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.
The inhibition of Escherichia coli isocitrate dehydrogenase by glyoxylate and oxaloacetate was examined. The shapes of the progress curves in the presence of the inhibitors depended on the order of addition of the assay components. When isocitrate dehydrogenase or NADP+ was added last, the rate slowly decreased until a new, inhibited, steady state was obtained. When isocitrate was added last, the initial rate was almost zero, but the rate increased slowly until the same steady-state value was obtained. Glyoxylate and oxaloacetate gave competitive inhibition against isocitrate and uncompetitive inhibition against NADP+. Product-inhibition studies showed that isocitrate dehydrogenase obeys a compulsory-order mechanism, with coenzyme binding first. Glyoxylate and oxaloacetate bind to and dissociate from isocitrate dehydrogenase slowly. These observations can account for the shapes of the progress curves observed in the presence of the inhibitors. Condensation of glyoxylate and oxaloacetate produced an extremely potent inhibitor of isocitrate dehydrogenase. Analysis of the reaction by h.p.l.c. showed that this correlated with the formation of oxalomalate. This compound decomposed spontaneously in assay mixtures, giving 4-hydroxy-2-oxoglutarate, which was a much less potent inhibitor of the enzyme. Oxalomalate inhibited isocitrate dehydrogenase competitively with respect to isocitrate and was a very poor substrate for the enzyme. The data suggest that the inhibition of isocitrate dehydrogenase by glyoxylate and oxaloacetate is not physiologically significant.  相似文献   

2.
Summary The utilization by yeast of two carbon sources is carried out through the operation of the glyoxylic acid cycle. Kinetic data from the isocitrate transforming enzymes suggest that the flow of isocitrate through the glyoxylic acid cycle depends upon the inhibition of the isocitrate decarboxylating enzymes. Both isocitrate dehydrogenases are inhibited by a mixture of glyoxylate + oxaloacetate, but for the reasons described in the text we consider that this inhibition is of no physiological significance. On the other hand, we have found that NADPH is a competitive inhibitor of NADP-isocitrate dehydrogenase with respect to NADP+, with a KI similar to its KM. It also produces an additive effect on the NADH-produced inhibition of NAD-isocitrate dehydrogenase. We propose NADPH as the compound that channels the utilization of isocitrate into the glyoxylic acid cycle. This is supported by the finding of an increased NADPH/NADP+ ratio in acetate grown yeast with respect to glucose grown cells.  相似文献   

3.
The RS-isomers of beta-mercapto-alpha-ketoglutarate, beta-methylmercapto-alpha-ketoglutarate and beta-methylmercapto-alpha-hydroxyglutarate have been synthesized. Beta-Mercapto-alpha-ketoglutarate was a potent inhibitor, competitive with isocitrate and noncompetitive with NADP+, of the mitochondrial NADP-specific isozyme from pig heart (Ki = 5 nM; Km (DL-isocitrate)/Ki(RS-beta-mercapto-alpha-ketoglutarate) = 650) and pig liver, the cytosolic isozyme from pig liver (I0.5 = 23 nM), and the NADP-linked enzymes from yeast (Ki = 58 nM) and Escherichia coli (Ki = 58 nM) at pH 7.4 and with Mg2+ as activator. beta-Mercapto-alpha-ketoglutarate was also an effective inhibitor of NADP-isocitrate-dehydrogenase activity in intact liver mitochondria. beta-Mercapto-alpha-ketoglutarate was a much less potent inhibitor for heart NAD-isocitrate dehydrogenase (Ki = 520 nM) than for the NADP-specific enzyme. beta-Methylmercapto-alpha-ketoglutarate (I0.5 = 10 microM) was a much less effective inhibitor than the beta-mercapto derivative for heart NADP-isocitrate dehydrogenase. The beta-sulfur substituted alpha-ketoglutarates were substrates for the oxidation of NADPH by heart NADP-isocitrate dehydrogenase without requiring CO2. beta-Methylmercapto-alpha-hydroxyglutarate, the expected product of reduction of beta-methylmercapto-alpha-ketoglutarate, did not cause reduction of NADP+ but it was an inhibitor competitive with isocitrate for NADP-isocitrate dehydrogenase. The beta-sulfur substituted alpha-ketoglutarate derivatives were alternate substrates for alpha-ketoglutarate dehydrogenase and the cytosolic and mitochondrial isozymes of heart aspartate aminotransferase but had no effect on glutamate dehydrogenase or alanine aminotransferase.  相似文献   

4.
R S Ehrlich  R F Colman 《Biochemistry》1976,15(18):4034-4041
The interaction of manganous ions with pig heart triphosphopyridine nucleotide (TPN) specific isocitrate dehydrogenase has been studied by kinetic experiments and by direct ultrafiltration measurements of manganous ion binding. At low metal ion concentrations, a lag is observed in the time-dependent production of reduced triphosphopyridine nucleotide (TPNH) that can be eliminated by adding 20 muM TPNH to the initial reaction mixture. A plot of 1/upsilon vs. 1/ (Mn2+) obtained at relatively high TPNH concentrations (20 muM) is linear and yields of Km value of 2 muM for metal ion, which is comparable to the direct binding constant measured in the presence of isocitrate. A similar plot at low TPNH concentrations (2 muM) reveals a biphasic relationship: at high metal concentrations the points are collinear with those obtained at high levels of TPNH, but at low metal concentrations that line is characterized by a Km of 19 muM for Mn2+. A difference in the deuterium oxide solvent isotope effect on Vmax observed with 20 muM TPNH as compared with 2 muM TPNH suggests that at high TPNH concentrations or high manganous ion concentrations the rate-limiting step is the dehydrogenation of isocitrate, while at low manganous ion concentrations and low TPNH concentrations, the slow step is the decarboxylation of enzyme-bound oxalosuccinate. Evidence to support this hypothesis is provided by the sensitivity to isocitrate concentration of the Km for total manganese measured in the presence of 20 muM TPNH that contrasts with the relative insensitivity to isocitrate of the Km measured at 2 muM TPNH and low manganous ion concentration. Direct measurements of oxalosuccinate decarboxylation reveal that the Vmax and the Km for manganous ion are influenced by the presence of oxidized or reduced TPN with the Km being lowest (5-7 muM) in the presence of TPNH. The dependence of the Km for manganous ion on the presence of substrate, TPN, and TPNH, is responsible for the variation with conditions in the rate-determining step. The enzyme binds only 1 mol of metal ion and 1 mol of isocitrate/mol of protein under all conditions. The pH dependence of the binding of free manganous ion, free isocitrate, and manganous-isocitrate complex indicates differences in the interaction of these species with isocitrate dehydrogenase. These results can be described in terms of two functions for manganous ion in the reactions catalyzed by isocitrate dehydrogenase, each of which requires a distinct binding site for metal ion: in the dehydrogenation step, Mn2+ facilitates the binding of the substrate isocitrate, and in the decarboxylation step it may stabilize the enolate of alpha-ketoglutarate which is generated.  相似文献   

5.
The oxidative decarboxylation of D-isocitrate catalyzed by NADP-linked isocitrate dehydrogenase is activated by NADPH, the product of the reaction. We analyzed the autocatalytic behavior exhibited by the enzyme during the steady-state kinetics. NADP acts as a competitive inhibitor toward NADPH in the catalytic activation. In a large concentration range of the reduced and oxidized coenzymes, the activity of the enzyme is proportional to the ratio (NADPH)/(NADP). The results are compared with the results of experiments done with other NADP-linked decarboxylating dehydrogenases. Two different models are presented in order to explain the mechanism of action of isocitrate dehydrogenase, according to our data.  相似文献   

6.
J L Gabriel  G W Plaut 《Biochemistry》1991,30(10):2594-2599
The present results suggest that the enzyme modifier citrate and the substrate isocitrate are bound at different sites on yeast NAD-specific isocitrate dehydrogenase and that citrate diminishes the binding of the positive effector 5'-AMP, thereby causing a decreased rate of enzyme catalysis. This interpretation differs from the earlier proposal that citrate can replace isocitrate at an activator site on the enzyme and can cause inhibition by binding at its catalytic site [Atkinson et al. (1965) J. Biol. Chem. 240, 2682]. The present proposal is supported by the following observations: At constant subsaturating levels of isocitrate, NAD+, and Mg2+ without AMP, up to 10 mM citrate was an activator and not an inhibitor. Citrate decreased velocity for AMP-activated enzyme; however, with increasing citrate the specific activity with AMP asymptotically approached but did not decrease below the level of the enzyme maximally activated by citrate in the absence of AMP. When added singly, AMP decreased S0.5 for isocitrate without changing the Hill number (n), whereas citrate lowered n without changing S0.5 for isocitrate. The difference in action of these modifiers indicated that they were bound at separate sites on the enzyme. The binding of citrate appeared to cause a conformational change in the protein that lowered the enzyme's affinity for AMP. This was consistent with the findings that citrate (or the citrate agonist fluorocitrate) (i) resulted in an increase in S0.5 for isocitrate with the AMP-activated enzyme and (ii) decreased binding of the positive effector analogue TNP-AMP as measured by fluorescence change.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
1. The sensitivity of the NAD(+)-specific isocitrate dehydrogenase from baker's yeast towards inhibition by anions decreases with decrease in pH. The patterns of the pH-dependence of the enzymic activity can be explained by this effect. 2. In the presence of a high isocitrate concentration, citrate, unlike AMP, has no antagonizing effect on the inhibition of the enzyme by anions. In the presence of AMP, citrate inhibits the enzyme at high isocitrate concentration and activates at low isocitrate concentration. 3. The effects on the enzymic activity of the previous incubation of the enzyme were studied in relation to the substrate concentration, the chloride concentration and the presence of citrate and AMP.  相似文献   

8.
Isocitrate dehydrogenase (ICDH) from Hydrogenobacter thermophilus catalyzes the reduction of oxalosuccinate, which corresponds to the second step of the reductive carboxylation of 2-oxoglutarate in the reductive tricarboxylic acid cycle. In this study, the oxidation reaction catalyzed by H. thermophilus ICDH was kinetically analyzed. As a result, a rapid equilibrium random-order mechanism was suggested. The affinities of both substrates (isocitrate and NAD+) toward the enzyme were extremely low compared to other known ICDHs. The binding activities of isocitrate and NAD+ were not independent; rather, the binding of one substrate considerably promoted the binding of the other. A product inhibition assay demonstrated that NADH is a potent inhibitor, although 2-oxoglutarate did not exhibit an inhibitory effect. Further chromatographic analysis demonstrated that oxalosuccinate, rather than 2-oxoglutarate, is the reaction product. Thus, it was shown that H. thermophilus ICDH is a nondecarboxylating ICDH that catalyzes the conversion between isocitrate and oxalosuccinate by oxidation and reduction. This nondecarboxylating ICDH is distinct from well-known decarboxylating ICDHs and should be categorized as a new enzyme. Oxalosuccinate-reducing enzyme may be the ancestral form of ICDH, which evolved to the extant isocitrate oxidative decarboxylating enzyme by acquiring higher substrate affinities.  相似文献   

9.
NAD(+)-dependent isocitrate dehydrogenase from Saccharomyces cerevisiae is composed of two nonidentical subunits, designated IDH1 (Mr approximately 40,000) and IDH2 (Mr approximately 39,000). We have isolated and characterized a yeast genomic clone containing the IDH2 gene. The amino acid sequence deduced from the gene indicates that IDH2 is synthesized as a precursor of 369 amino acids (Mr 39,694) and is processed upon mitochondrial import to yield a mature protein of 354 amino acids (Mr 37,755). Amino acid sequence comparison between S. cerevisiae IDH2 and S. cerevisiae NADP(+)-dependent isocitrate dehydrogenase shows no significant sequence identity, whereas comparison of IDH2 and Escherichia coli NADP(+)-dependent isocitrate dehydrogenase reveals a 33% sequence identity. To confirm the identity of the IDH2 gene and examine the relationship between IDH1 and IDH2, the IDH2 gene was disrupted by genomic replacement in a haploid yeast strain. The disruption strain expressed no detectable IDH2, as determined by Western blot analysis, and was found to lack NAD(+)-dependent isocitrate dehydrogenase activity, indicating that IDH2 is essential for a functional enzyme. Overexpression of IDH2, however, did not result in increased NAD(+)-dependent isocitrate dehydrogenase activity, suggesting that both IDH1 and IDH2 subunits are required for catalytic activity. The disruption strain was unable to utilize acetate as a carbon source and exhibited a 2-fold slower growth rate than wild type strains on glycerol or lactate. This growth phenotype is consistent with NAD(+)-dependent isocitrate dehydrogenase performing an essential role in the oxidative function of the citric acid cycle.  相似文献   

10.
Oxalic acid plays a pivotal role in the adaptation of the soil microbe Pseudomonas fluorescens to aluminum (Al) stress. Its production via the oxidation of glyoxylate necessitates a major reconfiguration of the enzymatic reactions involved in the tricarboxylic acid (TCA) cycle. The demand for glyoxylate, the precursor of oxalic acid appears to enhance the activity of isocitrate lyase (ICL). The activity of ICL, an enzyme that participates in the cleavage of isocitrate to glyoxylate and succinate incurred a 4-fold increase in the Al-stressed cells. However, the activity of isocitrate dehydrogenase, a competitor for the substrate isocitrate, appeared to be diminished in cells exposed to Al compared to the control cells. While the demand for oxalate in Al-stressed cells also negatively influenced the activity of the enzyme alpha-ketoglutarate dehydrogenase complex, no apparent change in the activity of malate synthase was recorded. Thus, it appears that the TCA cycle is tailored in order to generate the necessary precursor for oxalate synthesis as a consequence of Al-stress.  相似文献   

11.
The DPN-specific isocitrate dehydrogenase of pig heart is totally and irreversibly inactivated by 0.05 M potassium cyanate at pH 7.4 A plot of the rate constant versus cyanate concentration is not linear, but rather exhibits saturation kinetics, implying that cyanate may bind to the enzyme to give an enzyme-cyanate complex (K equal 0.125 M) prior to the covalent reaction. In the presence of manganous ion the addition of isocitrate protects the enzyme against cyanate inactivation, indicating that chemical modification occurs in the active site region of the enzyme. The dependence of the decrease of the rate constant for inactivation on the isocitrate concentration yields a dissociation constant for the enzyme-manganese-isocitrate complex which agrees with the Michaelis constant. The allosteric activator ADP, which lowers the Michaelis constant for isocitrate, does not itself significantly affect the cyanate reaction; however, it strikingly enhances the protection by isocitrate. The addition of the chelator EDTA essentially prevents protection by isocitrate and manganous ion, demonstrating the importance of the metal ion in this process. The substrate alpha-ketoglutarate and the coenzymes DPN and DPNH do not significantly affect the rate of modification of the enzymes by cyanate. Incubation of isocitrate dehydrogenase with 14C-labeled potassium cyanate leads to the incorporation of approximately 1 mol of radioactive cyanate per peptide chain concomitant with inactivation. Analysis of acid hydrolysates of the radioactive enzyme reveals that lysyl residues are the sole amino acids modified. These results suggest that cyanate, or isocyanic acid, may bind to the active site of this enzyme as an analogue of carbon dioxide and carbamylate a lysyl residue at the active site.  相似文献   

12.
13.
Mitochondrial NAD(H)-specific isocitrate dehydrogenase was purified from Saccharomyces cerevisiae for analyses of subunit structure and expression. Two subunits of the enzyme with different molecular weights (39,000 and 40,000) and slightly different isoelectric points were resolved by denaturing electrophoretic techniques. Sequence analysis of the purified subunits showed that the polypeptides have different amino termini. By using an antiserum to the native enzyme prepared in rabbits, subunit-specific immunoglobulin G fractions were obtained by affinity purification, indicating that the subunits are also immunochemically distinct. The levels of NAD(H)-specific isocitrate dehydrogenase activity and immunoreactivity were found to correlate closely with those of a second tricarboxylic acid cycle enzyme, malate dehydrogenase, in yeast cells grown under a variety of conditions. S. cerevisiae mutants with defects in NAD(H)-specific isocitrate dehydrogenase were identified by screening a collection of yeast mutants with acetate-negative growth phenotypes. Immunochemical assays were used to demonstrate that one mutant strain lacks the 40,000-molecular-weight subunit (IDH1) and that a second strain lacks the 39,000-molecular-weight subunit (IDH2). Mitochondria isolated from the IDH1 and IDH2 mutants exhibited a markedly reduced capacity for utilization of either isocitrate or citrate for respiratory O2 consumption. This confirms an essential role for NAD(H)-specific isocitrate dehydrogenase in oxidative functions in the tricarboxylic acid cycle.  相似文献   

14.
The present results show that the NADP specific isocitrate dehydrogenase from pig heart exhibits a time lag before the reaction rate approaches a constant value at low metal ion concentrations. Addition of NADPH or EDTA to the assay mixture abolished the lag, and will under certain conditions activate the enzyme.The lag time increased with increasing concentrations of isocitrate and decreased with increasing enzyme concentration. The NADP and metal ion concentration affected the lag in a complex manner. At low NADP and isocitrate concentration, the lag was reduced 50% by an NADPH concentration of less than 2 μm. Stopped flow experiments showed that premixing of NADP or NADPH with the enzyme abolished the effect of NADPH on the lag time. NADPH activated the enzyme at high NADP concentrations. This activating effect could be accounted for by removal of substrate inhibition by NADP.Evidence was obtained to show that the effect of NADPH on the activity was caused by binding of the reduced coenzyme to a site separate from the normal coenzyme binding site. Binding of metal ions by the reduced coenzyme is probably of importance as EDTA affects the lag time and activity in a manner similar to NADPH. The NADPH effect seems to be a general property of NADP-linked isocitrate dehydrogenases.  相似文献   

15.
Aluminum ion inhibited yeast glycerol kinase competitively with respect to the substrate MgATP. The K value of the enzyme for aluminum ion was about 3 M. Spermine at physiological concentrations prevented glycerol kinase from the inhibition by aluminum ion. Nuclear magnetic resonance spectroscopy showed the specific elimination by spermine of aluminum from the metal-ATP complex, but no dissociation of MgATP complex by spermine. Inhibition by aluminum ion of glycerol kinase as well as hexokinase can reduce the utilization of energy fuel in yeast. Change in polyamine concentration may control energy production in vivo, and is responsible for the development of age-related aluminum toxicity. © Rapid Science 1998.  相似文献   

16.
17.
Doyle SA  Beernink PT  Koshland DE 《Biochemistry》2001,40(14):4234-4241
Isocitrate dehydrogenase (IDH) catalyzes the oxidative decarboxylation of isocitrate and has negligible activity toward other (R)-malate-type substrates. The S113E mutant of IDH significantly improves its ability to utilize isopropylmalate as a substrate and switches the substrate specificity (k(cat)/K(M)) from isocitrate to isopropylmalate. To understand the structural basis for this switch in substrate specificity, we have determined the crystal structure of IDH S113E in a complex with isopropylmalate, NADP, and Mg(2+) to 2.0 A resolution. On the basis of a comparison with previously determined structures, we identify distinct changes caused by the amino acid substitution and by the binding of substrates. The S113E complex exhibits alterations in global and active site conformations compared with other IDH structures that include loop and helix conformational changes near the active site. In addition, the angle of the hinge that relates the two domains was altered in this structure, which suggests that the S113E substitution and the binding of substrates act together to promote catalysis of isopropylmalate. Ligand binding results in reorientation of the active site helix that contains residues 113 through 116. E113 exhibits new interactions, including van der Waals contacts with the isopropyl group of isopropylmalate and a hydrogen bond with N115, which in turn forms a hydrogen bond with NADP. In addition, the loop and helix regions that bind NADP are altered, as is the loop that connects the NADP binding region to the active site helix, changing the relationship between substrates and enzyme. In combination, these interactions appear to provide the basis for the switch in substrate specificity.  相似文献   

18.
Alpha-Methylisocitrate (3-hydroxy-1,2,3-butanetricarboxylate) is a potent inhibitor, competitive with isocitrate (1-hydroxy-1,2,3-propanetricarboxylate), of the TPN-linked isocitrate dehydrogenase from bovine heart and rat liver; it does not inhibit the DPN-specific enzyme from these tissues. In the presence of magnesium ion, values of Kis for DL-alpha-methylisocitrate for purified bovine heart enzyme, rat liver cytosol, and rat liver mitochondrial extract were in the range of 0.1 muM to 0.3 muM. This compared to values of apparent Km for DL-isocitrate for the same tissue preparations of 14 muM to 20 muM. One of the DL isomer pairs of alpha-methylisocitrate was inactive; the observations suggest that it is threo-alpha-methylisocitrate which inhibits TPN-linked isocitrate dehydrogenase. A method of synthesis of DL-threo-alpha-methylisocitric lactone (2-methyl-5-oxo-2,3-furandicarboxylic acid) from dimethyl trans-epoxymethylsuccinate and dimethylmalonate is described.  相似文献   

19.
A thermostable homodimeric isocitrate dehydrogenase from the hyperthermophilic sulfate-reducing archaeon Archaeoglobus fulgidus was purified and characterized. The mol. mass of the isocitrate dehydrogenase subunit was 42 kDa as determined by SDS-PAGE. Following separation by SDS-PAGE, A. fulgidus isocitrate dehydrogenase could be renatured and detected in situ by activity staining. The enzyme showed dual coenzyme specificity with a high preference for NADP+. Optimal temperature for activity was 90° C or above, and a half-life of 22 min was found for the enzyme when incubated at 90° C in a 50 mM Tricine-KOH buffer (pH 8.0). Based on the N-terminal amino acid sequence, the gene encoding the isocitrate dehydrogenase was cloned. DNA sequencing identified the icd gene as an open reading frame encoding a protein of 412 amino acids with a molecular mass corresponding to that determined for the purified enzyme. The deduced amino acid sequence closely resembled that of the isocitrate dehydrogenase from the archaeon Caldococcus noboribetus (59% identity) and bacterial isocitrate dehydrogenases, with 57% identity with isocitrate dehydrogenase from Escherichia coli. All the amino acid residues directly contacting substrate and coenzyme (except Ile-320) in E. coli isocitrate dehydrogenase are conserved in the enzyme from A. fulgidus. The primary structure of A. fulgidus isocitrate dehydrogenase confirmes the presence of Bacteria-type isocitrate dehydrogenases among Archaea. Multiple alignment of all the available amino acid sequences of di- and multimeric isocitrate dehydrogenases from the three domains of life shows that they can be divided into three distinct phylogenetic groups. Received: 6 February 1997 / Accepted: 12 June 1997  相似文献   

20.
J.B. Hoek  J. Rydstrm  L. Ernster 《BBA》1973,305(3):669-674
A recent claim in the literature (Moyle, J. and Mitchell, P. (1973) Biochem. J. 132, 571–585) that the NAD-dependent isocitrate oxidation observed in extracts of ratliver mitochondria proceeds entirely via the NADP-linked isocitrate dehydrogenase and nicotinamide nucleotide transhydrogenase, and that rat-liver mitochondria contain no NAD-linked isocitrate dehydrogenase, has been examined by using palmityl-CoA as a selective inhibitor of transhydrogenase and ADP as a selective activator of the NAD-linked isocitrate dehydrogenase. The results unambiguously demonstrate that the NAD-dependent oxidation of isocitrate observed under the conditions employed by Moyle and Mitchell proceeds predominantly via the NAD-linked isocitrate dehydrogenase. It is also shown that, by an unfortunate choice of assay conditions, these authors have considerably overestimated the rate of the transhydrogenase reaction.  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