitochondrial complex. Chloroplast complex has a higher magnesium requirement and a more alkaline pH optimum than mitochondrial complex, and these properties are consistent with light-mediated regulation in vivo. The chloroplast pyruvate dehydrogenase complex is not, however, regulated by ATP-dependent inactivation. The properties and subcellular localization of the chloroplast pyruvate dehydrogenase complex are consistent with its role of providing acetyl-CoA and NADH for fatty acid synthesis.  相似文献   

3.
A substance capable of stimulating the activities of pyruvate dehydrogenase and low Km cyclic AMP phosphodiesterase was prepared from H4-II-EC3′ hepatoma cells by acid extraction and partially purified by molecular exclusion chromatography. The material thus prepared by gel chromatography was found to stimulate the activities of these enzymes in a concentration-dependent manner. The amount or activity of the pyruvate dehydrogenase stimulating factor was increased in cells which had been treated with physiological concentrations of insulin (0.2 mU/ml). Increasing the concentration of insulin increased the amount or activity of the factor generated. High concentrations of insulin did not cause a reversal of the effects of insulin. The stimulation of pyruvate dehydrogenase activity by the factor was eliminated when sodium fluoride (75 mm) was present in the enzyme assay, implying that activation was mediated by the pyruvate dehydrogenase phosphatase. The enzyme-stimulating factor isolated from hepatoma cells shares a number of important characteristics with the putative second messenger of insulin prepared from other cell types: (1) it is heat and acid stable, (2) it has a similar apparent molecular weight, (3) it is generated in an insulin-dependent manner, (4) it stimulates the activity of pyruvate dehydrogenase by a fluoride-sensitive mechanism, and (5) it elutes from the anion-exchange resin AG 1-X8 at an ionic strength of 0.4 m. These findings suggest that the stimulator of pyruvate dehydrogenase and of low Km cyclic AMP phosphodiesterase isolated from hepatoma cells has chemical properties identical with those of the putative second messenger of insulin action isolated from a number of other insulin-sensitive tissues.  相似文献   

4.
Binding of fructose-6-P and Pi to rabbit liver fructose bisphosphatase has been analyzed in terms of four negatively cooperative binding sites per enzyme tetramer. The association of fructose-6-P occurs in the absence of divalent metal ion, although the extent of binding is increased in the order Mg2+ < Zn2+ < Mn2+. The binding of Pi shows an absolute requirement for divalent metal ion with Mn2+ being more effective than Mg2+. The interaction of the enzyme with the substrate analog, (α + β) methyl-d-fructofuranoside-1,6-P2 in the presence of Mn2+ closely resembles that found for fructose-1,6-P2 in the absence of Mn2+, although the measured constants are on average an order of magnitude smaller. Combination experiments with the three ligands show that the binding follows an identical ordered sequence, i.e., the tighter sites are initially occupied regardless of the ligand's identity. The binding of Pi or fructose-6-P is not altered by the presence of the other. Comparison of binding constant with Ki values obtained from steady-state assays permits identification of the catalytic sites expressed in the latter. The association of Mn2+ at the catalytic site can be induced by fructose-6-P or the substrate analog suggesting that a 1-phosphoryl group enhances but is not necessary for Mn2+ binding at this site. The binding of AMP is decreased in the presence of substrate analog relative to fructose-1,6-P2, suggesting that the 2-hydroxyl serves as a “molecular signal.” From the single and combined binding experiments, a calculation of the equilibrium constant for the overall hydrolysis reaction on the enzyme surface in the presence of Mn2+ has been carried out and an estimate made for the Mg2+ case.  相似文献   

5.
We have shown that the active form of the pyruvate dehydrogenase (PDHa) component exhibits at least a 9-fold greater affinity for sites on the dihydrolipoyl transacetylase core of the pyruvate dehydrogenase complex than does the inactive (phosphorylated) form of pyruvate dehydrogenase (PDHb). Consistent with a higher rate of dissociation for PDHb than for PDHa, free PDHa rapidly replaces PDHb whereas, even at high levels, free PDHb only slowly replaces PDHa. Dissociation of newly formed PDHb, during phosphorylation by the immobile PDHa kinase, leads to an increased association of free PDHa as observed by protection against inactivation of the complex, even though PDHa kinase activity is increased.  相似文献   

6.
Most bacterial pyruvate dehydrogenase complexes from either Gram-positive or Gram-negative bacteria have E1 components with an α2 homodimeric quaternary structure. In a sequel to our previous publications, we present the first NMR study on the flexible regions of the E1 component from Escherichia coli and its biological relevance. We report sequence-specific NMR assignments for 6 residues in the N-terminal 1–55 region and for a glycine in each of the two mobile active center loops of the E1 component, a 200-kDa homodimer. This was accomplished by using site-specific substitutions and appropriate labeling patterns along with a peptide with the sequence corresponding to the N-terminal 1–35 amino acids of the E1 component. To study the functions of these mobile regions, we also examined the spectra in the presence of (a) a reaction intermediate analog known to affect the mobility of the active center loops, (b) an E2 component construct consisting of a lipoyl domain and peripheral subunit binding domain, and (c) a peptide corresponding to the amino acid sequence of the E2 peripheral subunit binding domain. Deductions from the NMR studies are in excellent agreement with our functional finding, providing a clear indication that the N-terminal region of the E1 interacts with the E2 peripheral subunit binding domain and that this interaction precedes reductive acetylation. The results provide the first structural support to the notion that the N-terminal region of the E1 component of this entire class of bacterial pyruvate dehydrogenase complexes is responsible for binding the E2 component.  相似文献   

7.
Shiga toxin type 2 (Stx2a) is clinically most closely associated with enterohemorrhagic E. coli O157:H7-mediated hemorrhagic colitis that sometimes progresses to hemolytic-uremic syndrome. The ability to express the toxin has been acquired by other Escherichia coli strains, and outbreaks of food poisoning have caused significant mortality rates as, for example, in the 2011 outbreak in northern Germany. Stx2a, an AB5 toxin, gains entry into human cells via the glycosphingolipid receptor Gb3. We have determined the first crystal structure of a disaccharide analog of Gb3 bound to the B5 pentamer of Stx2a holotoxin. In this Gb3 analog, α-GalNAc replaces the terminal α-Gal residue. This co-crystal structure confirms previous inferences that two of the primary binding sites identified in the B5 pentamer of Stx1 are also functional in Stx2a. This knowledge provides a rationale for the synthesis and evaluation of heterobifunctional antagonists for E. coli toxins that target Stx2a. Incorporation of GalNAc Gb3 trisaccharide in a heterobifunctional ligand with an attached pyruvate acetal, a ligand for human amyloid P component, and conjugation to poly[acrylamide-co-(3-azidopropylmethacrylamide)] produced a polymer that neutralized Stx2a in a mouse model of Shigatoxemia.  相似文献   

8.
The influence of oxygen and temperature on the inactivation of pyruvate, Pi dikinase and NADP-malate dehydrogenase was studied in Zea mays. O2 was required for inactivation of both pyruvate, Pi dikinase and NADP-malate dehydrogenase in the dark in vivo. The rate of inactivation under 2% O2 was only slightly lower than that at 21% O2. The in vitro inactivation of pyruvate, Pi dikinase, while dependent on adenine nucleotides (ADP + ATP), did not require O2.

The postillumination inactivation of pyruvate, Pi dikinase in leaves was strongly dependent on temperature. As temperature was decreased in the dark, there was a lag period of increasing length (e.g. at 17°C there was a lag of about 25 minutes) before inactivation proceeded. Following the lag period, the rate of inactivation decreased with decreasing temperature. The half-time for dark inactivation was about 7 minutes at 32°C and 45 minutes at 17°C. The inactivation of pyruvate, Pi dikinase in vitro following extraction from illuminated leaves was also strongly dependent on temperature, but occurred without a lag period. In contrast, NADP-malate dehydrogenase was rapidly inactivated in leaves (half-time of approximately 3 minutes) during the postillumination period without a lag, and there was little effect of temperature between 10 and 32°C. The results are discussed in relation to known differences in the mechanism of activation/inactivation of the two enzymes.

  相似文献   

9.
The dihydrolipoyl transacetylase core component of the bovine kidney and heart pyruvate dehydrogenase complexes were covalently attached through the lipoyl moiety to Sepharose by the thiol-crosslinking reagent, N, N′-p-phenylenedimaleimide.In one approach, the N, N′-p-phenylenedimaleimide was allowed to react with glutathione which was in turn linked by its N-terminal to Sepharose CL-6B. In addition, we found the N, N′-p-phenylenedimaleimide would react directly with Sepharose CL-6B (at undetermined sites) and could be used as the sole bridge in forming a stable linkage of the transacetylase core to Sepharose. With the latter approach the extent of multiple-linkage of the 60-subunit core could more easily be controlled. This should be a generally useful approach for linking proteins with reactive surface thiol residues.Insolubilization of the core of the pyruvate dehydrogenase complex by these methods did not appear to significantly alter the binding of other protein components of the complex, but the catalytic activities of the complex requiring the lipoyl moiety were appreciably altered. Procedures for coupling the transacetylase core to various derivatives of phenylenedimaleimide-Sepharose and techniques described for studying the protein products should be useful in preparation of specialized matrices for both protein purification and the study of protein-protein interactions.  相似文献   

10.
The overall reaction catalyzed by the pyruvate dehydrogenase complex from rat epididymal fat tissue is inhibited by glyoxylate at concentrations greater than 10 μm. The inhibition is competitive with respect to pyruvate; Ki was found to be 80 μm. Qualitatively similar results were observed using pyruvate dehydrogenase from rat liver, kidney, and heart. Glyoxylate also inhibits the pyruvate dehydrogenase phosphate phosphatase from rat epididymal fat, with the inhibition being readily detectable using 50 μm glyoxylate. These effects of glyoxylate are largely reversed by millimolar concentrations of thiols (especially cysteine) because such compounds form relatively stable adducts with glyoxylate. Presumably these inhibitions by low levels of glyoxylate had not been previously observed, because others have used high concentrations of thiols in pyruvate dehydrogenase assays. Since the inhibitory effects are seen with suspected physiological concentrations, it seems likely that glyoxylate partially controls the activity of pyruvate dehydrogenase in vivo.  相似文献   

11.
1. The effects of phenylpyruvate, a metabolite produced in phenylketonuria, on the pyruvate dehydrogenase-complex activity were investigated in rat brain mitochondria. 2. Pyruvate dehydrogenase activity was measured by two methods, one measuring the release of 14CO2 from [1-14C]pyruvate and the other measuring the acetyl-CoA formed by means of the coupling enzyme, pigeon liver arylamine acetyltransferase (EC 2.3.1.5). In neither case was there significant inhibition of the pyruvate dehydrogenase complex by phenylpyruvate at concentrations below 2mm. 3. However, phenylpyruvate acted as a classical competitive inhibitor of the coupling enzyme arylamine acetyltransferase, with a Ki of 100μm. 4. It was concluded that the inhibition of pyruvate dehydrogenase by phenylpyruvate is unlikely to be a primary enzyme defect in phenylketonuria.  相似文献   

12.
The rare earth gadolinium forms a binary enzyme-metal complex with muscle pyruvate kinase which enhances the water proton relaxation rate (?b = 12 ± 2). Analysis of a Scatchard plot of the binding data indicates 3.7 ± 0.5 gadolinium binding sites with Kd = 26 ± 10 μM per protein of 237,000 daltons. The transition metal ion, manganese, is displaced from the enzyme by the rare earths, gadolinium, neodymium, thulium, and lanthanum as well as the alkaline earths, magnesium and calcium suggesting all of these metal ions bind to the same site on the protein. Upon addition of ATP to a solution of gadolinium and enzyme a decrease in enhancement is observed which is consistent with the formation of a metal bridge complex. Because of the low dissociation constant for the Gd-ATP complex (0.1 μm) it is possible to directly measure the dissociation of the Gd-ATP complex from the ternary enzyme-Gd-ATP complex, K2 = 13 μM ± 4 μM. However, a ternary complex of phospho-enolpyruvate-Gd-enzyme is not detected by water proton relaxation rate enhancement measurements which leads to speculation that the ionic radius of gadolinium (0.94 Å) is so large that it results in a distortion of the phosphoenolypyruvate binding site on pyruvate kinase thus preventing phosphoenolpyruvate binding.  相似文献   

13.
Despite extensive research, the regulation of mitochondrial function is still not understood completely. Ample evidence shows that cytosolic Ca2+ has a strategic task in co-ordinating the cellular work load and the regeneration of ATP by mitochondria. Currently, the paradigmatic view is that Cacyt2+ taken up by the Ca2+ uniporter activates the matrix enzymes pyruvate dehydrogenase, α-ketoglutarate dehydrogenase and isocitrate dehydrogenase. However, we have recently found that Ca2+ regulates the glutamate-dependent state 3 respiration by the supply of glutamate to mitochondria via aralar, a mitochondrial glutamate/aspartate carrier. Since this activation is not affected by ruthenium red, glutamate transport into mitochondria is controlled exclusively by extramitochondrial Ca2+. Therefore, this discovery shows that besides intramitochondrial also extramitochondrial Ca2+ regulates oxidative phosphorylation. This new mechanism acts as a mitochondrial “gas pedal”, supplying the OXPHOS with substrate on demand. These results are in line with recent findings of Satrustegui and Palmieri showing that aralar as part of the malate–aspartate shuttle is involved in the Ca2+-dependent transport of reducing hydrogen equivalents (from NADH) into mitochondria. This review summarises results and evidence as well as hypothetical interpretations of data supporting the view that at the surface of mitochondria different regulatory Ca2+-binding sites exist and can contribute to cellular energy homeostasis. Moreover, on the basis of our own data, we propose that these surface Ca2+-binding sites may act as targets for neurotoxic proteins such as mutated huntingtin and others. The binding of these proteins to Ca2+-binding sites can impair the regulation by Ca2+, causing energetic depression and neurodegeneration.  相似文献   

14.
A subcomplex consisting of dihydrolipoyl transacetylase and dihydrolipoyl dehydrogenase, two of the three enzymes comprising the Escherichia coli pyruvate dehydrogenase complex, has been crystallized. X-ray diffraction data establish that the space group is P213 with unit cell dimension a=211 .5A?. The unit cell contains four molecules of the subcomplex, each possessing 3-fold crystallographic and molecular symmetry. This finding, together with biochemical and electron microscopic data reported elsewhere, establish unequivocally that dihydrolipoyl transacetylase, the core enzyme of the pyruvate dehydrogenase complex, consists of 24 identical subunits with octahedral (432) symmetry. In the case presented here, the 432 symmetry of the transacetylase is reduced to 3-fold symmetry in the subcomplex by the addition of dihydrolipoyl dehydrogenase subunits. Crystal density measurements indicate that the dihydrolipoyl transacetylase present in these crystals is considerably smaller than the core mass generally reported for intact transacetylase. The implications of these findings are discussed with respect to the subunit stoichiometry and structure of the E. coli pyruvate dehydrogenase complex.  相似文献   

15.
35Cl nmr relaxation rate measurements have been used to study anion-binding sites in pig heart lactate dehydrogenase. These studies reveal two types of sites, one is intimately associated with the active site, the other is not. The nonactive site has been ascribed to a subunit site in analogy with crystallographic results from the dogfish M4 enzyme. The binding of either the reduced or the oxidized form of NAD results in an increase in the 35Cl nmr relaxation rate by a factor of 1.8–2. The enhanced nmr relaxation rate of the binary lactate dehydrogenase-NAD complex is reduced on binding of the substrate inhibitor molecules oxamate or oxalate to a value less than that exhibited by lactate dehydrogenase alone. The enhancement of the nmr relaxation rate is attributed to a decrease in the dissociation constant of Cl for the enzyme. The Kp values for Cl binding to the active center site of lactate dehydrogenase is 0.85 m and for lactate dehydrogenase-NADH is 0.25 m. The ratio of these constants, 3.4, agrees well with the measured enhancement value 3.7. The effect of coenzyme analogs on the 35Cl nmr relaxation rate has been examined. 3-Acetylpyridine NAD produces an enhancement of 4.3, thionicotinamide NAD of 2.3, whereas 3-pyridinealdehyde, adenosinediphosphoribose, and adenosine diphosphate do not affect the nmr relaxation state of Cl bound to lactate dehydrogenase.  相似文献   

16.
In vitro, the pyruvate dehydrogenase complex is sensitive to product inhibition by NADH and acetyl-coenzyme A (CoA). Based upon Km and Ki relationships, it was suggested that NADH can play a primary role in control of pyruvate dehydrogenase complex activity in vivo (JA Miernyk, DD Randall [1987] Plant Physiol 83:306-310). We have now extended the in vitro studies of product inhibition by assaying pyruvate dehydrogenase complex activity in situ, using purified intact mitochondria from green pea (Pisum sativum) seedlings. In situ activity of the pyruvate dehydrogenase complex is inhibited when mitochondria are incubated with malonate. In some instances, isolated mitochondria show an apparent lack of coupling during pyruvate oxidation. The inhibition by malonate, and the apparent lack of coupling, can both be explained by an accumulation of acetyl-CoA. Inhibition could be alleviated by addition of oxalacetate, high levels of malate, or l-carnitine. The CoA pool in nonrespiring mitochondria was approximately 150 micromolar, but doubled during pyruvate oxidation, when 60 to 95% of the total was in the form of acetyl-CoA. Our results indicate that in situ activity of the mitochondrial pyruvate dehydrogenase complex can be controlled in part by acetyl-CoA product inhibition.  相似文献   

17.
18.
A vinyl phosphonate analog of adenosine 5′-phosphate (AMP) was synthesized in which the CH2OP system of AMP is replaced by CHCHP. The Vmax values of this analog relative to AMP were 0.7% with rabbit muscle AMP aminohydrolase, 13.4% with rabbit muscle AMP kinase, and 6.6% with pig muscle AMP kinase. The vinyl analog of ADP produced by the kinases was a substrate of rabbit muscle pyruvate kinase. These results, together with substrate specificity properties at the AMP sites of the enzymes indicate that the C(4′)-C(5′)-O(5′)-P system of AMP is of trans character during conversion of AMP to ADP by pig or rabbit AMP kinase.  相似文献   

19.
A photolabile analog of N-1-naphthylphthalamic acid (NPA), 5′-azido-N-1-naphthylphthalamic acid (Az-NPA), has been synthesized and characterized. This potential photoaffinity label for the plasma membrane NPA binding protein competes with [3H]NPA for binding sites on Curcurbita pepo L. (zucchini) hypocotyl cell membranes with K0.5 = 2.8 × 10−7 molar. The K0.5 for NPA under these conditions is 2 × 10−8 molar, indicating that the affinity of Az-NPA for the membranes is only 14-fold lower than NPA. While the binding of Az-NPA to NPA binding sites is reversible in the dark, exposure of the Az-NPA treated membranes to light results in a 30% loss in [3H]NPA binding ability. Pretreatment of the membranes with NPA protects the membranes against photodestruction of [3H]NPA binding sites by Az-NPA supporting the conclusion that Az-NPA destroys these sites by specific covalent attachment.  相似文献   

20.
Dichloroacetate (2 mm) stimulated the conversion of [1-14C]lactate to glucose in hepatocytes from fed rats. In hepatocytes from rats starved for 24 h, where the mitochondrial NADHNAD+ ratio is elevated, dichloroacetate inhibited the conversion of [1-14C]lactate to glucose. Dichloroacetate stimulated 14CO2 production from [1-14C]lactate in both cases. It also completely activated pyruvate dehydrogenase and increased flux through the enzyme. The addition of β-hydroxybutyrate, which elevates the intramitochondrial NADHNAD+ ratio, changed the metabolism of [1-14C]lactate in hepatocytes from fed rats to a pattern similar to that seen in hepatocytes from starved rats. Thus, the effect of dichloroacetate on labeled glucose synthesis from lactate appears to depend on the mitochondrial oxidation-reduction state of the hepatocytes. Glucagon (10 nm) stimulated labeled glucose synthesis from lactate or alanine in hepatocytes from both fed and starved rats and in the absence or presence of dichloroacetate. The hormone had no effect on pyruvate dehydrogenase activity whether or not the enzyme had been activated by dichloroacetate. Thus, it appears that pyruvate dehydrogenase is not involved in the hormonal regulation of gluconeogenesis. Glucagon inhibited the incorporation of 10 mm [1-14C]pyruvate into glucose in hepatocytes from starved rats. This inhibition has been attributed to an inhibition of pyruvate dehydrogenase by the hormone (Zahlten et al., 1973, Proc. Nat. Acad. Sci. USA70, 3213–3218). However, dichloroacetate did not prevent the inhibition of glucose synthesis. Nor did glucagon alter the activity of pyruvate dehydrogenase in homogenates of cells that had been incubated with 10 mm pyruvate in the absence or presence of dichloroacetate. Thus, the inhibition by glucagon of pyruvate gluconeogenesis does not appear to be due to an inhibition of pyruvate dehydrogenase.  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