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 15 毫秒
1.
Incubation of pig heart NADP-dependent isocitrate dehydrogenase with ethoxyformic anhydride (diethylpyrocarbonate) at pH 6.2 results in a 9-fold greater rate of loss of dehydrogenase than of oxalosuccinate decarboxylase activity. The rate constants for loss of dehydrogenase and decarboxylase activities depend on the basic form of ionizable groups with pK values of 5.67 and 7.05, respectively, suggesting that inactivation of the two catalytic functions results from reaction with different amino acid residues. The rate of loss of dehydrogenase activity is decreased only slightly in the presence of manganous isocitrate, but is reduced up to 10-fold by addition of the coenzymes or coenzyme analogues, such as 2'-phosphoadenosine 5'-diphosphoribose (Rib-P2-Ado-P). Enzyme modified at pH 5.8 fails to bind NADPH, but exhibits manganese-enhanced isocitrate binding typical of native enzyme, indicating that reaction takes place in the region of the nucleotide binding site. Dissociation constants for enzyme . coenzyme-analogue complexes have been calculated from the decrease in the rate of inactivation as a function of analogue concentration. In the presence of isocitrate, activating metals (Mn2+, Mg2+, Zn2+) decrease the Kd value for enzyme . Rib-P2-Ado-P, while the inhibitor Ca2+ increases Kd. The strengthened binding of nucleotide produced by activating metal-isocitrate complexes may be essential for the catalytic reaction, reflecting an optimal orientation of NADP+ to facilitate hydride transfer. Measurements of ethoxyformyl-histidine formation at 240 nm and of incorporation of [14C]ethoxy groups in the presence and absence of Rib-P2-Ado-P indicate that loss of activity may be related to modification of approximately one histidine. The critical histidine appears to be located in the nucleotide binding site in a region distal from the substrate binding site.  相似文献   

2.
R S Ehrlich  R F Colman 《Biochemistry》1975,14(22):5008-5016
The binding of TPNH to native and chemically modified pig heart TPN-dependent isocitrate dehydrogenase was studied by the techniques of ultrafiltration and fluorescence enhancement. A single site (per peptide chain) was found for TPNH with a dissociation constant (KD = 1.45 muM) that is quantitatively comparable to the Michaelis constant. The oxidized coenzyme, TPN+, weakens the binding of TPNH. The substrate manganous isocitrate also inhibits the binding of TPNH and, reciprocally, TPNH inhibits the binding of manganous isocitrate, suggesting that binding to the reduced coenzyme and substrate sites is mutually exclusive. Ultrafiltration experiments with carbonyl [14C]TPN+ revealed the existence of two sites with a dissociation constant (49 muM) more than ten times higher than the Michaelis constant. This observation excludes a random mechanism for isocitrate dehydrogenase or a sequential mechanism in which TPN+ binds first. Four chemically modified isocitrate dehydrogenases have been prepared: enzyme inactivated by reaction of a single methionyl residue with iodoacetate, by modification of a glutamyl residue by glycinamide (in the presence of a water soluble carbodiimide), by reaction of four cysteines successively with 5,5'-dithiobis(2-nitrobenzoic acid) and potassium cyanide, or by addition of two cysteine residues to N-ethylmaleimide. These enzymes were tested for their ability to bind TPN+, TPNH, and manganous isocitrate. In the cases of the cysteinyl and glutamyl-modified enzymes, inactivation appears to be due primarily to loss of the ability to bind the substrate manganous isocitrate. In constrast, the methionyl residue may participate in the coenzyme binding site or, more likely, may be involved in a step in catalysis subsequent to binding.  相似文献   

3.
Catalytic reaction of the 2', 3'-dialdehyde analog of TPN (oTPN) with pig heart TPN-dependent isocitrate dehydrogenase in the presence of the substrate manganous isocitrate results in the formation of the dialdehyde derivative of TPNH (oTPNH). In the absence of the substrate, modification by oTPN leads to a progressive inactivation of the enzyme. The dependence of the pseudo-first order rate constants on the reagent concentration indicates the formation of a reversible complex with the enzyme prior to covalent modification (kmax = 5.5 X 10(-2) min-1; K1 = 290 microM). Reaction of [14C]oTPN with the enzyme results in the incorporation of 2 mol of oTPN/mol of peptide chain. No appreciable protection against either inactivation or incorporation by the natural ligands TPN and TPNH was obtained, suggesting different modes of binding of the analog in the presence and absence of the substrate isocitrate. Enzymatically synthesized oTPNH has been isolated and demonstrated to act as an affinity label for a TPNH-binding site of isocitrate dehydrogenase. The inactivation process exhibits saturation kinetics (kmax = 2.67 X 10(-3) min-1; K1 = 33 microM). Protection against activity loss, as well as a decrease in incorporation from 2 to 1 eq of [14C]oTPNH bound/peptide chain was observed in the presence of 1 mM TPNH. From the TPNH concentration dependence of the inactivation rate by oTPNH, a dissociation constant of 3.4 microM is calculated for TPNH, indicating binding of the analog to a specific TPNH-binding site on the enzyme. Although dialdehyde derivatives are frequently assumed to form Schiff bases with proteins, the evidence presented suggests the formation of morpholino derivatives as the products of the covalent reaction of isocitrate dehydrogenase with the dialdehyde derivatives of TPN and TPNH. The new reagent, oTPNH, may serve as an affinity label for other dehydrogenases.  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.
V G Neef  F M Huennekens 《Biochemistry》1976,15(18):4042-4047
The 1,N6-ethenoadenine derivatives of triphosphopyridine and reduced triphosphopyridine nucleotides (TPN and TPNH) epsilon-TPN and epsilon-TPNH) have been synthesized and used as fluorescent probes to examine the pyridine nucleotide binding site of L1210 dihydrofolate reductase. Epsilon-TPNH (Km = 16.7 muM) was able to replace TPNH (Km = 3.8 muM) in the enzyme-catalyzed reduction of dihyrdofolate, and both epsilon-TPN and epsilon-TPNH formed binary complexes with the enzyme that were stable to polyacrylamide gel electrophoresis. The fluorescence of epsilon-TPN was enhanced and the emission maximum shifted from 415 to 405 nm when the nucleotide was bound to the enzyme. The ethenoadenine moiety in epsilon-TPNH behaved similarily, but the fluorescence changes were complicated by concurrent effects of binding upon the dihydronicotinamide fluorophore. Fluorescence enhancement titrations yielded values of 1.8 and 0.59 muM, respectively, for the dissociation constants of the enzyme-epsilon-TPN and enzyme-epsilon-TPNH complexes. Titration experiments based upon quenching of enzyme fluorescence gave similar values, viz., 2.1 and 0.53 muM for the dissociation constants of these complexes. Fluorimetric titration of the enzyme-TPNH complex with epsilon-TPN (or of the enzyme-TPN complex with epsilon-TPNH) failed to reveal the presence of a second pyridine nucleotide binding site. The fluorescence enhancement of enzyme-bound epsilon-TPN or dihydrofolate was quenched when amethopterin or epsilon-TPN, respectively, was added to form a ternary complex. These results provide information concerning the nature of the pyridine nucleotide binding site and its spatial relationship to the dihydrofolate/amethopterin binding site.  相似文献   

6.
A protease from Tetrahymena pyriformis inactivated eight of nine commercially available enzymes tested, including lactate deyhdrogenase, isocitrate dehydrogenase (TPN-specific), glucose-6 phosphate dehydrogenase, D-amino acid oxidase, fumarase, pyruvate kinase, hexokinase, and citrate synthase. Urate oxidase was not inactivated. Inactivation occurred at neutral pH, was prevented by inhibitors of the protease, and followed first order kinetics. In those cases tested, inactivation was enhanced by mercaptoethanol. Most of the enzyme-inactivating activity was due to a protease of molecular weight 25,000 that eluted from DEAE-Sephadex at 0.3 M KCl. A second protease of this molecular weight, which was not retained by the gel, inactivated only isocitrate dehydrogenase and D-amino acid oxidase. These two proteases could also be distinguished by temperature and inhibitor sensitivity. Two other protease peaks obtained by DEAE-Sephadex chromatography had little or no no enzyme inactivating activity, while another attacked only D-amino acid oxidase. At least six of the enzymes could be protected from proteolytic inactivation by various ligands. Isocitrates dehydrogenase was protected by isocitrate, TPN, or TPNH, glucose-6-dehydrogenase by glucose-6-P or TPN, pyruvate kinase by phosphoenolypyruvate or ADP, hexokinase by glucose, and fumarase by a mixture of fumarate and malate. Lactate dehdrogenase was not protected by either of its substrates of coenzymes. Citrate synthase was probably protected by oxalacetate. Our data suggest that the protease or proteases discussed here may participate in the inactivation or degradation of a least some enzymes in Tetrahymena. Since the inactivation occurs at neutral pH, this process could be regulated by variations in the cellular levels of substrates, coenzymes, or allosteric regulators resulting form changes in growth conditions or growth state. Such a mechanism would permit the selective retention of enzymes of metabolically active pathways.  相似文献   

7.
Chemical modification of potato apyrase suggests that tryptophan residues are close to the nucleotide binding site. Kd values (+/- Ca2+) for the complexes of apyrase with the non-hydrolysable phosphonate adenine nucleotide analogues, adenosine 5'-(beta,gamma-methylene) triphosphate and adenosine 5'-(alpha,beta-methylene) diphosphate, were obtained from quenching of the intrinsic enzyme fluorescence. Other fluorescent nucleotide analogues (2'(3')-O-(2,4,6-trinitrophenyl) adenosine 5'-triphosphate, 2'(3')-O-(2,4,6-trinitrophenyl) adenosine 5'-diphosphate. 1,N6-ethenoadenosine triphosphate and 1,N6-ethenoadenosine diphosphate) were hydrolysed by apyrase in the presence of Ca2+, indicating binding to the active site. The dissociation constants for the binding of these analogues were calculated from both the decrease of the protein (tryptophan) fluorescence and enhancement of the nucleotide fluorescence. Using the sensitised acceptor (nucleotide analogue) fluorescence method, energy transfer was observed between enzyme tryptophans and ethene-derivatives. These results support the view that tryptophan residues are present in the nucleotide-binding region of the protein, appropriately oriented to allow the energy transfer process to occur.  相似文献   

8.
The binding of glucono-1,5-lactone (gluconolactone) with almond beta-glucosidase was studied at pH 5.0 and 25 degrees C, in the absence and presence of glucose, by monitoring the enzyme fluorescence as a probe. From the results of fluorometric titration, the dissociation constant Kd and the maximum fluorescence intensity increase (percent) of the enzyme-gluconolactone complex relative to the enzyme alone, delta Fmax, were determined to be 12.7 microM and 14.7%, respectively. From the study of the temperature dependence of Kd, delta G degrees, delta H degrees and delta S degrees for the binding were evaluated to be -6.7 kcal mol-1, -3.5 kcal mol-1, and 10.8 e.u. (cal mol-1 deg-1), respectively, at 25 degrees C. The analysis of the fluorometric titration data in the presence of glucose revealed that these ligands bind competitively to the enzyme, probably at the same site. The results of a stopped-flow kinetic study are consistent with the following two-step mechanism: (formula; see text) which indicates that gluconolactone (L) and the enzyme (E) transiently form a loosely bound complex, ELtr (k-1/k+1 = 4.5 mM), in the first rapid bimolecular association step, and ELtr is converted into a more tightly bound complex EL (k+2 = 94 s-1, k-2 = 0.36 s-1) in the subsequent slow unimolecular process. The fluorescence intensity increase occurs solely in the latter step.  相似文献   

9.
Fluorescence correlation spectroscopy (FCS) was used to characterize the interaction of fluorescence labeled single-stranded DNA (ssDNA) with hexameric RepA DNA-helicase (hRepA) encoded by plasmid RSF1010. The apparent dissociation constants, Kd(app), for the equilibrium binding of 12mer, 30mer, and 45mer ssDNA 5'-labeled with BFL to hRepA dimer in the presence of 0.5 mM ATPgammaS at pH 5.8 and 25 degrees C were determined to be 0.58 +/- 0.12, 0.52 +/- 0.07, and 1.66 +/- 0.32 microM, respectively. Binding curves are compatible with one binding site for ssDNA present on hRepA dimer, with no indication of cooperativity. At pH 7.6 in the presence of ATPgammaS and at pH 5.8 in the absence of ATPgammaS, complex formation between ssDNA and hRepA was too weak for measuring complete binding curves by FCS. Under these conditions, the dissociation constant, Kd(app), is in the range between 10 and 250 microM. The kinetics of complex formation at pH 5.8 are faster than the time resolution (approximately 10-20 s) of FCS experiments under pseudo-first-order conditions, with respect to BFL-ssDNA. Photon correlation spectroscopy (PCS) experiments yielded, within the experimental error range, the same values for the apparent hydrodynamic radii, R(h), of hRepA dimer and its complex with ssDNA as determined by FCS (R(h) = 6.6 +/- 1 nm). hRepA starts to aggregate under acidic conditions (相似文献   

10.
Sheep liver cytoplasmic aldehyde dehydrogenase was purified to homogeneity to give a sample with a specific activity of 380 nmol NADH min(-1) mg(-1). An amino acid analysis of the enzyme gave results similar to those reported for aldehyde dehydrogenases from other sources. The isoelectric point was at pH 5.25 and the enzyme contained no significant amounts of metal ions. On the binding of NADH to the enzyme there is a shift in absorption maximum of NADH to 344 nm, and a 5.6-fold enhancement of nucleotide fluorescence. The protein fluorescence (lambdaexcit = 290 nm, lambdaemisson = 340 nm) is quenched on the binding of NAD+ and NADH. The enhancement of nucleotide fluorescence on the binding of NADH has been utilised to determine the dissociation constant for the enzyme . NADH complex (Kd = 1.2 +/- 0.2 muM). A Hill plot of the data gave a straight line with a slope of 1.0 +/- 0.3 indicating the absence of co-operative effects. Ellman's reagent reacted only slowly with the enzyme but in the presence of sodium dodecylsulphate complete reaction occurred within a few minutes to an extent corresponding to 36 thiol groups/enzyme. Molecular weights were determined for both cytoplasmic and mitochondrial aldehyde dehydrogenases and were 212 000 +/- 8 000 and 205 000 respectively. Each enzyme consisted of four subunits with molecular weight of 53 000 +/- 2 000. Properties of the cytoplasmic and mitochondrial aldehyde dehydrogenases from sheep liver were compared with other mammalian liver aldehyde dehydrogenases.  相似文献   

11.
F Widmer  N O Kaplan 《Biochemistry》1976,15(21):4693-4699
Mechanisms involved in the action of the pyridine nucleotide transhydrogenase from Pseudomonas aeruginosa (EC 1.6.1.1) have been investigated by means of kinetic studies and fluorescence titration. Our results, as well as those from previous investigations, suggest that the allosteric MWC model (Monod, J., Wyman, J., and Changeux, J. P. (1965), J. Mol. Biol. 12, 88-118) may be used as a first step for the explanation of the properties of the transhydrogenase. The basic reaction of the enzyme is the oxidation of reduced triphosphopyridine nucleotide (TPNH) by diphosphopyridine nucleotide (DPN+). In terms of the model, the functional R state is favored by TPNH, whereas the product triphosphopyridine nucleotide (TPN+) behaves as an allosteric inhibitor, and is therefore assumed to favor the nonfunctional T state. To a slight extent, the T state is also favored by inorganic phosphate. On the other hand, adenosine 2'-monophosphate and several other 2'-phosphate nucleotides function as activators, and hence are presumed to shift the allosteric equilibrium toward the R state. The studies in this paper suggest a specific regulatory site for the transhydrogenase.  相似文献   

12.
Pyridoxal kinase has been purified 2000-fold from pig brain. The enzyme preparation migrates as a single protein and activity band on analytical gel electrophoresis. The interactions of the substrate pyridoxal and the inhibitor N-dansyl-2-oxopyrrolidine (dansyl = 5-dimethylaminonaphthalene-1-sulfonyl) with the catalytic site were examined by means of fluorescence spectroscopy. The increase in emission anisotropy that follows the binding of pyridoxal to the kinase was used to determine the equilibrium dissociation constant. Pyridoxal kinase binds one molecule of substrate with a Kd = 11 microns at pH 6. The emission anisotropy spectrum of bound pyridoxal reveals that the substrate is not rigidly trapped by the protein matrix. N-Dansyl-2-oxopyrrolidine is a competitive inhibitor with respect to ATP at saturating concentrations of pyridoxal. It binds to the enzyme with a dissociation constant of 6 microns. N-Dansyl-2-oxopyrrolidine is immobilized by strong interactions with the enzyme, but it is displaced from the catalytic site by ATP. The results are consistent with the hypothesis that N-dansyl-2-oxopyrrolidine binds at the nucleotide binding site of pyridoxal kinase.  相似文献   

13.
R S Levy  J J Villafranca 《Biochemistry》1977,16(15):3293-3231
Electron paramagnetic resonance (EPR) spectra were obtained for various isocitrate dehydrogenase-Mn(II) complexes. The qualitative effects of the binding of substrates, nucleotides, and substrate analogues on the isotropic character of the electronic environment of enzyme-bound Mn(II) were subsequently investigated. The addition of isocitrate produces a markedly anisotropic spectrum whereas alpha-ketoglutarate does not alter the spectrum of enzyme-Mn(II) substantially. This suggests direct coordination of isocitrate to the Mn(II) but perphaps a different mode of binding for alpha-ketoglutarate. Other studies demonstrated mutually exclusive binding relationships between TPN and TPNH, between Mn-isocitrate and TPNH, and between HCO3-(CO2) and formate or thiocyanate. Indirect evidence supporting CO2 rather than HCO3-as the actual reactive species which binds to the enzyme in the reductive carboxylation reaction is presented on the basis of the results of the formate and thiocyanate studies. From the EPR results recorded for ternary, quaternary, and quinary enzyme-substrate complexes, correlations between the appearance of fine structure signals and the binding of individual substrates and/or nucleotides are found, and tentative assignments of such signals are made on this basis. Additional studies were conducted to determine binding constants for Mg(II) Co(II), and Co-isocitrate, and a comparison was made with kinetically determined binding constants.  相似文献   

14.
It was previously reported that 2',3'-O-(2,4,6-trinitrocyclohexadienylidene) (TNP)-nucleotides bind with high affinity to the sarcoplasmic reticulum Ca-ATPase (Dupont, Y., Chapron, Y., and Pougeois, R. (1982) Biochem. Biophys. Res. Commun. 106, 1272-1279 and Watanabe, T., and Inesi, G. (1982) J. Biol. Chem. 257, 11510-11516). Here we report a study of the Ca-ATPase nucleotide binding sites using TNP-nucleotides. Competition at equilibrium between TNP-nucleotides and ATP was measured in the absence of calcium; it was found that TNP-nucleotides and ATP competitively bind to two classes of sites of equal concentration (3.5 nmol/mg). The ATP dissociation constants for the two classes of sites were found to be sensitive to H+ and Mg2+ concentrations. In the absence of Mg2+ (independently of pH) or at acid pH (independently of Mg2+ concentration), the nucleotide sites behave like one single family of sites of intermediate affinity (Kd = 20 microM). They split into two classes of sites of high (Kd = 2-4 microM) and low (Kd greater than 1 mM) affinity at pH values higher than neutral and in the presence of Mg2+. The calcium-activated ATP hydrolysis is accelerated by TNP-ATP (or TNP-AMP-PNP) binding on the phosphorylated enzyme. It is concluded 1) that the Ca-ATPase enzyme possesses two classes of ATP binding sites, 2) that the affinity of these two sites and the nature of their interaction is modulated by the H+ and Mg2+ concentrations, and 3) that the hydrolytic activity of the high affinity ATP binding site is activated by ATP or TNP-AMP-PNP (or TNP-ATP) binding in a low affinity ATP binding site.  相似文献   

15.
When α-ketoglutarate is the substrate, malate is a considerably more effective inhibitor of glutamate dehydrogenase than glutamate, oxalacetate, aspartate, or glutarate. Malate is a considerably poorer inhibitor when glutamate is the substrate. Malate is competitive with α-ketoglutarate, uncompetitive with TPNH, and noncompetitive with glutamate. The above, plus the fact that malate is a considerably more potent inhibitor when TPNH rather than TPN is the coenzyme, indicates that malate is predominantly bound to the α-ketoglutarate site of the enzyme-TPNH complex and has a considerably lower affinity for the enzyme-TPN complex. Ligands which decrease binding of TPNH to the enzyme such as ADP and leucine markedly decrease inhibition by malate. Conversely, GTP, which increases binding of TPNH to the enzyme also enhances inhibition by malate. Malate also decreases interaction between mitochondrial aspartate aminotransferase and glutamate dehydrogenase. This effect of malate on enzyme-enzyme interaction is enhanced by DPNH and GTP which also increase inhibition of glutamate dehydrogenase by malate and is decreased by TPN, ADP, ATP, α-ketoglutarate, and leucine which decrease inhibition of glutamate dehydrogenase by malate. These results indicate that malate could decrease α-ketoglutarate utilization by inhibiting glutamate dehydrogenase and retarding transfer of α-ketoglutarate from the aminotransferase to glutamate dehydrogenase. These effects of malate would be most pronounced when the mitochondrial level of α-ketoglutarate is low and the level of malate and reduced pyridine nucleotide is high.  相似文献   

16.
Pig heart NADP-dependent isocitrate dehydrogenase is 65% inactivated by 3-bromo-2-ketoglutarate (Ehrlich, R.S., and Colman, R.F., 1987, J. Biol. Chem. 262, 12,614-12,619) and 90% inactivated by 2-(4-bromo-2,3-dioxobutylthio)-1,N6- ethenoadenosine 2',5'-bisphosphate (2-BDB-T epsilon A-2',5'-DP) (Bailey, J.M., and Colman, R.F., 1987, J. Biol. Chem. 262, 12,620-12,626). Both inactivation reactions result in enzyme with an incorporation of 1.0 mol reagent/mol enzyme dimer and both modified enzymes bind only 1.0 mol manganous isocitrate or NADPH/mol enzyme dimer as compared to 2.0 mol manganous isocitrate or NADPH/mol enzyme dimer for unmodified enzyme. The inactivation reactions, which occur at or near the nucleotide binding site, are mutually exclusive. Reaction with either affinity reagent led to the isolation of the same modified triskaidekapeptide, DLAGXIHGLSNVK. We have isolated from isocitrate dehydrogenase a peptide, DLAGCIHGLSNVK, that had been modified by N-ethylmaleimide (NEM) with no loss of enzymatic activity. We now show that enzyme modified by NEM in the presence of isocitrate plus Mn2+ retains full catalytic activity but is not inactivated by either of the affinity reagents; thus, all three reagents appear to react at the same site. The analysis of HPLC tryptic maps of isocitrate dehydrogenase treated under denaturing conditions with iodo[3H]acetic acid or [3H]NEM demonstrates that both bromoketoglutarate and 2-BDB-T epsilon A-2',5'-DP react with the cysteine residue of DLAGCIHGLSNVK. We conclude that the cysteine of this triskaidekapeptide is close to the coenzyme binding site but is not essential for catalytic function.  相似文献   

17.
Nucleotide binding to sarcoplasmic reticulum vesicles was investigated in the absence of calcium using both filtration and fluorescence measurements. Filtration assays of binding of radioactive nucleotides at concentrations up to 0.1 mM gave a stoichiometry of one ATP-binding site/sarcoplasmic reticulum ATPase molecule. When measured in the presence of calcium under otherwise similar conditions, ATPase velocity rose 4-8-fold (depending on pH and magnesium concentration) when the ATP concentration was increased from 1 microM to 0.1 mM. Binding of ATP and ADP enhanced the intrinsic fluorescence of sarcoplasmic reticulum ATPase, but AMP and adenosine did not affect it. Both filtration and fluorescence measurements showed that binding of metal-free ATP is independent of pH (Kd = 20-25 microM) but that the presence of magnesium induces pH dependence of the binding of the Mg.ATP complex (Kd = 10 microM at pH 6.0 and 1.5 microM at pH 8.0). Binding of metal-free ADP was pH-dependent but was not affected by magnesium. High magnesium concentrations inhibited nucleotide binding. These results suggest that ATP interacts with two different domains of Ca-ATPase that form the catalytic site. The first domain may bind the adenine moiety of the substrate, and the pH dependence of ADP binding suggests the participation of His683 in this region. The second domain of the catalytic site may bind the gamma-phosphate and the magnesium ion of the Mg.ATP complex and constitute the locus of the electrostatic interactions between the substrate and the enzyme.  相似文献   

18.
J M Bailey  R F Colman 《Biochemistry》1987,26(15):4893-4900
When the substrate isocitrate-Mn2+ is present, the fluorescent nucleotide analogue 2-[(4-bromo-2,3-dioxobutyl)thio]-1,N6-ethenoadenosine 2',5'-bisphosphate (2-BDB-T epsilon A-2',5'-DP) reacts irreversibly with pig heart NADP+-specific isocitrate dehydrogenase at the coenzyme binding site on one subunit of the dimeric enzyme [Bailey, J. M., & Colman, R. F. (1985) Biochemistry 24, 5367-5377]. The modified enzyme, which retains partial activity, binds 1 mol of NADPH or 1 mol of the coenzyme analogue, reduced thionicotinamide adenine dinucleotide phosphate (TNADPH), per dimer. TNADPH quenches the fluorescence of enzyme-bound 2-BDB-T epsilon A-2',5'-DP with an efficiency of energy transfer of 9.8%. From this value and the spectral properties of the donor and acceptor chromophores, a distance of 32 A was calculated as the average distance between coenzyme sites on the two subunits. Isocitrate dehydrogenase activity requires a divalent metal ion, such as Mn2+, Co2+, or Ni2+. Co2+ and Ni2+ have absorption spectra that overlap the emission spectra of enzyme-bound 2-BDB-T epsilon A-2',5'-DP. In the presence of isocitrate, each of these two metal ions quenches the fluorescence of the enzyme-bound reagent with an efficiency of energy transfer of 28-29%. From this value and the spectral characteristics of the energy donor and acceptors, an average distance of 8.0 A was estimated between the metal-isocitrate site and the labeled coenzyme site. These distances have provided constraints in formulating a model of the spatial arrangement of active-site ligands on isocitrate dehydrogenase.  相似文献   

19.
Terbium ions and terbium formycin triphosphate have been used to investigate the interactions between the cation and nucleotide binding sites of the sarcoplasmic reticulum Ca2+-ATPase. Three classes of Tb3+-binding sites have been found: a first class of low-affinity (Kd = 10 microM) corresponds to magnesium binding sites, located near a tryptophan residue of the protein; a second class of much higher affinity (less than 0.1 microM) corresponds to the calcium transport sites, their occupancy by terbium induces the E1 to E2 conformational change of the Ca2+-ATPase; a third class of sites is revealed by following the fluorescence transfer from formycin triphosphate (FTP) to terbium, evidencing that terbium ions can also bind into the nucleotide binding site at the same time as FTP. Substitution of H2O by D2O shows that Tb-FTP binding to the enzyme nucleotide site is associated with an important dehydration of the terbium ions associated with FTP. Two terbium ions, at least, bind to the Ca2+-ATPase in the close vicinity of FTP when this nucleotide is bound to the ATPase nucleotide site. Addition of calcium quenches the fluorescence signal of the terbium-FTP complex bound to the enzyme. Calcium concentration dependence shows that this effect is associated with the replacement of terbium by calcium in the transport sites, inducing the E2----E1 transconformation when calcium is bound. One interpretation of this fluorescence quenching is that the E1----E2 transition induces an important structural change in the nucleotide site. Another interpretation is that the high-affinity calcium sites are located very close to the Tb-FTP complex bound to the nucleotide site.  相似文献   

20.
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   

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