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.
In this study changes in alternative pathways of glucose metabolism are examined in the rat lens using radiolabelled glucose in a 1 hr in vitro incubation of 50 mM or 10 mM glucose with or without 0.1 mM phenazine methosulphate (PMS). PMS which reoxidizes NADPH ensures that the pentose phosphate pathway (PPP) is not limited by the supply of NADP+. The data shows that maximal activation of the PPP (with PMS) is 40% greater at high glucose concentrations than normal glucose. This difference in maximal stimulation may be explained by the increase glucose uptake in the hyperglycaemic incubation. In the high-glucose incubation with PMS, hexokinase activity and the glucose 6-phosphate pool is not limiting for the PPP. Under these conditions, PMS alter the NAD+/NADH and NADP+/NADPH ratio. The change in the redox state alters the flux through the polyol pathway, the glycerol 3-phosphate shuttle and the glycolytic control sites, glyceraldehyde 3-phosphate, pyruvate and lactate dehydrogenases. These results are discussed in relation to hyperglycaemia-induced oxidative stress.  相似文献   

2.
Transhydrogenase is a proton pump. It has separate binding sites for NAD+/NADH (on domain I of the protein) and for NADP+/NADPH (on domain III). Purified, detergent-dispersed transhydrogenase from Escherichia coli catalyses the reduction of the NAD+ analogue, acetylpyridine adenine dinucleotide (AcPdAD+), by NADH at a slow rate in the absence of added NADP+ or NADPH. Although it is slow, this reaction is surprising, since transhydrogenase is generally thought to catalyse hydride transfer between NAD(H)--or its analogues and NADP(H)--or its analogues, by a ternary complex mechanism. It is shown that hydride transfer occurs between the 4A position on the nicotinamide ring of NADH and the 4A position of AcPdAD+. On the basis of the known stereospecificity of the enzyme, this eliminates the possibilities of transhydrogenation(a) from NADH in domain I to AcPdAD+ wrongly located in domain III; and (b) from NADH wrongly located in domain III to AcPdAD+ in domain I. In the presence of low concentrations of added NADP+ or NADPH, detergent-dispersed E. coli transhydrogenase catalyses the very rapid reduction of AcPdAD+ by NADH. This reaction is cyclic; it takes place via the alternate oxidation of NADPH by AcPdAD+ and the reduction of NADP+ by NADH, while the NADPH and NADP+ remain tightly bound to the enzyme. In the present work, it is shown that the rate of the cyclic reaction and the rate of reduction of AcPdAD+ by NADH in the absence of added NADP+/NADPH, have similar dependences on pH and on MgSO4 concentration and that they have a similar kinetic character. It is therefore suggested that the reduction of AcPdAD+ by NADH is actually a cyclic reaction operating, either with tightly bound NADP+/NADPH on a small fraction (< 5%) of the enzyme, or with NAD+/NADH (or AcPdAD+/AcPdADH) unnaturally occluded within the domain III site. Transhydrogenase associated with membrane vesicles (chromatophores) of Rhodospirillum rubrum also catalyses the reduction of AcPdAD+ by NADH in the absence of added NADP+/NADPH. When the chromatophores were stripped of transhydrogenase domain I, that reaction was lost in parallel with 'normal reverse' transhydrogenation (e.g., the reduction of AcPdAD+ by NADPH). The two reactions were fully recovered upon reconstitution with recombinant domain I protein. However, after repeated washing of the domain I-depleted chromatophores, reverse transhydrogenation activity (when assayed in the presence of domain I) was retained, whereas the reduction of AcPdAD+ by NADH declined in activity. Addition of low concentrations of NADP+ or NADPH always supported the same high rate of the NADH-->AcPdAD+ reaction independently of how often the membranes were washed. It is concluded that, as with the purified E. coli enzyme, the reduction of AcPdAD+ by NADH in chromatophores is a cyclic reaction involving nucleotides that are tightly bound in the domain III site of transhydrogenase. However, in the case of R. rubrum membranes it can be shown with some certainty that the bound nucleotides are NADP+ or NADPH. The data are thus adequately explained without recourse to suggestions of multiple nucleotide-binding sites on transhydrogenase.  相似文献   

3.
Diabetes increases the incidence of cardiovascular disease as well as the complications of myocardial infarction. Studies using animal models of diabetes have demonstrated that the metabolic alterations occurring at the myocyte level may contribute to the severity of ischemic injury in diabetic hearts. Of the several mechanisms being investigated to understand the pathogenesis of diabetic complications, the increased metabolism of glucose via the polyol pathway has received considerable attention. Deviant metabolic regulation due to increased flux through aldose reductase in diabetic hearts may influence the ability of the myocardium to withstand ischemia insult. To determine if aldose reductase inhibition improves tolerance to ischemia, hearts from acute type I diabetic and nondiabetic control rats were isolated and retrograde perfused. Each group was exposed to 1 micromol/l zopolrestat, a specific inhibitor of aldose reductase, for 10 min, followed by 20 min of global ischemia and 60 min of reperfusion in the absence of zopolrestat. Zopolrestat reduced sorbitol levels before ischemia in diabetic hearts. The cytosolic redox state (NADH/NAD+), as measured by lactate-to-pyruvate ratios, was significantly lowered under baseline, ischemic, and reperfusion conditions in diabetic hearts perfused with zopolrestat. In these diabetic hearts, ATP was significantly higher in zopolrestat hearts during ischemia, as were phosphocreatine and left ventricular-developed pressure on reperfusion. Zopolrestat provided similar metabolic and functional benefits in nondiabetic hearts. Creatine kinase release was reduced by approximately 50% in both nondiabetic and diabetic hearts treated with zopolrestat. These data indicate that inhibition of aldose reductase activity preserves high-energy phosphates, maintains a lower cytosolic NADH/NAD+ ratio, and markedly protects both diabetic and nondiabetic hearts during ischemia and reperfusion.  相似文献   

4.
Nicotinamide nucleotide transhydrogenase catalyzes the reversible reduction of NADP+ by NADH and a concomitant proton translocation. It was demonstrated (Glavas, N.A. and Bragg, P.D. (1995) Biochim. Biophys. Acta 1231, 297-303) that the Escherichia coli transhydrogenase also catalyzed a reduction of the NAD-analogue 3-acetylpyridine-NAD+ (AcPyAD+) by NADH at low pH and in the absence of (added) NADP(H) and high salt concentrations The mechanism of this reaction has as yet not been explained. In the present study, the E. coli transhydrogenase was purified by affinity chromatography through the NADP(H)-site, rendering the pure enzyme free of NADP(H). Using this preparation it was confirmed that the enzyme readily catalyzes the above reaction. Inhibitors specific for the NADP(H)-site, e.g., palmitoyl-Coenzyme A and adenosine-2'-monophosphate-5'-diphosphoribose, strongly inhibited the reduction of AcPyAD+ by NADH, whereas an inhibitor of the NAD(H)-site, adenosine 5'-diphosphoribose, was less inhibitory. This suggests that a lack of metal ions or other ions at low pH induces an unspecific interaction of the NADP(H)-site with AcPyAD+ or NADH, presumably NADH, producing a cyclic reduction of AcPyAD+ by NADH via NAD(H) bound in the NADP(H) site. A stimulation of reduction of AcPyAD+ by NADPH by Mg2+ present during reconstitution of transhydrogenase in phospholipid vesicles was observed, but it is presently unclear whether this effect is related to that seen with the detergent-dispersed enzyme.  相似文献   

5.
Flavin reductase catalyzes the reduction of free flavins by NAD(P)H. As isolated, Escherichia coli flavin reductase does not contain any flavin prosthetic group but accommodates both the reduced pyridine nucleotide and the flavin substrate in a ternary complex prior to oxidoreduction. The reduction of riboflavin by NADPH catalyzed by flavin reductase has been studied by static and rapid kinetics absorption spectroscopies. Static absorption spectroscopy experiments revealed that, in the presence of riboflavin and reduced pyridine nucleotide, flavin reductase stabilizes, although to a small extent, a charge-transfer complex of NADP+ and reduced riboflavin. In addition, reduction of riboflavin was found to be essentially irreversible. Rapid kinetics absorption spectroscopy studies demonstrated the occurrence of two intermediates with long-wavelength absorption during the catalytic cycle. Such intermediate species exhibit spectroscopic properties similar to those of charge-transfer complexes of oxidized flavin and NAD(P)H, and reduced flavin and NAD(P)+, respectively, which have been identified as intermediates during the reaction of flavoenzymes of the ferredoxin-NADP+ reductase family. Thus, a minimal kinetic scheme for the reaction of flavin reductase with NADPH and riboflavin can be proposed. After formation of the Michaelis complex of flavin reductase with NADPH and riboflavin, a first intermediate, identified as a charge-transfer complex of NADPH and riboflavin, is formed. It is followed by a second charge-transfer intermediate of enzyme-bound NADP+ and reduced riboflavin. The latter decays, yielding the Michaelis complex of flavin reductase with NADP+ and reduced riboflavin, which then dissociates to complete the reaction. These results support the initial hypothesis of a structural similarity between flavin reductase and the enzymes of the ferredoxin-NADP+ reductase family and extend it at a functional level.  相似文献   

6.
There is strong evidence to show that diabetes is associated with increased oxidative stress. However, the source of this oxidative stress remains unclear. Using transgenic mice that overexpress aldose reductase (AR) in their lenses, we found that the flux of glucose through the polyol pathway is the major cause of hyperglycemic oxidative stress in this tissue. The substantial decrease in the level of reduced glutathione (GSH) with concomitant rise in the level of lipid peroxidation product malondialdehyde (MDA) in the lens of transgenic mice, but not in the nontransgenic mice, suggests that glucose autoxidation and nonenzymatic glycation do not contribute significantly to oxidative stress in diabetic lenses. AR reduction of glucose to sorbitol probably contributes to oxidative stress by depleting its cofactor NADPH, which is also required for the regeneration of GSH. Sorbitol dehydrogenase, the second enzyme in the polyol pathway that converts sorbitol to fructose, also contributes to oxidative stress, most likely because depletion of its cofactor NAD+ leads to more glucose being channeled through the polyol pathway. Despite a more than 100% increase of MDA, oxidative stress plays only a minor role in the development of cataract in this acute diabetic cataract model. However, chronic oxidative stress generated by the polyol pathway is likely to be an important contributing factor in the slow-developing diabetic cataract as well as in the development of other diabetic complications.--Lee, A. Y. W., Chung, S. S. M. Contributions of polyol pathway to oxidative stress in diabetic cataract. FASEB J. 13, 23-30 (1999)  相似文献   

7.
During batch growth of Lactococcus lactis subsp. lactis NCDO 2118 on various sugars, the shift from homolactic to mixed-acid metabolism was directly dependent on the sugar consumption rate. This orientation of pyruvate metabolism was related to the flux-controlling activity of glyceraldehyde-3-phosphate dehydrogenase under conditions of high glycolytic flux on glucose due to the NADH/NAD+ ratio. The flux limitation at the level of glyceraldehyde-3-phosphate dehydrogenase led to an increase in the pool concentrations of both glyceraldehyde-3-phosphate and dihydroxyacetone-phosphate and inhibition of pyruvate formate lyase activity. Under such conditions, metabolism was homolactic. Lactose and to a lesser extent galactose supported less rapid growth, with a diminished flux through glycolysis, and a lower NADH/NAD+ ratio. Under such conditions, the major pathway bottleneck was most probably at the level of sugar transport rather than glyceraldehyde-3-phosphate dehydrogenase. Consequently, the pool concentrations of phosphorylated glycolytic intermediates upstream of glyceraldehyde-3-phosphate dehydrogenase decreased. However, the intracellular concentration of fructose-1,6-bisphosphate remained sufficiently high to ensure full activation of lactate dehydrogenase and had no in vivo role in controlling pyruvate metabolism, contrary to the generally accepted opinion. Regulation of pyruvate formate lyase activity by triose phosphates was relaxed, and mixed-acid fermentation occurred (no significant production of lactate on lactose) due mostly to the strong inhibition of lactate dehydrogenase by the in vivo NADH/NAD+ ratio.  相似文献   

8.
It has been shown that treatment of bovine mitochondrial complex I (NADH-ubiquinone oxidoreductase) with NADH or NADPH, but not with NAD or NADP, increases the susceptibility of a number of subunits to tryptic degradation. This increased susceptibility involved subunits that contain electron carriers, such as FMN and iron-sulfur clusters, as well as subunits that lack electron carriers. Results shown elsewhere on changes in the cross-linking pattern of complex I subunits when the enzyme was pretreated with NADH or NADPH (Belogrudov, G., and Hatefi, Y. (1994) Biochemistry 33, 4571-4576) also indicated that complex I undergoes extensive conformation changes when reduced by substrate. Furthermore, we had previously shown that in submitochondrial particles the affinity of complex I for NAD increases by >/=20-fold in electron transfer from succinate to NAD when the particles are energized by ATP hydrolysis. Together, these results suggest that energy coupling in complex I may involve protein conformation changes as a key step. In addition, it has been shown here that treatment of complex I with trypsin in the presence of NADPH, but not NADH or NAD(P), produced from the 39-kDa subunit a 33-kDa degradation product that resisted further hydrolysis. Like the 39-kDa subunit, the 33-kDa product bound to a NADP-agarose affinity column, and could be eluted with a buffer containing NADPH. It is possible that together with the acyl carrier protein of complex I the NADP(H)-binding 39-kDa subunit is involved in intramitochondrial fatty acid synthesis.  相似文献   

9.
Both NADH and NADPH reduce the permeability of the mitochondrial outer membrane to ADP. This is specific for the outer membrane and uncorrelated with the respiratory control ratio. This could result in a 7-fold difference between the concentration of ADP in the intermembrane space and that in the external environment (at 5 microM ADP). In both cases the permeability declines by a factor of 5, but NADH is more potent: KD = 86 microM for NADH versus 580 microM for NADPH. The lower apparent affinity for NADPH is partly explained by Mg2+-NADPH being the active species, and under our conditions only 30% of the NADPH is in this form. The corrected KD is 184 microM. Free NADH has the same charge as the Mg2+-NADPH complex, and thus both likely bind to the same site. The ability of NADH and NADPH to induce the closure of reconstituted VDAC channels is consistent with VDAC being the main pathway for metabolite flow across the outer membrane. Oncotic pressure, effective at inducing VDAC closure, also decreases the outer membrane permeability. Thus, in the presence of cytosolic colloidal osmotic pressure NAD(P)H may inhibit mitochondrial catabolic pathways and divert reducing equivalents to anabolic pathways.  相似文献   

10.
Epoxide metabolism in the aerobic bacterium Xanthobacter strain Py2 proceeds by an NADPH- and NAD+-dependent carboxylation reaction that forms beta-keto acids as products. Epoxide carboxylase, the enzyme catalyzing this reaction, was resolved from the soluble fraction of cell-free extracts into four protein components that are obligately required for functional reconstitution of epoxide carboxylase activity. One of these components, component II, has previously been purified and characterized as an NADPH:disulfide oxidoreductase. In the present study, the three additional epoxide carboxylase components have been purified to homogeneity and characterized. These component proteins are as follows: component I, a homohexameric protein consisting of 41.7-kDa subunits; component III, a dimeric protein consisting of 26.0- and 26.2-kDa polypeptides; and component IV, a dimeric protein consisting of a single 25.4-kDa polypeptide. Component I contained 5 mol of tightly bound zinc per mol of protein. Component I was specifically inactivated by methylepoxypropane, a time-dependent irreversible inactivator of epoxide carboxylase activity, suggesting that this component plays an integral role in epoxide binding and activation. No metals or organic cofactors were detected for components III and IV. The molecular weights, N-terminal sequences, and amino acid compositions of the purified epoxide carboxylase components were determined and found to correlate with open reading frames within and adjacent to a cloned fragment of DNA that complements Xanthobacter Py2 mutants defective in epoxide degradation. Using the purified epoxide carboxylase system, epoxide carboxylation was found to be stoichiometrically coupled to the transhydrogenation of pyridine nucleotide cofactors according to the following equation: epoxypropane + CO2 + NADPH + NAD+ --> acetoacetate + H+ + NADP+ + NADH.  相似文献   

11.
In differentiated tissues, such as muscle and brain, increased adenosine monophosphate (AMP) levels stimulate glycolytic flux rates. In the breast cancer cell line MCF-7, which characteristically has a constantly high glycolytic flux rate, AMP induces a strong inhibition of glycolysis. The human breast cancer cell line MDA-MB-453, on the other hand, is characterized by a more differentiated metabolic phenotype. MDA-MB-453 cells have a lower glycolytic flux rate and higher pyruvate consumption than MCF-7 cells. In addition, they have an active glycerol 3-phosphate shuttle. AMP inhibits cell proliferation as well as NAD and NADH synthesis in both MCF-7 and MDA-MB-453 cells. However, in MDA-MB-453 cells glycolysis is slightly activated by AMP. This disparate response of glycolytic flux rate to AMP treatment is presumably caused by the fact that the reduced NAD and NADH levels in AMP-treated MDA-MB-453 cells reduce lactate dehydrogenase but not cytosolic glycerol-3-phosphate dehydrogenase reaction. Due to the different enzymatic complement in MCF-7 cells, proliferation is inhibited under glucose starvation, whereas MDA-MB-453 cells grow under these conditions. The inhibition of cell proliferation correlates with a reduction in glycolytic carbon flow to synthetic processes and a decrease in phosphotyrosine content of several proteins in both cell lines.  相似文献   

12.
The hyperthermophilic archaeum Thermoproteus tenax possesses two glyceraldehyde-3-phosphate dehydrogenases differing in cosubstrate specificity and phosphate dependence of the catalyzed reaction. NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase catalyzes the phosphate-independent irreversible oxidation of D-glyceraldehyde 3-phosphate to 3-phosphoglycerate. The coding gene was cloned, sequenced, and expressed in Escherichia coli. Sequence comparisons showed no similarity to phosphorylating glyceraldehyde-3-phosphate dehydrogenases but revealed a relationship to aldehyde dehydrogenases, with the highest similarity to the subgroup of nonphosphorylating glyceraldehyde-3-phosphate dehydrogenases. The activity of the enzyme is affected by a series of metabolites. All effectors tested influence the affinity of the enzyme for its cosubstrate NAD+. Whereas NADP(H), NADH, and ATP reduce the affinity for the cosubstrate, AMP, ADP, glucose 1-phosphate, and fructose 6-phosphate increase the affinity for NAD+. Additionally, most of the effectors investigated induce cooperativity of NAD+ binding. The irreversible catabolic oxidation of glyceraldehyde 3-phosphate, the control of the enzyme by energy charge of the cell, and the regulation by intermediates of glycolysis and glucan degradation identify the NAD+-dependent glyceraldehyde-3-phosphate dehydrogenase as an integral constituent of glycolysis in T. tenax. Its regulatory properties substitute for those lacking in the reversible nonregulated pyrophosphate-dependent phosphofructokinase in this variant of the Embden-Meyerhof-Parnas pathway.  相似文献   

13.
The two diastereotopic protons at C4 of NAD(P)H are seen separately in 1H-NMR spectra. This fact was used to determine the stereospecificity at C4 of NAD(P) for the NADP-dependent alcohol dehydrogenase from Methanogenium organophilum and for the NAD-dependent (R)-2-hydroxyglutarate dehydrogenase from Acidaminococcus fermentans. The reduction of NADP+ with [2H6]ethanol was found to yield (4R)-[4-2H1]NADPH and the oxidation of (4R)-[4-2H1]NADH with 2-oxoglutarate to yield unlabelled [4-1H]NAD+. These results indicate that both enzymes are Re-face stereospecific at C4 of the pyridine nucleotides.  相似文献   

14.
Addition of insulin or a physiological ratio of ketone bodies to buffer with 10 mM glucose increased efficiency (hydraulic work/energy from O2 consumed) of working rat heart by 25%, and the two in combination increased efficiency by 36%. These additions increased the content of acetyl CoA by 9- to 18-fold, increased the contents of metabolites of the first third of the tricarboxylic acid (TCA) cycle 2- to 5-fold, and decreased succinate, oxaloacetate, and aspartate 2- to 3-fold. Succinyl CoA, fumarate, and malate were essentially unchanged. The changes in content of TCA metabolites resulted from a reduction of the free mitochondrial NAD couple by 2- to 10-fold and oxidation of the mitochondrial coenzyme Q couple by 2- to 4-fold. Cytosolic pH, measured using 31P-NMR spectra, was invariant at about 7.0. The total intracellular bicarbonate indicated an increase in mitochondrial pH from 7.1 with glucose to 7.2, 7.5 and 7.4 with insulin, ketones, and the combination, respectively. The decrease in Eh7 of the mitochondrial NAD couple, Eh7NAD+/NADH, from -280 to -300 mV and the increase in Eh7 of the coenzyme Q couple, Eh7Q/QH2, from -4 to +12 mV was equivalent to an increase from -53 kJ to -60 kJ/2 mol e in the reaction catalyzed by the mitochondrial NADH dehydrogenase multienzyme complex (EC 1.6.5.3). The increase in the redox energy of the mitochondrial cofactor couples paralleled the increase in the free energy of cytosolic ATP hydrolysis, delta GATP. The potential of the mitochondrial relative to the cytosolic phases, Emito/cyto, calculated from delta GATP and delta pH on the assumption of a 4 H+ transfer for each ATP synthesized, was -143 mV during perfusion with glucose or glucose plus insulin, and decreased to -120 mV on addition of ketones. Viewed in this light, the moderate ketosis characteristic of prolonged fasting or type II diabetes appears to be an elegant compensation for the defects in mitochondrial energy transduction associated with acute insulin deficiency or mitochondrial senescence.  相似文献   

15.
Conserved and semiconserved acidic and basic residues of the beta subunit of the proton-pumping nicotinamide nucleotide transhydrogenase from Escherichia coli potentially involved in proton pumping were investigated. Out of 16 charged residues studied, 6 have not been previously investigated. The most dramatic effects of mutation were observed with beta H91, beta D392, and beta K424. beta H91E showed a pronounced shift of the pH optimum for both reduction of thio-NADP+ by NADH (forward reaction) and reduction of 3-acetylpyridine-NAD+ by NADPH (reverse reaction) to lower pH. This mutant catalyzed a cyclic reduction of 3-acetylpyridine-NAD+ by NADH in the presence of NADP(H) with a pH profile also shifted toward a lower pH. These results are consistent with a mechanism where the normal forward and reverse reactions are indeed limited by protonation/deprotonation of beta H91. The cyclic reaction was affected by mutations of beta H91, probably through conformational changes involving the active NADP(H) site. The beta D392A mutant was inactive with regard to forward and reverse reactions, but showed a wild-type-like pH dependence for the partly active cyclic reaction. However, Km,app for NADP(H) in this reaction was elevated 50-100-fold, suggesting that beta D392 is located in or near the NADP(H)-binding site. Transhydrogenases contain a conserved beta K424-beta R425-beta S426 sequence that has been proposed to be important for NADP(H) binding. beta K424R was strongly inhibited and showed an 18-fold increased Km,app for NADPH in the reverse reaction as compared to wild type. Consequently, this mutation affected all NADP(H)-linked activities and essentially abolished the unspecific interaction of NAD(H) with this site. The pH dependences of the forward and reverse reactions, as well as the cyclic reaction, were shifted to a lower pH as compared to the wild-type enzyme, and the salt dependence was also altered.  相似文献   

16.
To compare roles of specific enzymes in supply of NADPH for cellular biosynthesis, collections of yeast mutants were constructed by gene disruptions and matings. These mutants include haploid strains containing all possible combinations of deletions in yeast genes encoding three differentially compartmentalized isozymes of NADP+-specific isocitrate dehydrogenase and in the gene encoding glucose-6-phosphate dehydrogenase (Zwf1p). Growth phenotype analyses of the mutants indicate that either cytosolic NADP+-specific isocitrate dehydrogenase (Idp2p) or the hexose monophosphate shunt is essential for growth with fatty acids as carbon sources and for sporulation of diploid strains, a condition associated with high levels of fatty acid synthesis. No new biosynthetic roles were identified for mitochondrial (Idp1p) or peroxisomal (Idp3p) NADP+-specific isocitrate dehydrogenase isozymes. These and other results suggest that several major presumed sources of biosynthetic reducing equivalents are non-essential in yeast cells grown under many cultivation conditions. To develop an in vivo system for analysis of metabolic function, mammalian mitochondrial and cytosolic isozymes of NADP+-specific isocitrate dehydrogenase were expressed in yeast using promoters from the cognate yeast genes. The mammalian mitochondrial isozyme was found to be imported efficiently into yeast mitochondria when fused to the Idp1p targeting sequence and to substitute functionally for Idp1p for production of alpha-ketoglutarate. The mammalian cytosolic isozyme was found to partition between cytosolic and organellar compartments and to replace functionally Idp2p for production of alpha-ketoglutarate or for growth on fatty acids in a mutant lacking Zwf1p. The mammalian cytosolic isozyme also functionally substitutes for Idp3p allowing growth on petroselinic acid as a carbon source, suggesting partial localization to peroxisomes and provision of NADPH for beta-oxidation of that fatty acid.  相似文献   

17.
Previous studies demonstrating reduced plasma concentrations of ascorbic acid (AA) in diabetes and interactions between this vitamin and biochemical mechanisms such as synthesis of structural proteins, oxidative stress, polyol pathway and nonenzymatic glycation of proteins suggest that disturbed AA metabolism may be important in the pathogenesis of diabetic microangiopathy. However, limited information is available on the concentration of AA in tissues which develop diabetic complications. This study demonstrates reduced renal but not sciatic nerve or plasma AA concentration in two animal models of insulin-dependent diabetes mellitus, namely the STZ-diabetic rat and the spontaneously diabetic BB rat. Decreased lens AA concentration was also observed in STZ-diabetic rats. Improvement of glycaemic control by insulin treatment (albeit insufficient to achieve normoglycaemia) partially corrected lens and renal AA concentration in STZ-diabetic rats. AA treatment increased kidney and lens AA concentrations of STZ-diabetic and non-diabetic rats and corrected the abnormalities observed for untreated diabetic rats. Sciatic nerve AA concentration was not increased by AA treatment in any group. Tissue ratios of dehydroascorbic acid (DHAA)/AA, one index of oxidative stress, were not different between the diabetic and non-diabetic groups and were unaltered by AA supplementation. AA treatment of STZ-diabetic rats had no effect on elevated tissue concentrations of glucose, sorbitol and fructose or reduced myo-inositol concentration. The effect of reduced tissue AA levels in diabetes on either collagen synthesis or ability to combat increased free radical production is not known. However, correction of abnormal kidney and lens AA concentrations in experimental diabetes by AA supplementation suggests that if AA does have a role in the development or progression of the renal and ocular complications of diabetes, this treatment could be beneficial.  相似文献   

18.
3-Deoxy-3-fluoro-D-glucose (3FG) administered by injection is toxic to adult Locusta migratoria or Schistocerca gregaria (LD50, 4.8 mg/g). temperature-programmed and isothermal gas chromatographic analysis of poisoned locust haemolymph reveals the presence of a fluorinated metabolite identified as 3-deoxy-3-fluoro-D-glucitol (3FGL). The enzymes responsible for the accumulation of this metabolite are located in the fat body of the insect and partially purified as aldose reductase (alditol: NADP+ 1-oxidoreductase, EC 1.1.1.21) and sorbitol dehydrogenase (L-iditol: NAD+ 5-oxidoreductase EC 1.1.1.14) 3FGL is shown to be both a competitive inhibitor of the NAD-linked sorbitol dehydrogenase with Ki 8?x 10(-2) M as well as a substrate with Km 0.5 M. A kinetic rate equation is derived and verified to account for the kinetic duality of 3FGL. These results partially explain the toxic effects of 3FG and are consistent with the presence of a hitherto undetected sorbitol metabolism in locusts.  相似文献   

19.
We have measured the levels of glucose and sugar alcohols in the lens epithelium of 24 non-diabetic and 21 diabetic patients undergoing extracapsular cataract extraction for age-related cataract. Lens status was assessed preoperatively according to the Lens Opacities Classification System II. In comparison to non-diabetics, the lens epithelia of diabetic patients have increased levels of glucose and sorbitol, and lower content of myo-inositol. Both myo-inositol decrease and sorbitol accumulation are strictly related to the fasting blood sugar level. No correlation or trend was detected in diabetic subjects between myo-inositol or sorbitol level and cataract type. In non-diabetic patients nuclear opacification was associated with a significant increase of myo-inositol level in lens epithelium.  相似文献   

20.
Reduction of the antioxidant lipoic acid has been proposed to be catalyzed in vivo by lipoamide dehydrogenase (LipDH) or glutathione reductase (GR). We have found that thioredoxin reductase (TR) from calf thymus, calf liver, human placenta, and rat liver efficiently reduced both lipoic acid and lipoamide with Michaelis-Menten type kinetics in NADPH-dependent reactions. In contrast to LipDH, lipoic acid was reduced almost as efficiently as lipoamide. Under equivalent conditions at 20 degrees C, pH 8.0, mammalian TR reduced lipoic acid by NADPH 15 times more efficiently than the corresponding NADH dependent reduction catalyzed by LipDH (297 min-1 for TR vs. 20.3 min-1 for LipDH). Moreover, TR was 2.5 times faster in reducing lipoic acid with NADPH than in catalyzing the reverse reaction (oxidation of dihydrolipoic acid with NADP+). In contrast, LipDH was only 0.048 times as efficient in the forward reaction as compared to the reverse reaction (using NADH and NAD+). We conclude that all or part of the previously described NADPH-dependent lipoamide dehydrogenase (diaphorase) activities in mammalian systems should be attributed to TR. Our results suggest that in mammalian cells a significant part of the therapeutically important reduction of lipoic acid is catalyzed by thioredoxin reductase.  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