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 21 毫秒
1.
Incubation of malic enzyme (L-malate:NADP+ oxidoreductase (oxaloacetate-decarboxylating), EC 1.1.1.40) with ethoxyformic anhydride caused the time-dependent loss of its ability to catalyze reactions requiring the nucleotide cofactor NADP+ or NADPH, such as the oxidative decarboxylase, the NADP+ - stimualted oxalacetate decarboxylase, the pyruvate reductase, and the pyruvate-medium proton exchange activities. Similar loss of oxidative decarboxylase and pyruvate reductase activities was affected by photo-oxidation in the presence of rose bengal. The inactivation of oxidative decarboxylase activity by ethoxyformic anhydride was accompanied by the reaction of greater than or equal to 2.3 histidyl residues per enzyme site and was strongly inhibited by NADP+. Ethoxyformylation also impaired the ability of malic enzyme to bind NADP+ or NADPH. These results support the involvement of histidyl residue(s) at the nucleotide binding site of malic enzyme.  相似文献   

2.
Reactions catalyzed by NAD-linked malic enzyme from Escherichia coli were investigated. In addition to L-malate oxidative decarboxylase activity (Activity 1) and oxaloacetate decarboxylase activity (Activity 2), the enzyme exhibited oxaloacetate reductase activity (Activity 3) and pyruvate reductase activity (Activity 4). Optimum pH's for Activities 3 and 4 were 4.0 and 5.0, and their specific activities were 1.7 and 0.07, respectively. Upon reaction with N-ethylmaleimide (NEM), Activity 1 decreased following pseudo-first order kinetics. Activity 2 decreased in parallel with Activity 1, while Activities 3 and 4 were about ten-fold enhanced by NEM modification. Modification of one or two sulfhydryl groups per enzyme subunit caused an alteration of the activities. Tartronate, a substrate analog, NAD+, and Mn2+ protected the enzyme against the modification. The Km values for the substrates and coenzymes were not significantly affected by NEM modification. Similarly, other sulfhydryl reagents such as p-hydroxymercuribenzoate (PMB), 5,5'-dithiobis(2-nitrobenzoate) (DTNB), and iodoacetate inhibited the decarboxylase activities and activated the reductase activities to various extents. Modification of the enzyme with PMB or DTNB was reversed by the addition of a sulfhydryl compound such as dithiothreitol or 2-mercaptoethanol. Based on the above results, the mechanism of the alteration of enzyme activities by sulfhydryl group modification is discussed.  相似文献   

3.
G G Chang  R Y Hsu 《Biochemistry》1977,16(2):311-320
Malic enzyme from pigeon liver is alkylated by the substrate analogue bromopyruvate, resulting in the concomitant loss of its oxidative decarboxylase and oxalacetate decarboxylase activities, but not its ability to reduce alpha-keto acids. The inactivation of oxidative decarboxylase activity follows saturation kinetics, indicating the formation of an enzyme-bromopyruvate complex (K congruent to 8 mM) prior to alkylation. The inactivation is inhibited by metal ions and pyridine nucleotide cofactors. Protection of malic enzyme by the substrates L-malate and pyruvate and the inhibitors tartronate and oxalate requires the presence of the above cofactors, which tighten the binding of these carboxylic acids in accord with the ordered kinetic scheme (Hsu, R. Y., Lardy, H. A., and Cleland, W. W. (1967), J. Biol. Chem. 242, 5315-5322). Bromopyruvate is reduced to L-bromolactate by malic enzyme and is an effective inhibitor of L-malate and pyruvate in the overall reaction. The apparent kinetic constants (90 muM-0.8 mM) are one to two orders of magnitude lower than the half-saturation constant (K) of inactivation, indicating a similar tightening of bromopyruvate binding in the E-NADP+ (NADPH)-Mn2+ (Mg2+)-BP complexes. During alkylation, bromopyruvate interacts initially at the carboxylic acid substrate pocket of the active site, as indicated by the protective effect of substrates and the ability of this compound to form kinetically viable complexes with malic enzyme, particularly as a competitive inhibitor of pyruvate carboxylation with a Ki (90 muM) in the same order as its apparent Michaelis constant of 98 muM. Subsequent alkylation of a cysteinyl residue blocks the C-C bond cleavage step. The incorporation of radioactivity from [14C]bromopyruvate gives a half-site stoichiometry of two carboxyketomethyl residues per tetramer, indicating strong negative cooperativity between the four subunits of equal size, or alternatively the presence of structurally dissimilar active sites.  相似文献   

4.
Malic enzyme (S)-malate: NADP+ oxidoreductase (oxaloacetate-decarboxylating, EC 1.1.1.40) purified from the thermoacidophilic archaebacterium Sulfolobus solfataricus, strain MT-4, catalyzed the metal-dependent decarboxylation of oxaloacetate at optimum pH 7.6 at a rate comparable to the decarboxylation of L-malate. The oxaloacetate decarboxylase activity was stimulated about 50% by NADP but only in the presence of MgCl2, and was strongly inhibited by L-malate and NADPH which abolished the NADP activation. In the presence of MnCl2 and in the absence of NADP, the Michaelis constant and Vm for oxaloacetate were 1.7 mM and 2.3 mumol.min-1.mg-1, respectively. When MgCl2 replaced MnCl2, the kinetic parameters for oxaloacetate remained substantially unvaried, whereas the Km and Vm values for L-malate have been found to vary depending on the metal ion. The enzyme carried out the reverse reaction (malate synthesis) at about 70% of the forward reaction, at pH 7.2 and in the presence of relatively high concentrations of bicarbonate and pyruvate. Sulfhydryl residues (three cysteine residues per subunit) have been shown to be essential for the enzymatic activity of the Sulfolobus solfataricus malic enzyme. 5,5'-Dithiobis(2-nitrobenzoic acid), p-hydroxymercuribenzoate and N-ethylmaleimide caused the inactivation of the oxidative decarboxylase activity, but at different rates. The inactivation of the overall activity by p-hydroxymercuribenzoate was partially prevented by NADP singly or in combination with both L-malate and MnCl2, and strongly enhanced by the carboxylic acid substrates; NADP + malate + MnCl2 afforded total protection. The inactivation of the oxaloacetate decarboxylase activity by p-hydroxymercuribenzoate treatment was found to occur at a slower rate than that of the oxidative decarboxylase activity.  相似文献   

5.
Pigeon liver malic enzyme (malate dehydrogenase (decarboxylating), EC 1.1.1.40) was reversibly inactivated by periodate-oxidized NADP in a biphasic manner. The reversibility could be made irreversible by treating the modified enzyme with sodium borohydride. The inactivation showed saturation kinetics and could be prevented by nucleotide (NADP or NADPH). Fully protection was afforded by the combination of NADP, Mn2+ and L-malate. Oxidized NADP was also found to be a coenzyme and noncompetitive inhibitor of L-malate in the oxidative decarboxylase reaction catalyzed by malic enzyme.  相似文献   

6.
NADP-linked malic enzyme [EC 1.1.1.40] was highly purified from Escherichia coli W cells. The purified enzyme was homogeneous as judged by ultracentrifugation and gel electrophoresis. The apparent molecular weights obtained by sedimentation equilibrium analysis, from diffusion and sedimentation constants, and by disc electrophoresis at various gel concentrations were 471,000, 438,000, and 495,000, respectively. The subunit molecular weights obtained by sedimentation equilibrium analysis in the presence of 6 M guanidine hydrochloride and gel electrophoresis in the presence of sodium dodecyl sulfate were 76,000 and 82,000, respectively. The sedimentation coefficient (S(0)20, W) was 13.8S, and the molecular activity was 44,700 min-1 at 30 degrees C. The amino acid composition of the enzyme was determined, and the results were compared with those of NAD-linked malic enzyme from the same organism and those of pigeon liver NADP-linked malic enzyme. The partial specific volume was calculated to be 0.738 ml/g. The Km value for L-malate was 2.3 mM at pH 7.4. Malonate, tartronate, glutarate, and DL-tartrate competitively inhibited the activity. The saturation profile for L-malate exhibited a marked cooperativity in the presence of both chloride ions and acetyl-CoA. However, acetyl-CoA alone did not show cooperativity or produce inhibition in the absence of chloride ions. Vmax and Km were determined as a function of pH. The optimum pH for the reaction was 7.8. Inspection of the Dixon plots suggested that three ionizable groups of the enzyme are essential for the enzyme activity. In addition to the oxidative decarboxylase activity, the enzyme preparation exhibited divalent metal ion-dependent oxaloacetate decarboxylase and alpha-keto acid reductase activities. Based on the above results, the molecular properties of the enzymatic reaction are discussed.  相似文献   

7.
Pigeon liver malic enzyme   总被引:7,自引:0,他引:7  
Summary Malic enzyme of pigeon liver is a tetrameric molecule with identical, or nearly-identical subunits. It catalyzes, in addition to oxidative decarboxylation of L-malate, the following metal activated component reactions: Oxalacetate decarboxylase; reductase with broad specificity on -ketocarboxylic acids; a NADP+-dependent dismutation of L-malate to L-lactate; and proton exchange between pyruvate and medium water. The kinetic mechanism of oxidative decarboxylase is sequential and ordered, with NADP+ adding first to the metal enzyme, followed by L-malate, and by the release of products CO2, pyruvate, and NADPH. NADPH release, or a conformation change preceeding it, is rate-limiting in the overall reaction.Chemical modification studies indicate the presence of histidyl and lysyl residues at the nucleotide site, and tyrosyl residues at the carboxylic acid site. The involvement of protonated histidine(s) in NADPH binding is implicated by results of direct titration experiments, which also suggest a role of this residue as a proton sink in the catalytic reaction.A cysteinyl SH group is located near (but not at) each of the substrate-sites on the enzyme tetramer. Reaction of these groups with SH reagents causes selective loss of activities involving decarboxylation (i.e., oxidative decarboxylase, reductive carboxylase, and oxalacetate decarboxylase), owing to blockage of the reversible carbon-carbon cleavage step by the bulky substituent. All-of-the-sites reactivity is observed for non-specific thiol reagents such as 5,5 dithiobis-(2-nitrobenzoic acid), N-ethylmaleimide, iodoacetate, and iodoacetamide. While bromopyruvate, which is reduced by the enzyme to L-bromolactate under catalytic conditions, alkylates these groups in an active-site directed manner with half-of-the-sites stoichiometry. The remaining two SH groups are reactive toward non-specific reagents, but at rates 2.4 - 3.6 fold lower than do the same groups on the unalkylated enzyme. This behavior is interpreted in terms of the ligand-induced negative cooperativity concept of Koshland, et al. (Biochemistry 5: 365–385, 1966): Reaction of bromopyruvate induces a conformation change on the alkylated subunit which is transmitted to the unoccupied subunit neighbor, turning off its catalytic site for reaction with L-malate, as well as converting the initial fast SH groups into slow, or unreactive SH groups.In equilibrium binding experiments, all-of-the-sites reactivity is seen with nucleotide cofactors NADP+ and NADPH. Binding of Mn2+, or L-malate in the presence of Mn2+ and NADPH is biphasic, showing two tight sites with dissociation constants in the micromolar range, and two weak sites with 10–100 fold lower affinities. The presence of tight and weak L-malate sites is confirmed by fluorescence titration experiments which also yields similar affinities for the substrate molecule. In kinetic studies, two types of non-equivalent, and functionally distinct sites are detected. At saturating NADP+, and Mn2+ and L-malate levels corresponding to binding at tight sites, typical Michaelian behavior is observed. The reaction is inhibited uncompentitively by L-malate at higher concentrations corresponding to occupancy at all of the L-malate sites. Occupancy of Mn2+ at weak metal sites as well has no effect at low L-malate, but prevents substrate inhibition at high L-malate.A tentative half-of-the-sites model consistent with results of chemical modification, binding, and kinetic experiments is proposed for this enzyme. This model implicates involvement of subunit cooperativity in the catalytic process. Malic enzyme is depicted as a tetramer composed of inititally identical subunits, each containing an active-site capable of binding all reactants. Mn2+ and L-malate bind anticooperatively to the tight and weak sites, in contrast to NADP+ which binds equivalently to all sites. On the fully active enzymes, only half (or the tight) of the subunits are simultaneously undergoing catalysis. Binding of L-malate (but not Mn2+) at the adjacent weak subunits causes a slow isomerization of the enzyme, and inhibition of NADPH dissociation from the catalytic subunits. Binding of Mn2+ at the same sites prevents this change and thereby relieving substrate inhibition. This model is further supported by results of active-site titration experiments, such as the half-size burst of enzyme-bound NADPH in the transient state, and half-of-the-sites reactivity of oxalate, an analog for the transition state intermediate of the reaction.Abbreviations DTNB 5,5 dithiobis-(2-nitrobenzoic acid) - NEM N-ethylmaleimide - BP bromopyruvate - DTT dithiothreitol  相似文献   

8.
Bromopyruvate behaves as an active-site-directed inhibitor of the pyruvate decarboxylase (E1) component of the pyruvate dehydrogenase complex of Escherichia coli. It requires the cofactor thiamin pyrophosphate (TPP) and acts initially as an inhibitor competitive with pyruvate (Ki ca. 90 microM) but then proceeds to react irreversibly with the enzyme, probably with the thiol group of a cysteine residue. E1 catalyzes the decomposition of bromopyruvate, the enzyme becoming inactivated once every 40-60 turnovers. Bromopyruvate also inactivates the intact pyruvate dehydrogenase complex in a TPP-dependent process, but the inhibition is more rapid and is mechanistically different. Under these conditions, bromopyruvate is decarboxylated, and the lipoic acid residues in the lipoate acetyltransferase (E2) component become reductively bromoacetylated. Further bromopyruvate then reacts with the new thiol groups thus generated in the lipoic acid residues, inactivating the complex. If reaction with the lipoic acid residues is prevented by prior treatment of the complex with N-ethylmaleimide in the presence of pyruvate, the mode of inhibition reverts to irreversible reaction with the E1 component. In both types of inhibition of E1, reaction of 1 mol of bromopyruvate/mol of E1 chain is required for complete inactivation, and all the evidence is consistent with reaction taking place at or near the pyruvate binding site.  相似文献   

9.
Malic enzyme of duck liver is alkylated by bromopyruvate with half-of-the-sites stoichiometry, and with accompanying loss of oxidative decarboxylase and enhancement of pyruvate reductase activities as was previously shown for the pigeon enzyme (Hsu, R.Y. (1982) Mol. Cell. Biochem. 43, 3-26). In the present work, the alkylated enzyme is shown to bind NADPH, but not L-malate in the presence of MnCl2, indicating impairment of the enzyme site for the substrate and/or divalent metal. The enzyme was differentially labeled by 3-bromo-1-[14C]-pyruvate and digested with TPCK-treated trypsin. Two peptides bearing the susceptible residue were purified by high-performance liquid chromatography and sequenced. Peptide II has the sequence of FMPIVYTPTVGLAXQQYGLAFR, corresponding to residues 86-107 (temporary numbering) of the duck enzyme; cysteine-99(x) is not detected, indicating that it is the target of modification by bromopyruvate. Peptide I is a truncated form of peptide II lacking five amino acid residues at the C-terminal. Cysteine-99 is conserved in malic enzymes from duck, rat, mouse, maize, human, Flaveria trinervia and Bacillus stearothermophilus.  相似文献   

10.
Pigeon liver fatty acid synthetase (FAS) was inactivated irreversibly by stoichiometric concentration of o-phthalaldehyde exhibiting a bimolecular kinetic process. FAS-o-phthalaldehyde adduct gave a characteristic absorption maxima at 337 nm. Moreover this derivative showed fluorescence emission maxima at 412 nm when excited at 337 nm. These results were consistent with isoindole ring formation in which the -SH group of cysteine and epsilon-NH2 group of lysine participate in the reaction. The inactivation is caused by the reaction of the phosphopantetheine -SH group since it is protected by either acetyl- or malonyl-CoA. The enzyme incubated with iodoacetamide followed by o-phthalaldehyde showed no change in fluorescence intensity but decrease in intensity was found in the treatment of 2,4,6-trinitrobenzenesulphonic acid (TNBS), a lysine specific reagent with the enzyme prior to o-phthalaldehyde addition. As o-phthalaldehyde did not inhibit enoyl-CoA reductase activity, so nonessential lysine is involved in the o-phthalaldehyde reaction. Double inhibition experiments showed that 5,5'-dithiobis-(2-nitrobenzoic acid) (DTNB), a thiol specific reagent, binds to the same cysteine which is also involved in the o-phthalaldehyde reaction. Stoichiometric results indicated that 2 moles of o-phthalaldehyde were incorporated per mole of enzyme molecule upon complete inactivation.  相似文献   

11.
Methanosarcina acetivorans, a strictly anaerobic methane-producing species belonging to the domain Archaea, contains a gene cluster annotated with homologs encoding oxidative stress proteins. One of the genes (MA3736) is annotated as a gene encoding an uncharacterized carboxymuconolactone decarboxylase, an enzyme required for aerobic growth with aromatic compounds by species in the domain Bacteria. Methane-producing species are not known to utilize aromatic compounds, suggesting that MA3736 is incorrectly annotated. The product of MA3736, overproduced in Escherichia coli, had protein disulfide reductase activity dependent on a C(67)XXC(70) motif not found in carboxymuconolactone decarboxylase. We propose that MA3736 be renamed mdrA (methanosarcina disulfide reductase). Further, unlike carboxymuconolactone decarboxylase, MdrA contained an Fe-S cluster. Binding of the Fe-S cluster was dependent on essential cysteines C(67) and C(70), while cysteines C(39) and C(107) were not required. Loss of the Fe-S cluster resulted in conversion of MdrA from an inactive hexamer to a trimer with protein disulfide reductase activity. The data suggest that MdrA is the prototype of a previously unrecognized protein disulfide reductase family which contains an intermolecular Fe-S cluster that controls oligomerization as a mechanism to regulate protein disulfide reductase activity.  相似文献   

12.
The effect of Zn2- on nitrate reductase (NR, EC 1.6.6.1) activity was studied in botá wheat (Triticum aestivum cv. Oasis) leaves and in the NR enzyme partially purified from wheat leaves. Leaf segments were floated on 0 to 5 mM ZnSO4 solutions (pH 6.0) for 24 h under continuous light. Zn2- at 250 M decreased NR activity and increased membrane permeability. However, parameters of cellular oxidative damage were scarcely affected by Zn2- treatments. Accordingly, the decrease of NR activity induced by Zn2- was not prevented by benzoate (a scavenger of oxygen radicals). The effect of Zn2- was dependent on leaf age: it decreased NR activity in mature but not in young leaves. Zn2 inhibited the partially purified NR. This inhibition was not reversed by either co- or post-incubation with cysteine, and the amount of -SH groups of the purified NR was not affected by Zn2+ indicating that Zn2- inhibition does not involve key -SH groups of the enzyme. However, o-phenantroline both prevented and reversed Zn2+-induced NR inhibition. We concluded that the effect of Zn2+ on NR activity in vivo is not associated with an increase in active oxygen generation and involves a direct and reversible inhibition of the enzyme.  相似文献   

13.
Solubilization and partial purification of heme oxygenase from rat liver.   总被引:6,自引:0,他引:6  
Hepatic microsomal heme oxygenase was solubilized, partially purified, and characterized from Co2+-treated rats. The enzyme on sodium dodecyl sulfate-polyacrylamide gel electrophoresis exhibited a minimum molecular weight of greater than or equal to 68,000. The solubilized enzyme was totally devoid of contamination with cytochrome P-450 or b5. The requirement for reduced pyridine nucleotides was absolute, and ascorbate could not support heme oxidative activity. However, both TPNH and DPNH could serve as electron donors, with TPNH being more effective. The presence of an appropriate flavoprotein reductase was essential for heme oxidation. The enzyme had an apparent Km of 40 micrometer, a pH optimum of 7.5, and lost substantial activity upon freezing and thawing. Methemoglobin was 30% as effective a substrate for the enzyme as was heme. Free porphyrins could not serve as substrates for the enzyme. The activity of the enzyme was inhibited by HgCl2, p-chloromercuribenzoate, iodoacetamide, mercaptoethanol, and dithiothrietol indicating that free -SH group(s) is necessary for enzyme activity.  相似文献   

14.
Mitochondria from L-1210 mouse ascites tumor show a very high rate of oxidation of L-malate in comparison with mitochondria from normal tissues. They were found to contain large amounts of malic enzyme (E.C.1.1.1.39) catalyzing oxidative decarboxylation of L-malate to pyruvate. Malic enzyme in extracts of tumor mitochondria requires Mn2+ or Mg2+, utilizes either NAD+ or NADP+ as electron acceptor, and shows positive cooperativity in binding of L-malate. These observations suggest that L-1210 tumor mitochondria actively convert excess tricarboxylate cycle intermediates and their precursors into pyruvate for further oxidation.  相似文献   

15.
The pH variation of the kinetic parameters for the oxidative decarboxylation of L-malate and decarboxylation of oxalacetate catalyzed by malic enzyme has been used to gain information on the catalytic mechanism of this enzyme. With Mn2+ as the activator, an active-site residue with a pK of 5.4 must be protonated for oxalacetate decarboxylation and ionized for the oxidative decarboxylation of L-malate. With Mg2+ as the metal, this pK is 6, and, at high pH, V/K for L-malate decreases when groups with pKs of 7.8 and 9 are deprotonated. The group at 7.8 is a neutral acid (thought to be water coordinated to Mg2+), while the group at 9 is a cationic acid such as lysine. The V profile for reaction of malate shows these pKs displaced outward by 1.4 pH units, since the rate-limiting step is normally TPNH release, and the chemical reaction, which is pH sensitive, is 25 times faster. TPN binding is decreased by ionization of a group with pK 9.3 or protonation of a group with pK 5.3. The pH variation of the Km for Mg shows that protonation of a group with pK 8.7 (possibly SH) decreases metal binding in the presence of malate by a factor of 1400, and in the absence of malate by a factor of 20. A catalytic mechanism is proposed in which hydride transfer is accompanied by transfer of a proton to the group with pK 5.4-6, and enolpyruvate is protonated by water coordinated to the Mg2+ (pK 7.8) after decarboxylation and release of CO2.  相似文献   

16.
The structural requirements of the NADP+ molecule as a coenzyme in the oxidative decarboxylation reaction catalysed by pigeon liver malic enzyme were studied by kinetic and fluorimetric analyses with various NADP+ analogues and fragments. The substrate L-malate had little effect on the nucleotide binding. Etheno-NADP+, 3-acetylpyridine-adenine dinucleotide phosphate, and nicotinamide-hypoxanthine dinucleotide phosphate act as alternative coenzymes for the enzyme. Their kinetic parameters were similar to that of NADP+. Thionicotinamide-adenine dinucleotide phosphate, 3-aminopyridine-adenine dinucleotide phosphate, 5'-adenylyl imidodiphosphate, nicotinamide-adenine dinucleotide 3'-phosphate and NAD+ act as inhibitors for the enzyme. The first two were competitive with respect to NADP+ and non-competitive with respect to L-malate; the other inhibitors were non-competitive with NADP+. All NADP+ fragments were inhibitory to the enzyme, with a wide range of affinity, depending on the presence or absence of a 2'-phosphate group. Compounds with this group bind to the enzyme 2-3 orders of magnitude more tightly than those without this group. Only compounds with this group were competitive inhibitors with respect to NADP+. We conclude that the 2'-phosphate group is crucial for the nucleotide binding of this enzyme, whereas the carboxyamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity. There is a strong synergistic effect between the binding of the nicotinamide and adenosine moieties of the nucleotide molecule.  相似文献   

17.
Pig heart TPN-dependent isocitrate dehydrogenase is inactivated by reaction with 5,5′-dithiobis (2-nitrobenzoic acid) (DTNB). The dependence of the rate constant for inactivation on the reagent concentration is nonlinear, and can be analyzed in terms of the existence of two mechanisms for reaction with the enzyme, one involving reversible binding prior to inactivation and the other a bimolecular reaction. Cyanide reacts with the inactive modified enzyme to yield thiocyano-isocitrate dehydrogenase without increasing the catalytic activity; this result suggests that inactivation by DTNB is not due to steric hindrance by the bulky thionitrobenzoate group bound to the enzyme. The inactive thiocyano enzyme binds manganous ion normally. In contrast to its effect on native enzyme, however, isocitrate does not strengthen the binding of Mn2+ to the thiocyano enzyme; the tightened binding of manganous-isocitrate may be critical for the catalytic activity of the enzyme. Protection against inactivation by DTNB is provided by isocitrate plus the activator, manganous ion, or the competitive inhibitor, calcium ion. The concerted inhibitors oxalacetate and glyoxylate, when present together with Mn2+ and TPN, also protect against loss of activity. A marked decrease in the inactivation rate constant to a finite limiting value is caused by saturating concentrations of TPNH and Mn2+, indicating that these ligands do not bind directly at the sites attacked by DTNB. The number of cysteine residues which react with DTNB concomitant with inactivation depends on the ligands present in the reaction mixture. In all cases, the equivalent of one -SH reacts without affecting activity. In the presence of Mn2+ and α-ketoglutarate, which do not appreciably affect the inactivation rate, loss of activity is proportional to reaction with two -SH groups. These results suggest that the integrity of a maximum of two cysteine residues is essential for the function of the pig heart isocitrate dehydrogenase, and that at least one cysteine residue may be located within the manganous-isocitrate binding site.  相似文献   

18.
We have partially purified the CO2 reductase, present in green potato tuber chloroplasts, as a latent form. Illumination of the chloroplasts in the absence of substrate, bicarbonate, activated the enzyme, which could then be obtained in soluble forms. Purification of the enzyme was achieved by (NH4)2SO4 fractionation (0-30%) and adsorption and elution from a DEAE-Sephadex A-50 column. The final preparation showed 15-fold purification and 50% recovery of the activity. The pH optimum for CO2 reductase was 8.0. Hepes and Tricine buffers showed maximum activity whereas Tris/phosphate or borate failed to show any activity. The enzyme reaction was sensitive to the presence of metal ions like Fe3+, Hg2+, Cu2+, Mo6+ and Zn2+, however, a threefold activation was observed with Fe2+. The metal requirement for CO2 reductase was evident from the observed inhibition by metal chelators like o-phenanthroline, alpha, alpha'-dipyridyl, bathocuproine, 8-hydroxyquinoline etc. Out of these o-phenanthroline was the strongest inhibitor and its concentration for 50% inhibition was 40 microM. The presence of Fe2+ ions in the reaction mixture protected the enzyme from heat denaturation upto 50 degrees C. Maximum enzyme activity was observed at 15 degrees C. The enzyme activity showed a 30-s lag period and the maximum was reached in 90 s. Supplementation of sodium dithionite in the reaction activated enzyme activity threefold, suggesting involvement of dithiol groups in the catalytic activity. There was strong inhibition by -SH inhibitors like 5,5'-dithiobis(2-nitrobenzoic acid) and N-ethylmaleimide and -SH reagents like dithiothreitol, 2-mercaptoethanol and cysteine. Various nucleotide coenzyme tried inhibited the enzyme strongly.  相似文献   

19.
Bromopyruvate was shown to inhibit E. coli glutamate decarboxylase competitively with respect to L-glutamate. High concentrations of bromopyruvate caused a time-dependent inactivation of glutamate decarboxylase. However, the apoenzyme was rapidly and irreversibly inactivated by bromopyruvate with an inactivation constant of 490 1 mole?1 min?1 at pH 5.7. Studies with labeled bromopyruvate indicated that approximately 1.7 moles of inhibitor were bound per subunit of apoenzyme.  相似文献   

20.
Fatty acid synthetase from goose uropygial gland was inactivated by treatment with pyridoxal 5′-phosphate. Malonyl-CoA and acetyl-CoA did not protect the enzyme whereas NADPH provided about 70% protection against this inactivation. 2′-Monophospho-ADP-ribose was nearly as effective as NADPH while 2′-AMP, 5′-AMP, ADP-ribose, and NADH were ineffective suggesting that pyridoxal 5′-phosphate modified a group that interacts with the 5′-pyrophosphoryl group of NADPH and that the 2′-phosphate is necessary for the binding of the coenzyme to the enzyme. Of the seven component activities catalyzed by fatty acid synthetase only the enoyl-CoA reductase activity was inhibited. Inactivation of both the overall activity and enoyl-CoA reductase of fatty acid synthetase by this compound was reversed by dialysis or dilution but not after reduction with NaBH4. The modified protein showed a characteristic Schiff base absorption (maximum at 425 nm) that disappeared on reduction with NaBH4 resulting in a new absorption spectrum with a maximum at 325 nm. After reduction the protein showed a fluorescence spectrum with a maximum at 394 nm. Reduction of pyridoxal phosphate-treated protein with NaB3H4 resulted in incorporation of 3H into the protein and paper chromatography of the acid hydrolysate of the modified protein showed only one fluorescent spot which was labeled and ninhydrin positive and had an Rf identical to that of authentic N6-pyridoxyllysine. When [4-3H]pyridoxal phosphate was used all of the 3H, incorporated into the protein, was found in pyridoxyllysine. All of these results strongly suggest that pyridoxal phosphate inhibited fatty acid synthetase by forming a Schiff base with the ?-amino group of lysine in the enoyl-CoA reductase domain of the enzyme. The number of lysine residues modified was estimated with [4-3H]pyridoxal-5′-phosphate/NaBH4 and by pyridoxal-5′-phosphate/NaB3H4. Scatchard analysis showed that modification of two lysine residues per subunit resulted in complete inactivation of the overall activity and enoyl-CoA reductase of fatty acid synthetase. NADPH prevented the inactivation of the enzyme by protecting one of these two lysine residues from modification. The present results are consistent with the hypothesis that each subunit of the enzyme contains an enoyl-CoA reductase domain in which a lysine residue, at or near the active site, interacts with NADPH.  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