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 46 毫秒
1.
C T Grubmeyer  K W Chu  S Insinga 《Biochemistry》1987,26(12):3369-3373
Salmonella typhimurium histidinol dehydrogenase produces histidine from the amino alcohol histidinol by two sequential NAD-linked oxidations which form and oxidize a stable enzyme-bound histidinaldehyde intermediate. The enzyme was found to catalyze the exchange of 3H between histidinol and [4(R)-3H]NADH and between NAD and [4(S)-3H]NADH. The latter reaction proceeded at rates greater than kcat for the net reaction and was about 3-fold faster than the former. Histidine did not support an NAD/NADH exchange, demonstrating kinetic irreversibility in the second half-reaction. Specific activity measurements on [3H]histidinol produced during the histidinol/NADH exchange reaction showed that only a single hydrogen was exchanged between the two reactants, demonstrating that under the conditions employed this exchange reaction arises only from the reversal of the alcohol dehydrogenase step and not the aldehyde dehydrogenase reaction. The kinetics of the NAD/NADH exchange reaction demonstrated a hyperbolic dependence on the concentration of NAD and NADH when the two were present in a 1:2 molar ratio. The histidinol/NADH exchange showed severe inhibition by high NAD and NADH under the same conditions, indicating that histidinol cannot dissociate directly from the ternary enzyme-NAD-histidinol complex; in other words, the binding of substrate is ordered with histidinol leading. Binding studies indicated that [3H]histidinol bound to 1.7 sites on the dimeric enzyme (0.85 site/monomer) with a KD of 10 microM. No binding of [3H]NAD or [3H]NADH was detected. The nucleotides could, however, displace histidinol dehydrogenase from Cibacron Blue-agarose.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The kinetic mechanism and the substrate specificity of liver alcohol dehydrogenase are changed when 3-benzoylpyridine-adenine dinucleotide is used as coenzyme. Only primary alcohols are substrates of the enzyme and with ethanol the mechanism becomes rapid-equilibrium random bi-bi. According to model building experiments on a graphic display, the benzoyl group partially enters the substrate binding site, whereas the essential interactions between coenzyme and enzyme are preserved. This restraint on the substrate binding site provides a molecular explanation for the observed dependence between coenzyme and substrate chemical structures.  相似文献   

3.
The biochemistry of interallelic complementation within the Salmonella typhimurium hisD gene was investigated by in vitro protein complementation of mutant histidinol dehydrogenases (EC 1.1.1.23). Double-mutant strains were constructed containing the hisO1242 (constitutive overproducer) attenuator mutation and selected hisDa or hisDb mutations. Extracts from such hisDa986 and hisDb1799 mutant cells failed to show histidinol dehydrogenase activity but complemented to produce active enzyme. Inactive mutant histidinol dehydrogenases were purified from each of the two mutants by ion-exchange chromatography. Complementation by the purified mutant proteins required the presence of 2-mercaptoethanol and MnCl2, and protein-protein titrations indicated that heterodimers were strongly preferred in mixtures of the complementary mutant enzymes. Neither mutant protein showed negative complementation with wild-type enzyme. The Vmax for hybrid histidinol dehydrogenase was 11% of that for native enzyme, with only minor changes in Km values for substrate or coenzyme. Both purified mutant proteins failed to catalyze NAD-NADH exchange reactions reflective of the first catalytic step of the two-step reaction. The inactive enzymes bound 54Mn2+ weakly or not at all in the presence of 2-mercaptoethanol, in contrast to wild-type enzyme which bound 54Mn2+ to 0.6 sites per monomer under the same conditions. The mutant proteins, like wild-type histidinol dehydrogenase, behaved as dimers on analytical gel filtration chromatography, but dissociated to form monomers in the presence of 2-mercaptoethanol. This effect of 2-mercaptoethanol was prevented by low levels of MnCl2. It thus appears that mutant histidinol dehydrogenase molecules bind metal ion poorly. The complementation procedure may allow for formation of a functional Mn2+-binding site, perhaps at the subunit interface.  相似文献   

4.
The enzyme 2,3-dihydro-2,3-dihydroxybenzoate dehydrogenase (2,3-diDHB dehydrogenase, hereafter Ent A), the product of the enterobactin biosynthetic gene entA, catalyzes the NAD(+)-dependent oxidation of the dihydroaromatic substrate 2,3-dihydro-2,3-dihydroxybenzoate (2,3-diDHB) to the aromatic catecholic product 2,3-dihydroxybenzoate (2,3-DHB). The catechol 2,3-DHB is one of the key siderophore units of enterobactin, a potent iron chelator secreted by Escherichia coli. To probe the reaction mechanism of this oxidation, a variety of 2,3-diDHB analogues were synthesized and tested as substrates. Specifically, we set out to elucidate both the regio- and stereospecificity of alcohol oxidation as well as the stereochemistry of NAD+ reduction. Of those analogues tested, only those with a C3-hydroxyl group (but not a C2-hydroxyl group) were oxidized to the corresponding ketone products. Reversibility of the Ent A catalyzed reaction was demonstrated with the corresponding NADH-dependent reduction of 3-ketocyclohexane- and cyclohexene-1-carboxylates but not the 2-keto compounds. These results establish that Ent A functions as an alcohol dehydrogenase to specifically oxidize the C3-hydroxyl group of 2,3-diDHB to produce the corresponding 2-hydroxy-3-oxo-4,6-cyclohexadiene-1-carboxylate (Scheme II) as a transient species that undergoes rapid aromatization to give 2,3-DHB. Stereospecificity of the C3 allylic alcohol group oxidation was confirmed to be 3R in a 1R,3R dihydro substrate, 3, and hydride transfer occurs to the si face of enzyme-bound NAD+.  相似文献   

5.
Drosophila alcohol dehydrogenase (DADH) is an NAD+-dependent enzyme that catalyzes the oxidation of alcohols to aldehydes/ketones and that is also able to further oxidize aldehydes to their corresponding carboxylic acids. The structure of the ternary enzyme-NADH-acetate complex of the slow alleloform of Drosophila melanogaster ADH (DmADH-S) was solved at 1.6 A resolution by X-ray crystallography. The coenzyme stereochemistry of the aldehyde dismutation reaction showed that the obtained enzyme-NADH-acetate complex reflects a productive ternary complex although no enzymatic reaction occurs. The stereochemistry of the acetate binding in the bifurcated substrate-binding site, along with previous stereochemical studies of aldehyde reduction and alcohol oxidation shows that the methyl group of the aldehyde in the reduction reaction binds to the R1 and in the oxidation reaction to the R2 sub-site. NMR studies along with previous kinetic studies show that the formed acetaldehyde intermediate in the oxidation of ethanol to acetate leaves the substrate site prior to the reduced coenzyme, and then binds to the newly formed enzyme-NAD+ complex. Here, we compare the three-dimensional structure of D.melanogaster ADH-S and a previous theoretically built model, evaluate the differences with the crystal structures of five Drosophila lebanonensis ADHs in numerous complexed forms that explain the substrate specificity as well as subtle kinetic differences between these two enzymes based on their crystal structures. We also re-examine the electrostatic influence of charged residues on the surface of the protein on the catalytic efficiency of the enzyme.  相似文献   

6.
L C Hsu  M Okamoto  E E Snell 《Biochimie》1989,71(4):477-489
A coupled assay with alpha-hydroxyglutarate dehydrogenase was used to analyze the kinetic behavior of histidinol phosphate aminotransferase from Salmonella typhymurium. Data obtained from studies of initial velocity, inhibition by products or substrate analogues, isotope exchange rates, and the determination of the equilibrium constant were consistent only with a Ping-Pong Bi Bi mechanism. Variations in inhibition patterns by different substrate analogues indicate that the microenvironment about the pyridoxal phosphate and the pyridoxamine phosphate forms of histidinol phosphate amino-transferase are different, and favor the presence of one active site with partially overlapping substrate-binding subsites for these 2 forms of the enzyme. Histidinol phosphate aminotransferase also catalyzes decomposition of beta-chloro-L-alanine to pyruvate, NH3 and Cl-; no transamination of this substrate occurs and inactivation of the enzyme accompanies this reaction. After reduction of histidinol-P aminotransferase with [3H]NaBH4, carboxymethylation, and tryptic digestion, one major radioactive peptide absorbing at 325 nm was isolated. Its primary structure was determined to be TLSK*AFALAGLR, where K* is the P-pyridoxyllysine residue. Although this peptide is only 30-40% homologous with the corresponding segment reported for other transaminases, all of these peptides are similar in placement of an hydroxyamino acid residue three residues upstream from the lysine residue, and in the cluster of hydrophobic amino acid residues immediately following the lysine residue.  相似文献   

7.
Cytoplasmic malate dehydrogenase (cMDH) is a key enzyme in several metabolic pathways. Though its activity has been examined extensively, there are lingering mechanistic uncertainties involving substrate and cofactor binding. To more completely understand this enzyme's interactions with cofactor and substrate ligands, a fluorescent reporter group was introduced into the enzyme's structure. This was accomplished by selective modification of Cys 110. The reaction placed an aminonaphthaline sulfonic acid group near the enzyme's active site. Substrate, inhibitor, and NAD binding activities were characterized using changes in this label's fluorescence. Results demonstrated that both substrate and cofactor molecules bound to the enzyme in the absence of their companion ligands. This is in contrast to strictly ordered cofactor then substrate binding as has been suggested by kinetic analyses of closely related enzymes. Binding results also indicated that the cofactor, NAD, bound to cMDH in a negatively cooperative manner, but substrates and the inhibitor, hydroxymalonate, bound non-cooperatively. Multiple substrate binding modes were identified and interactions between substrate and cofactor binding were found.  相似文献   

8.
The transient kinetics of aldehyde reduction by NADH catalyzed by liver alcohol dehydrogenase consist of two kinetic processes. This biphasic rate behavior is consistent with a model in which one of the two identical subunits in the enzyme is inactive during the reaction at the adjacent protomer. Alternatively, enzyme heterogeneity could result in such biphasic behavior. We have prepared liver alcohol dehydrogenase containing a single major isozyme; and the transient kinetics of this purified enzyme are biphasic.Addition of two [14C]carboxymethyl groups per dimer to the two “reactive” sulfhydryl groups (Cys46) yields enzyme which is catalytically inactive toward alcohol oxidation. Alkylated enzyme, as initially isolated by gel filtration chromatography at pH 7·5, forms an NAD+-pyrazole complex. However, the ability to bind NAD+-pyrazole is rapidly lost in pH 8·75 buffer; therefore, our alkylated preparations, as isolated by chromatography at pH 8·75, are inactive toward NAD+-pyrazole complex formation. We have prepared partially inactivated enzyme by allowing iodoacetic acid to react with liver alcohol dehydrogenase until 50% of the NAD+-pyrazole binding capacity remains; under these reaction conditions one [14C]carboxymethyl group is added per dimer. This partially alkylated enzyme preparation is isolated by gel filtration and has been aged sufficiently to lose NAD+-pyrazole binding ability at alkylated subunits. When solutions of native liver alcohol dehydrogenase and partially alkylated liver alcohol dehydrogenase containing the same number of unmodified active sites are allowed to react with substrate under single turnover conditions, partially alkylated enzyme is only half as reactive as native enzyme. This indicates that some molecular species in partially alkylated liver alcohol dehydrogenase that react with pyrazole and NAD+ during the active site titration do not react with substrate. These data are consistent with a model in which a subunit adjacent to an alkylated protomer in the dimeric enzyme is inactive toward substrate. In addition, NAD+-pyrazole binding at the protomers adjacent to alkylated subunits is slowly lost so that 75% of the enzyme-NAD+-pyrazole binding capacity is lost in 50% alkylated enzyme. These data supply strong evidence for subunit interactions in liver alcohol dehydrogenase.Binding experiments performed on partially alkylated liver alcohol dehydrogenase indicate that coenzyme binding is normal at a subunit adjacent to an alkylated protomer even though active ternary complexes cannot be formed. One hypothesis consistent with these results is the unavailability of zinc for substrate binding at the active site in subunits adjacent to alkylated protomers in monoalkylated dimer.  相似文献   

9.
The molecular features of 19 synthetic substrates and ground-state analogues of cycloeucalenol, the natural substrate of cycloeucalenol - obtusifoliol isomerase, a membrane-bound enzyme specific to higher plants, and of 9 synthetic carbocationic analogues of the high-energy intermediate occurring during the reaction catalyzed by the isomerase, were related to their ability to be transformed by this enzyme (catalytical competence) and their potency as inhibitors of this enzyme. With substrates and ground-state analogues it has been possible to determine at least two critical domains: significant binding requires the presence of the 3 beta-hydroxyl group on the ring A with the correct stereochemistry together with absence of a 4 beta-methyl group. Moreover initial enzyme-substrate interaction appears to be dependent upon the accessibility of the 3 beta-oxygen. Substitutions on the ring B do not preclude binding whereas they are of great influence on substrate transformation. Modifications of the ring A and other modifications suggest that ground-state and high-energy intermediate analogues bind two different conformations of the isomerase active site.  相似文献   

10.
Sanghani PC  Robinson H  Bosron WF  Hurley TD 《Biochemistry》2002,41(35):10778-10786
The human glutathione-dependent formaldehyde dehydrogenase is unique among the structurally studied members of the alcohol dehydrogenase family in that it follows a random bi bi kinetic mechanism. The structures of an apo form of the enzyme, a binary complex with substrate 12-hydroxydodecanoic acid, and a ternary complex with NAD+ and the inhibitor dodecanoic acid were determined at 2.0, 2.3, and 2.3 A resolution by X-ray crystallography using the anomalous diffraction signal of zinc. The structures of the enzyme and its binary complex with the primary alcohol substrate, 12-hydroxydodecanoic acid, and the previously reported binary complex with the coenzyme show that the binding of the first substrate (alcohol or coenzyme) causes only minor changes to the overall structure of the enzyme. This is consistent with the random mechanism of the enzyme where either of the substrates binds to the free enzyme. The catalytic-domain position in these structures is intermediate to the "closed" and "open" conformations observed in class I alcohol dehydrogenases. More importantly, two different tetrahedral coordination environments of the active site zinc are observed in these structures. In the apoenzyme, the active site zinc is coordinated to Cys44, His66 and Cys173, and a water molecule. In the inhibitor complex, the coordination environment involves Glu67 instead of the solvent water molecule. The coordination environment involving Glu67 as the fourth ligand likely represents an intermediate step during ligand exchange at the active site zinc. These observations provide new insight into metal-assisted catalysis and substrate binding in glutathione-dependent formaldehyde dehydrogenase.  相似文献   

11.
Tsang WY  Amyes TL  Richard JP 《Biochemistry》2008,47(16):4575-4582
The ratio of the second-order rate constants for reduction of dihydroxyacetone phosphate (DHAP) and of the neutral truncated substrate glycolaldehyde (GLY) by glycerol 3-phosphate dehydrogenase (NAD (+), GPDH) saturated with NADH is (1.0 x 10 (6) M (-1) s (-1))/(8.7 x 10 (-3) M (-1) s (-1)) = 1.1 x 10 (8), which was used to calculate an intrinsic phosphate binding energy of at least 11.0 kcal/mol. Phosphite dianion binds very weakly to GPDH ( K d > 0.1 M), but the bound dianion strongly activates GLY toward enzyme-catalyzed reduction by NADH. Thus, the large intrinsic phosphite binding energy is expressed only at the transition state for the GPDH-catalyzed reaction. The ratio of rate constants for the phosphite-activated and the unactivated GPDH-catalyzed reduction of GLY by NADH is (4300 M (-2) s (-1))/(8.7 x 10 (-3) M (-1) s (-1)) = 5 x 10 (5) M (-1), which was used to calculate an intrinsic phosphite binding energy of -7.7 kcal/mol for the association of phosphite dianion with the transition state complex for the GPDH-catalyzed reduction of GLY. Phosphite dianion has now been shown to activate bound substrates for enzyme-catalyzed proton transfer, decarboxylation, hydride transfer, and phosphoryl transfer reactions. Structural data provide strong evidence that enzymic activation by the binding of phosphite dianion occurs at a modular active site featuring (1) a binding pocket complementary to the reactive substrate fragment which contains all the active site residues needed to catalyze the reaction of the substrate piece or of the whole substrate and (2) a phosphate/phosphite dianion binding pocket that is completed by the movement of flexible protein loop(s) to surround the nonreacting oxydianion. We propose that loop motion and associated protein conformational changes that accompany the binding of phosphite dianion and/or phosphodianion substrates lead to encapsulation of the substrate and/or its pieces in the protein interior, and to placement of the active site residues in positions where they provide optimal stabilization of the transition state for the catalyzed reaction.  相似文献   

12.
NAD(+)-dependent (R)-2-hydroxyglutarate dehydrogenase (HGDH) catalyses the reduction of 2-oxoglutarate to (R)-2-hydroxyglutarate and belongs to the d-2-hydroxyacid NAD(+)-dependent dehydrogenase (d-2-hydroxyacid dehydrogenase) protein family. Its crystal structure was determined by phase combination to 1.98 A resolution. Structure-function relationships obtained by the comparison of HGDH with other members of the d-2-hydroxyacid dehydrogenase family give a chemically satisfying view of the substrate stereoselectivity and catalytic requirements for the hydride transfer reaction. A model for substrate recognition and turnover is discussed. The HGDH active site architecture is structurally optimized to recognize and bind the negatively charged substrate 2-oxoglutarate. The structural position of the side chain of Arg52, and its counterparts in other family members, strongly correlates with substrate specificity towards substitutions at the C3 atom (linear or branched substrates). Arg235 interacts with the substrate's alpha-carboxylate and carbonyl groups, having a dual role in both substrate binding and activation, and the gamma-carboxylate group can dock at an arginine cluster. The proton-relay system built up by Glu264 and His297 permits His297 to act as acid-base catalyst and the 4Re-hydrogen from NADH is transferred as hydride to the carbonyl group Si-face leading to the formation of the correct enantiomer (R)-2-hydroxyglutarate.  相似文献   

13.
A three-dimensional model of yeast alcohol dehydrogenase, based on the homologous horse liver enzyme, was used to compare the substrate binding pockets of the three isozymes (I, II, and III) from Saccharomyces cerevisiae and the enzyme from Schizosaccharomyces pombe. Isozyme I and the S. pombe enzyme have methionine at position 294 (numbered as in the liver enzyme, corresponding to 270 in yeast), whereas isozymes II and III have leucine. Otherwise the active sites of the S. cerevisiae enzymes are the same. All four wild-type enzymes were produced from the cloned genes. In addition, oligonucleotide-directed mutagenesis was used to change Met-294 in alcohol dehydrogenase I to leucine. The mechanisms for all five enzymes were predominantly ordered with ethanol (but partially random with butanol) at pH 7.3 and 30 degrees C. The wild-type alcohol dehydrogenases and the leucine mutant had similar kinetic constants, except that isozyme II had 10-20-fold smaller Michaelis and inhibition constants for ethanol. Thus, residue 294 is not responsible for this difference. Apparently, substitutions outside of the substrate binding pocket indirectly affect the interactions of the alcohol dehydrogenases with ethanol. Nevertheless, the substitution of methionine with leucine in the substrate binding site of alcohol dehydrogenase I produced a 7-10-fold increase in reactivity (V/Km) with butanol, pentanol, and hexanol. The higher activity is due to tighter binding of the longer chain alcohols and to more rapid hydrogen transfer.  相似文献   

14.
S Ghisla  C Thorpe  V Massey 《Biochemistry》1984,23(14):3154-3161
Butyryl-CoA dehydrogenase from Megasphera elsdenii catalyzes the exchange of the alpha- and beta-hydrogens of substrate with solvent [Gomes, B., Fendrich, G., & Abeles, R. H. (1981) Biochemistry 20, 1481-1490]. The stoichiometry of this exchange was determined by using 3H2O label as 1.94 +/- 0.1 per substrate molecule. The rate of 3H label incorporation into substrate under anaerobic conditions is monophasic, indicating that both the alpha- and beta-hydrogens exchange at the same rate. The exchange in 2H2O leads to incorporation of one 2H each into the alpha- and the beta-positions of butyryl-CoA, as determined by companion 1H NMR experiments and confirmed by mass spectroscopic analysis. In contrast, with general acyl-CoA dehydrogenase from pig kidney, only exchange of the alpha-hydrogen was found. The beta-hydrogen is the one that is transferred (reversibly) to the flavin 5-position during substrate dehydrogenation. This was demonstrated by reacting 5-3H- and 5-2H-reduced 5-deaza-FAD-general acyl-CoA dehydrogenase with crotonyl-CoA. Only one face of the reduced flavin analogue is capable of transferring hydrogen to substrate. The rate of this reaction is 11.1 s-1 for 5-deaza-FAD-enzyme and 2.2 s-1 for [5-2H]deaza-FAD-enzyme, yielding an isotope effect of 5. These values compare with a rate of 2.6 s-1 for the reaction of native reduced enzyme with crotonyl-CoA. The two reduced enzymes (normal vs. 5-deaza-FAD-enzyme) thus react at similar rates, indicating a similar mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
The 2-oxoglutarate dehydrogenase complex was purified from Azotobacter vinelandii. The complex consists of three components, 2-oxoglutarate dehydrogenase/decarboxylase (E1o), lipoate succinyltransferase (E2o) and lipoamide dehydrogenase (E3). Upon purification, the E3 component dissociates partially from the complex. From reconstitution experiments, the Kd for E3 was found to be 26 nM, about 30 times higher than that for the pyruvate dehydrogenase complex. The Km values for the substrates 2-oxoglutarate, CoA and NAD+ were found to be 0.15, 0.014 and 0.17 mM, respectively. The system has a high specificity for 2-oxoglutarate, which is determined by the action of both E1o and E2o. Above 4 mM substrate inhibition is observed. From steady-state inhibition experiments with substrate analogs, two substrate-binding modes are revealed at different degrees of saturation of the enzyme with 2-oxoglutarate. At low substrate concentrations (10(-6) to 10(-5) M), the binding mainly depends on the interaction of the enzyme with the substrate carboxyl groups. At a higher degree of substrate saturation (10(-4) to 10(-3) M) the relative contribution of the 2-oxo group in the binding increases. A kinetic analysis points to a single binding site for a substrate analog under steady state conditions. Saturation of this site with an analog indicates that two kinetically different complexes are formed with 2-oxoglutarate in the course of catalysis. From competition studies with analogs it is concluded that one of these complexes is formed at the site that is sterically identical to the substrate inhibition site. The data obtained are represented by a minimal scheme that considers formation of a precatalytic complex SE between the substrate and E1o before the catalytic complex ES, in which the substrate is added to the thiamin diphosphate cofactor, is formed. The incorrect orientation of the substrate molecule in SE or the occupation of this site by analogs is supposed to cause substrate or analog inhibition, respectively.  相似文献   

16.
The structure of a phenylalanine aminomutase (TcPAM) from Taxus canadensis has been determined at 2.4 ? resolution. The active site of the TcPAM contains the signature 4-methylidene-1H-imidazol-5(4H)-one prosthesis, observed in all catalysts of the class I lyase-like family. This catalyst isomerizes (S)-α-phenylalanine to the (R)-β-isomer by exchange of the NH2/H pair. The stereochemistry of the TcPAM reaction product is opposite of the (S)-β-tyrosine made by the mechanistically related tyrosine aminomutase (SgTAM) from Streptomyces globisporus. Since TcPAM and SgTAM share similar tertiary- and quaternary-structures and have several highly conserved aliphatic residues positioned analogously in their active sites for substrate recognition, the divergent product stereochemistries of these catalysts likely cannot be explained by differences in active site architecture. The active site of the TcPAM structure also is in complex with (E)-cinnamate; the latter functions as both a substrate and an intermediate. To account for the distinct (3R)-β-amino acid stereochemistry catalyzed by TcPAM, the cinnamate skeleton must rotate the C1-Cα and Cipso-Cβ bonds 180° in the active site prior to exchange and rebinding of the NH2/H pair to the cinnamate, an event that is not required for the corresponding acrylate intermediate in the SgTAM reaction. Moreover, the aromatic ring of the intermediate makes only one direct hydrophobic interaction with Leu-104. A L104A mutant of TcPAM demonstrated an ~1.5-fold increase in kcat and a decrease in KM values for sterically demanding 3'-methyl-α-phenylalanine and styryl-α-alanine substrates, compared to the kinetic parameters for TcPAM. These parameters did not change significantly for the mutant with 4'-methyl-α-phenylalanine compared to those for TcPAM.  相似文献   

17.
J K Wright  M Takahashi 《Biochemistry》1977,16(8):1541-1548
The aspartokinase activity of the aspartokinase-homoserine dehydrogenase complex of Escherichia coli was affinity labeled with substrates ATP, aspartate, and feedback inhibitor threonine. Exchange-inert ternary adducts of Co(III)-aspartokinase and either ATP, aspartate or threonine were formed by oxidation of corresponding Co(II) ternary complexes with H2O2. The ternary enzyme-Co(III)-threonine adduct (I) had 3.8 threonine binding sites per tetramer, one-half that of the native enzyme. The binding of threonine to I was still cooperative as determined by equilibrium dialysis (nH = 2.2) or by studying inhibition of residual dehydrogenase activity (nH = 2.7). Threonine still protected the SH groups of I against 5,5'-dithiobis(2-nitrobenzoate) (DTNB) reaction but the number of SH groups reacting with thiol reagents (DTNB) was reduced by 1-2 per subunit in the absence of threonine. This suggests either that Co(III) is bound to the enzyme via sulfhydryl groups or that 1-2SH groups are buried or rendered inaccessible in I. The binding of threonine to sites not blocked by the affinity labeling produced changes in the circular dichroism of the complex comparable to changes produced by threonine binding to native enzyme and also protected against proteolytic digestion. The major conformational changes produced by threonine are thus ascribable to binding at this one class of regulatory sites. The interactions of kinase substrates with various aspartokinase-Co(III) complexes containing ATP, aspartate, or threonine and a threonine-insensitive homoserine dehydrogenase produced by mild proteolysis were studied. The inhibition of homoserine dehydrogenase by kinase substrates is not due to binding of these inhibitors at the kinase active site but was shown to be due to binding to sites within the dehydrogenase domain of the enzyme. L-alpha-Aminobutyrate, a presumed threonine analogue, also inhibits the dehydrogenase by binding at the same or similar sites in the dehydrogenase domain and not at threonine regulatory site.  相似文献   

18.
The synthesis of 5-(2-oxalylethyl)-NADH, a reduced nicotinamide adenine dinucleotide (NADH) derivate with pyruvate covalently attached to the 5 position of the dihydronicotinamide ring over an additional methylene group has been described previously (Trommer, W.E., Blume, H., and Kapmeyer, H. (1976) Justus Liebigs Ann. Chem., 848). In the presence of lactate dehydrogenase, the dihydropyridine ring of this coenzyme-substrate analogue is oxidized and the carbonyl function of the side chain is reduced to the corresponding L-hydroxy derivative with a maximum velocity of 1/3000 of the natural reaction. This reaction is intramolecular as shown by competition experiments with pyruvate. 5-(2-oxalylethyl)-NADH (pyr-NADH) appears to be a true transition state analogue, proving its postulated structure. Pyr-NADH is high specific for this enzyme as demonstrated by the facts that (1) D-lactate dehydrogenase does not catalyze the intramolecular redox reaction, although the substrate moiety of pyr-NADH is reduced in the presence of NADH; (2) when tested with malate dehydrogenase, alcohol dehydrogenase, glyceraldehyde phosphate dehydrogenase,glycerate dehydrogenase, and glycerol dehydrogenase pyr-NADH is not even oxidized in the presence of the corresponding substrates. However, a great similarity between the transition states of the reduction of pyruvate catalyzed by lactate dehydrogenase and alanine dehydrogenase could be shown. Alanine dehydrogenase catalyzes the intramolecular redox reaction as well. In the presence of ammonium ions, pyr-NADH is transformed to 5-(3-carboxyl-3-aminopropyl)-NAD+.  相似文献   

19.
The N-methylacridinium cation is bound to hydrophobic sites of horse liver alcohol dehydrogenase and human serum albumin with an observed stoichiometry of one molecule N-methyl-acridinium chloride per subunit of alcohol dehydrogenase and 2.5 molecules of the dye per molecule human serum albumin; the dissociation constants are 3.6 X 10(-5) M and 1.7 X 10(-5) M, respectively. In light, the proteins catalyze the dismutation of N-methylacridinium chloride to N-methylacridone and N-methyl-9,10-dihydroacridine. The presence or absence of oxygen has no effect upon the observed reaction rate. If horse liver alcohol dehydrogenase is used as catalyst, the reaction is inhibited by adenosine diphosphoribose and by 1,1'-dimethyl-4,4'-bipyridylium dichloride. It is concluded that the N-methylacridinium cation is bound within the catalytic site of the enzyme interacting with the binding sites of the nicotinium ring and/or the binding site of the lipophilic part of the substrate. The anaerobic photodismutation of N-methylacridinium chloride to N-methyl-9,10-dihydroacridine and N-methylacridone can be explained by several alternative patways (see Appendix by S. Hünig), the overall reaction being 2[N-Methylacridinium]+ + H2Ohw leads to N-Methyl-9,10-dihydroacridine + N-methylacridone + 2H+. The prerequisite, a high rate of proton transfer from the reaction site, seems to be common property of the hydrophobic binding regions for the N-methylacridinium cation in both horse liver alcohol dehydrogenase and human serum albumin.  相似文献   

20.
Initial velocities of the histidinol dehydrogenase reaction (EC 1.1.1.23) were measured as a function of the concentrations of the substrates histidinol and NAD+ and in the presence and absence of the product NADH. The data are consistent with a Bi Uni Uni Bi Ping Pong mechanism. The kinetic constants of this mechanism were determined; Km for histidinol was found to be 14 microM and for NAD+ 0.7 mV; Ki for NAD+ was 0.4 mM.  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