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 328 毫秒
1.
Isocitrate lyase of germinating castor seed endosperm catalyzes the reactions of succinate and of isocitrate (but not of glyoxylate) with tetranitromethane (TNM), giving rise to the nitroform anion (C-(NO2)3), analogous to the reaction of TNM with carbanions (O.P. Malhotra and U.N. Dwivedi, 1984, Ind. J. Biochem. Biophys. 21, 65-67). The kinetics of this reaction have been investigated under a variety of conditions. At a fixed TNM concentration, the initial rate of reaction exhibits a hyperbolic saturation of the enzyme with isocitrate. The reaction with succinate, however, shows "negative cooperativity" in succinate saturation and the data are consistent with the existence of two sets of succinate binding sites of unequal affinity ("tight" and "loose" sites). Equal reaction rates are observed at enzyme-saturating concentrations of succinate and isocitrate. In every case, the rate of reaction is proportional to the TNM concentration. In the presence of alpha-ketoglutarate, hyperbolic saturation curves are obtained for all the substrates (TNM and succinate or TNM and isocitrate). In the presence of this effector the Km of succinate and TNM are independent of the concentration of the second substrate. On the other hand, sets of parallel straight lines are obtained in the double-reciprocal plots for the enzymatic reaction of TNM with isocitrate in the presence of alpha-ketoglutarate. Studies on the effect of pH on the isocitrate lyase-catalyzed reactions of TNM with succinate, TNM with isocitrate, and succinate with glyoxylate in the absence as well as in the presence of alpha-ketoglutarate show that the proton behaves as an uncompetitive inhibitor in all these reactions, suggesting the presence of a "masked" basic group at the enzyme site, which is protonated in the presence of substrate only. The pKa value of this group lies in the range 6.7-6.9. The enzymatic reactions of TNM with succinate and isocitrate exhibit identical Mg2+ ion dependence. From a comparison of the data on the enzymatic reactions of TNM with the corresponding results on the physiological reaction catalyzed by this enzyme, it has been suggested that an ion pair intermediate (E+ X S-, in which E, S, and S- stand for enzyme, succinate, and succinate carbanion, respectively) lies on the pathway of catalysis by isocitrate lyase.  相似文献   

2.
Y H Ko  P Vanni  G R Munske  B A McFadden 《Biochemistry》1991,30(30):7451-7456
The inactivation of tetrameric 188-kDa isocitrate lyase from Escherichia coli at pH 6.8 (37 degrees C) by diethyl pyrocarbonate, exhibiting saturation kinetics, is accompanied by modification of histidine residues 266 and 306. Substrates isocitrate, glyoxylate, or glyoxylate plus succinate protect the enzyme from inactivation, but succinate alone does not. Removal of the carbethoxy groups from inactivated enzyme by treatment with hydroxylamine restores activity of isocitrate lyase. The present results suggest that the group-specific modifying reagent diethyl pyrocarbonate may be generally useful in determining the position of active site histidine residues in enzymes.  相似文献   

3.
Alkylation of isocitrate lyase from Escherichia coli by 3-bromopyruvate   总被引:5,自引:0,他引:5  
The inactivation of tetrameric isocitrate lyase from Escherichia coli by 3-bromopyruvate, exhibiting saturation kinetics, is accompanied by the loss of one sulfhydryl per subunit. The substrates glyoxylate and isocitrate protect against inactivation whereas the substrate succinate does not. The modification by 3-bromopyruvate (equimolar to subunits) imparts striking resistance to digestion of isocitrate lyase by trypsin, chymotrypsin, and V8 protease as well as a major decrease in the intensity of tryptophan fluorescence. After alkylation, the sequence Gly-His-Met-Gly-Gly-Lys is found following the modified Cys residue in the tryptic peptide representing positions 196-201. Thus Cys195 is alkylated by 3-bromopyruvate.  相似文献   

4.
Kinetic analysis of inactivation of isocitrate lyase from Pseudomonas indigofera by 3-bromopyruvate established that enzyme binds this compound prior to alkylation and that substrate, Ds-isocitrate, competes for the same site on the enzyme. The rate of inactivation was increased by EDTA which is a promoter of catalysis in the presence of activated (reduced) enzyme and substrate. The combination of products, glyoxylate plus succinate, also protected against inactivation. Glyoxylate plus itaconate, phosphoenolpyruvate, or maleate also protected. However, each of the latter three compounds or glyoxylate or succinate alone provided little or no protection. Pyruvate, a competitive inhibitor with respect to glyoxylate in the condensation reaction, also failed to protect. However, two dicarboxylates, meso-tartrate and oxalate, that are also competitive inhibitors with respect to glyoxylate provide some protection against inactivation by BrP perhaps by bridging across cationic sites that facilitate glyoxylate and succinate binding. These and other results imply that alkylation by 3-bromopyruvate occurs at the succinate part of the active site. A mechanism which includes a catalytic role for the cysteine residue at the active site is presented and discussed.  相似文献   

5.
Isocitrate lyase was purified from Phycomyces blakesleeanus N.R.R.L. 1555(-). The native enzyme has an Mr of 240,000. The enzyme appeared to be a tetramer with apparently identical subunits of Mr 62,000. The enzyme requires Mg2+ for activity, and the data suggest that the Mg2(+)-isocitrate complex is the true substrate and that Mg2+ ions act as a non-essential activator. The kinetic mechanism of the enzyme was investigated by using product and dead-end inhibitors of the cleavage and condensation reactions. The data indicated an ordered Uni Bi mechanism and the kinetic constants of the model were calculated. The spectrophotometric titration of thiol groups in Phycomyces isocitrate lyase with 5.5'-dithiobis-(2-nitrobenzoic acid) gave two free thiol groups per subunit of enzyme in the native state and three in the denatured state. The isocitrate lyase was completely inactivated by iodoacetate, with non-linear kinetics. The inactivation data suggest that the enzyme has two classes of modifiable thiol groups. The results are also in accord with the formation of a non-covalent enzyme-inhibitor complex before irreversible modification of the enzyme. Both the equilibrium constants for formation of the complex and the first-order rate constants for the irreversible modification step were determined. The partial protective effect of isocitrate and Mg2+ against iodoacetate inactivation was investigated in a preliminary form.  相似文献   

6.
The cleavage of Ds-isocitrate catalyzed by isocitrate lyase from Linum usitatissimum results in the ordered release of succinate and glyoxylate. The glyoxylate analog 3-bromopyruvate irreversibly inactivates the flax enzyme in a process exhibiting saturation kinetics and protection by glyoxylate or isocitrate or the competitive inhibitor l-tartrate. Succinate provides considerably less protection. Results with 3-bromopyruvate suggest that this reagent modifies plant and prokaryotic isocitrate lyases differently. Treatment of the tetrameric 264,000-dalton flax enzyme with carboxypeptidase A results in a release of one histidine/subunit which is concordant with loss of activity. The only N-terminal residue is methionine. Treatment of flax enzyme with diethylpyrocarbonate at pH 6.5 selectively modifies two histidines per 67,000-dalton subunit. The reaction of one histidine residue is abolished by the binding of l-tartrate and the modification of one is coincident with inactivation. The carboxy-terminal and active-site modifications establish that one histidine residue/monomer is essential in the flax enzyme and considerably extend information heretofore available only for fungal and bacterial isocitrate lyase.  相似文献   

7.
Isocitrate lyase from the mycelium of Phycomyces blakesleeanus was inactivated with thiol-reactive reagents, 5,5'-dithiobis-(2-nitrobenzoic)acid, p-hydroxymercuribenzoic acid, N-ethylmaleimide or iodoacetate, at pH 6.8 and 25 degrees C. In all cases the inactivation is characterized by a biphasic kinetic profile. The rapid initial phase of inactivation does not increase linearly with increasing reagent concentration, but exhibits an apparent saturation effect, suggesting the formation of a reversible complex between the enzyme and the reagent prior to the inactivation step. Re-activation of the enzyme was observed under thiol excess treatment. The pH dependence of the initial phase of inactivation suggests that a group on the enzyme with pKa = 6.8 is being modified. The effect of ligands was tested on the inactivation reaction. Mg(2+)-Ds-isocitrate and Ds-isocitrate provided total protection, whereas Mg2+ ions, succinate and oxalate provided only partial protection of the enzyme against inactivation. On the basis of these results, we would suggest that the thiol-reactive reagents modify at least one thiol group crucial for the enzymatic activity and probably located in the interface between succinate and glyoxylate subsite.  相似文献   

8.
Vanadate was used as a substrate analogue to modify and subsequently localize active site serine residues of isocitrate lyase from Escherichia coli. Irradiation of the enzyme on ice with UV light in the presence of vanadate resulted in inactivation. Inactivation was prevented by the substrates glyoxylate or Ds-isocitrate and to a much lesser extent by succinate. Reduction of photoinactivated isocitrate lyase by NaBH4 partially restored enzyme activity. The photomodified enzyme was labeled by reduction with NaB[3H]4 in the presence and absence of the substrates succinate plus glyoxylate. Highly differential labeling of serine residues 319 and 321 in the absence of substrates suggests their importance in the action of isocitrate lyase. These residues are highly conserved in all five known sequences of this enzyme.  相似文献   

9.
The storage stability of bilirubin oxidase was studied in water-in-oil CTAB microemulsions with a chloroformrich continuous organic phase. The kinetics of the inactivation process were best described by a double exponential equation. Approximately half of enzymatic activity was lost during a "fast" phase with a half life of ca. 50 min, whereas the remaining activity was lost much more slowly (half life ca. 1000 min). Rates of inactivation were not affected significantly by variation of either solvent composition or concentration of water droplets, but inactivation was more rapid when droplet size was very small. Steady-state enzyme kinetics were studied at various stages in the inactivation process, and it was shown that inactivation occurred without change in the K(m) of the enzyme for bilirubin. Stability was also studied in a liquid/solid two-phase system; it was found that the inactivation process in this system; it was found that the inactivation process in this system was best described by a single exponential term. The rate was similar to the "fast" phase rate observed in the water-in-oil microemulsion system. Inactivation of the enzyme slow. Addition of the surfactant CTAB to the aqueous environment increased the rate of inactivation to levels comparable to those of the "slow" phase observed in water-in-oil microemulsions. (c) 1993 Wiley & Sons, Inc.  相似文献   

10.
1. The enzymes citrate lyase and isocitrate lyase catalyse similar reactions in the cleavage of citrate to acetate plus oxaloacetate and of isocitrate to succinate plus glyoxylate, respectively. 2. Nevertheless, the mechanism of action of each enzyme appears to be different from each other. Citrate lyase is an acyl carrier protein-containing enzyme complex whereas isocitrate lyase is not. The active form of citrate lyase is an acetyl-S-enzyme but that of isocitrate lyase is not a corresponding succinyl-S-enzyme. 3. In contrast to citrate lyase, the isocitrate enzyme is not inhibited by hydroxylamine nor does it acquire label if treated with appropriately labelled radioactive substrate. 4. Isotopic exchange experiments performed in H18-2O with isocitrate as a substrate produced no labelling in the product succinate. This was shown by mass-spectrometric analysis. 5. The conclusion drawn from these results is that no activation of succinate takes place on the enzyme through transient formation of succinic anhydride or a covalently-linked succinyl-enzyme, derived from this anhydride.  相似文献   

11.
Creatine kinase from beef heart mitochondria is inactivated by 2,3-butanedione. The kinetics of inactivation of the mitochondrial enzyme is biphasic with a bend at a point corresponding to 50% inactivation. The inactivation rate constants of the first fast and the second slow phases of the reaction differ by one order of magnitude, thus suggesting the existence of two types of arginine residues, i.e. "fast" and "slow" ones, with different reactivities. The inactivation rate constant of the slow phase is very close to that for cytoplasmic creatine kinase. At saturating concentrations MgATP and MgADP afford complete protection of the slow phase of inactivation. It is assumed that the "slow" arginine is involved in the binding of metal-nucleotide substrates in the enzyme active center.  相似文献   

12.
In Aspergillus nidulans, activity of the glyoxylate cycle enzyme isocitrate lyase is finely regulated. Isocitrate lyase is induced by growth on C2 compounds and long-chain fatty acids and repressed by glucose. In addition, activity of isocitrate lyase is subject to a second mechanism of catabolite control, glucose-induced inactivation. Here, we demonstrate that the catabolite inactivation of A. nidulans isocitrate lyase, a process that takes place during glucose adaptation of cells grown under gluconeogenic conditions, occurs by proteolysis of the enzyme. Ultrastructural analyses were carried out in order to investigate the cellular processes that govern the catabolite inactivation of this peroxisomal enzyme. Addition of glucose to oleate-induced cells triggered the specific engulfment and sequestration of peroxisomes by the vacuoles. Sequestration of various peroxisomes by a single vacuole was a frequently observed phenomenon. Results obtained by immunoelectron microscopy using antibodies against A. nidulans isocitrate lyase showed that degradation of this peroxisomal enzyme occurred inside the vacuole. In addition, ultrastructural studies demonstrated that microautophagy was the autophagic pathway involved in degradation of redundant peroxisomes during glucose adaptation of oleate-induced cells of A. nidulans.  相似文献   

13.
The substrate analogue 3-bromo-2-ketoglutarate reacts with pig heart NADP+-dependent isocitrate dehydrogenase to yield partially inactive enzyme. Following 65% inactivation, no further inactivation was observed. Concomitant with this inactivation, incorporation of 1 mol of reagent/mol of enzyme dimer was measured. The dependence of the inactivation rate on bromoketoglutarate concentration is consistent with reversible binding of reagent (KI = 360 microM) prior to irreversible reaction. Manganous isocitrate reduces the rate of inactivation by 80% but does not provide complete protection even at saturating concentrations. Complete protection is obtained with NADP+ or the NADP+-alpha-ketoglutarate adduct. By modification with [14C]bromoketoglutarate or by NaB3H4 reduction of modified enzyme, a single major radiolabeled tryptic peptide was obtained by high performance liquid chromatography with the sequence: Asp-Leu-Ala-Gly-X-Ile-His-Gly-Leu-Ser-Asn-Val-Lys. Evidence in the following paper (Bailey, J.M., Colman, R.F. (1987) J. Biol. Chem. 262, 12620-12626) indicates that X is glutamic acid. Enzyme modified at the coenzyme site by 2-(bromo-2,3-dioxobutylthio)-1,N(6)-ethenoadenosine 2',5'-biphosphate in the presence of manganous isocitrate is not further inactivated by bromoketoglutarate. Bromoketoglutarate-modified enzyme exhibits a stoichiometry of binding isocitrate and NADPH equal to 1 mol/mol of enzyme dimer, half that of native enzyme. These results indicate that bromoketoglutarate modifies a residue in the nicotinamide region of the coenzyme site proximal to the substrate site and that reaction at one catalytic site of the enzyme dimer decreases the activity of the other site.  相似文献   

14.
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   

15.
Nocardia salmonicolor, grown on acetate, commercial D,L-lactate or hydrocarbon substrates, has high isocitrate lyase activities compared with those resulting from growth on other carbon sources. This presumably reflects the anaplerotic role of the glyoxylate cycle during growth on the former substrates. Amongst a variety of compounds tested, including glucose, pyruvate and tricarboxylic acid cycle intermediates, only succinate and fumarate prevented an increase in enzyme activity in the presence of acetate. When acetate (equimolar to the initial sugar concentration) was added to cultures growing on glucose, there followed de novo synthesis of isocitrated lyase and isocitrate dehydrogenase, with increases in growth rate and glucose utilization, and both acetate and glucose were metabolized simultaneously. A minute amount of acetate (40 muM) caused isocitrate lyase synthesis (a three-fold increase in activity within 3 min of addition) when added to glucose-limited continuous cultures, but even large amounts added to nitrogen-limited batch cultures were ineffective. Malonate, at a concentration that was not totally growth-inhibitory (1mM) prevented the inhibition of acetate-stimulated isocitrate lyase synthesis by succinate, but fumarate still inhibited in the presence of malonate. Phosphoenolpyruvate is a non-competitive inhibitor of the enzyme (apparent Ki 1-7 mM). The results are consistent with the induction of isocitrate or a closely related metabolite, and catabolite repression by a C-4 acid of the tricarboxylic acid cycle, possibly fumarate.  相似文献   

16.
Escherichia coli NZN111, which lacks activities for pyruvate-formate lyase and lactate dehydrogenase, and AFP111, a derivative which contains an additional mutation in ptsG (a gene encoding an enzyme of the glucose phophotransferase system), accumulate significant levels of succinic acid (succinate) under anaerobic conditions. Plasmid pTrc99A-pyc, which expresses the Rhizobium etli pyruvate carboxylase enzyme, was introduced into both strains. We compared growth, substrate consumption, product formation, and activities of seven key enzymes (acetate kinase, fumarate reductase, glucokinase, isocitrate dehydrogenase, isocitrate lyase, phosphoenolpyruvate carboxylase, and pyruvate carboxylase) from glucose for NZN111, NZN111/pTrc99A-pyc, AFP111, and AFP111/pTrc99A-pyc under both exclusively anaerobic and dual-phase conditions (an aerobic growth phase followed by an anaerobic production phase). The highest succinate mass yield was attained with AFP111/pTrc99A-pyc under dual-phase conditions with low pyruvate carboxylase activity. Dual-phase conditions led to significant isocitrate lyase activity in both NZN111 and AFP111, while under exclusively anaerobic conditions, an absence of isocitrate lyase activity resulted in significant pyruvate accumulation. Enzyme assays indicated that under dual-phase conditions, carbon flows not only through the reductive arm of the tricarboxylic acid cycle for succinate generation but also through the glyoxylate shunt and thus provides the cells with metabolic flexibility in the formation of succinate. Significant glucokinase activity in AFP111 compared to NZN111 similarly permits increased metabolic flexibility of AFP111. The differences between the strains and the benefit of pyruvate carboxylase under both exclusively anaerobic and dual-phase conditions are discussed in light of the cellular constraint for a redox balance.  相似文献   

17.
Inhibition of isocitrate lyase from Pseudomonas indigofera by itaconate   总被引:2,自引:0,他引:2  
The effect of the inhibitor itaconate on the activity of purified isocitrate lyase from Pseudomonas indigofera was examined for the reaction in both directions. Itaconate was found to equilibrate very slowly with its enzyme-bound form, so that a rapid change in itaconate concentration produced a gradual change in reaction velocity which eventually reached a new steady state. Kinetic studies of this relaxation phenomenon indicated that itaconate inhibited by binding the enzyme only after prior binding of glyoxylate, thus mimicking the kinetic behavior of succinate. On the basis of these studies, the dissociation constants for itaconate and glyoxylate from their respective enzyme-bound forms were calculated. More than half of the isocitrate lyase was complexed by glyoxylate during cleavage of saturating isocitrate. The rate constant for release of itaconate from the enzyme was calculated to be about 0.2 min?1. Direct binding of [14C]itaconate and [14C]succinate to isocitrate lyase at pH 6.8 was measured. Some binding of both ligands was found in the absence of glyoxylate, which was stimulated by the presence of 1 mm glyoxylate. These results suggest that there are up to three or more binding sites per active subunit, but that only one of these is catalytic.  相似文献   

18.
The interaction of clostridiopeptidase of Clostridium histolyticum with EDC, TNM and MA, the specific reagents for COOH-groups, tyrosine and lysine residues was studied. It was shown that at pH 6.0 EDC inactivates the enzyme. The inactivation process follows the pseudo-first order kinetics and is described by a second order rate constant equal to 1 M-1 min-1. The synthetic substrate does not prevent, in practical terms, the enzyme inactivation by EDC. At pH 8.0 TNM modifies about 19 tyrosine residues in the clostridiopeptidase molecule which is accompanied by marked inhibition of the enzyme activity (down to 70-90%). In this case, the inactivation process is not described by simple pseudo-first order kinetics but is characterized by two steps (fast and slow) with second order rate constants of approximately 14 and 3.5 M-1 min-1, respectively. The synthetic substrate partly prevents the inactivation of the enzyme by TNM and protects 11 tyrosine residues. The MA-induced incorporation of 13 +/- 3 maleyl groups into the clostridiopeptidase molecule in partially prevented by the synthetic substrate with protects the enzyme against inactivation. The data obtained suggest that lysine residues are seemingly included into the active center of clostridiopeptidase, whereas tyrosine residues provide for the maintenance of active conformation of the enzyme.  相似文献   

19.
During growth on succinate, Acinetobacter calcoaceticus contains two forms of the enzyme isocitrate dehydrogenase. Addition of acetate to a lag-phase culture grown on succinate causes a dramatic increase in activity of form II of isocitrate dehydrogenase and in isocitrate lyase. Form II of isocitrate dehydrogenase may be responsible for the partition of isocitrate between the TCA cycle and the glyoxylate by-pass. This report describes the phosphorylation of the enzyme isocitrate lyase from A. calcoaceticus. This phosphorylation may be a regulatory mechanism for the glyoxylate by-pass.  相似文献   

20.
Key enzymes of the glyoxylate cycle, isocitrate lyase and malate synthase, were identified in pupas of the butterfly Papilio machaon L. The activities of these enzymes in pupas were 0.056 and 0.108 unit per mg protein, respectively. Isocitrate lyase was purified by a combination of various chromatographic steps including ammonium sulfate fractionation, ion-exchange chromatography on DEAE-Toyopearl, and gel filtration. The specific activity of the purified enzyme was 5.5 units per mg protein, which corresponded to 98-fold purification and 6% yield. The enzyme followed Michaelis-Menten kinetics (Km for isocitrate, 1.4 mM) and was competitively inhibited by succinate (Ki = 1.8 mM) and malate (Ki = 1 mM). The study of physicochemical properties of the enzyme showed that it is a homodimer with a subunit molecular weight of 68 +/- 2 kD and a pH optimum of 7.5 (in Tris-HCl buffer).  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