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 31 毫秒
1.
Several denitrifying Pseudomonas strains contained an NADP+-specific 2-oxoglutarate dehydrogenase, in contrast to an NAD+-specific pyruvate dehydrogenase, if the cells were grown anaerobically with aromatic compounds. With non-aromatic substrates or after aerobic growth the coenzyme specificity of 2-oxoglutarate dehydrogenase changed to NAD+-specificity. The reaction stoichiometry and the apparent K m-values of the enriched enzymes were determined: pyruvate 0.5 mM, coenzyme A 0.05 mM, NAD+ 0.25 mM; 2-oxoglutarate 0.6 mM, coenzyme A 0.05 mM, NADP+ 0.03 mM. Isocitrate dehydrogenase was NADP+-specific. The findings suggest that these strains contained at least two lipoamide dehydrogenases, one NAD+-specific, the other NADP+-specific.  相似文献   

2.
In this work, we have postulated a comprehensive and unified chemical mechanism of action for yeast alcohol dehydrogenase (EC 1.1.1.1, constitutive, cytoplasmic), isolated from Saccharomyces cerevisiae. The chemical mechanism of yeast enzyme is based on the integrity of the proton relay system: His-51....NAD+....Thr-48....R.CH2OH(H2>O)....Zn++, stretching from His-51 on the surface of enzyme to the active site zinc atom in the substrate-binding site of enzyme. Further, it is based on extensive studies of steady-state kinetic properties of enzyme which were published recently. In this study, we have reported the pH-dependence of dissociation constants for several competitive dead-end inhibitors of yeast enzyme from their binary complexes with enzyme, or their ternary complexes with enzyme and NAD+ or NADH; inhibitors include: pyrazole, acetamide, sodium azide, 2-fluoroethanol, and 2,2,2-trifluorethanol. The unified mechanism describes the structures of four dissociation forms of apoenzyme, two forms of the binary complex E.NAD+, three forms of the ternary complex E.NAD+.alcohol, two forms of the ternary complex E.NADH.aldehyde and three binary complexes E.NADH. Appropriate pKa values have been ascribed to protonation forms of most of the above mentioned complexes of yeast enzyme with coenzymes and substrates.  相似文献   

3.
《Bioorganic chemistry》1987,15(1):31-42
The use of NAD+ analogs lacking a carbonyl function at position C-3 of the pyridinium moiety allowed the manipulation of the kinetic mechanism of calf spleen NAD+ glycohydrolase so as to render the cleavage of the pyridinium-ribose bond rate limiting. The analogs used in this study are relatively poor substrates of the enzyme. They present an affinity for the active site which is independent of the nature of their substituent (Ki = 10 ± 2 μm), suggesting that the specificity of the NAD+ glycohydrolase reflects the dynamic steps occurring after the formation of the Michaelis complex. The maximal rates of hydrolysis of the NAD+ analogs are very sensitive to the pKa of the departing pyridine; a Brönsted plot (r = 0.99) gave a βtg = −0.90 (at 37°C). From this plot we could estimate that for NAD+, the specific interaction of the 3-carboxamide group with the active site contributed to the catalysis by decreasing the energy barrier by about 2 kcal mol−1. We have also studied the nonenzymatic hydrolysis of NAD+ and its analogs under conditions (pH-independent hydrolysis) which favor a unimolecular mechanism. In this case a linear Brönsted plot was also found (r = 0.99) with βtg = −1.11 (at 37°C). Our data indicate that NAD+ glycohydrolase catalyzes the chemical cleavage of the pyridinium-ribose bond, over a 103 rate difference, according to a single mechanism involving a late transition state in which the scissile bond is broken. The present study strongly supports our previous hypothesis (F. Schuber, P. Travo, and M. Pascal (1979) Bioorg. Chem. 8, 83) according to which NAD+ glycohydrolase catalyses unimolecular decomposition of its substrates with generation of an ADP-ribosyl oxocarbonium ion intermediate which must be stabilized by the active site of the enzyme.  相似文献   

4.
The hydrolysis of NAD+ analogs bearing different substituents in the pyridinium moiety, catalyzed by solubilized calf spleen NAD+ glycohydrolase, was studied. The enzyme was specific for analogs possessing a carbonyl function at position C-3. The observed maximal rates showed no simple dependence either on the leaving ability of the parent pyridines or on the observed binding energies. The analogs without carbonyl substituents were not found to be hydrolyzed under our experimental conditions; and, compared to the substrates, they all presented much higher affinities for the active site, which could not be related to specific interactions. These results indicate that the rate-limiting step of the NAD+ hydrolysis, which is the formation of an enzyme-ADP-ribosyl intermediate, is probably more complex than a simple chemical step, i.e., pyridinium-ribose bond breakage. At a molecular level we favor a catalytic mechanism which, through nonbonded interactions between the substrate and the active site of the enzyme, results in the destabilization of the pyridinium-ribose bond, i.e., unimolecular decomposition involving an oxocarbonium ion intermediate.  相似文献   

5.
4,4-bis Dimethylaminodiphenylcarbinol (BDC-OH) has recently been reported to be a highly sensitive reagent for the quantitative determination of sulfhydryl residues in biological materials (1). In this communication the effectiveness of BDC-OH as a reagent for selective chemical modification of “active center” cysteine residues was investigated. The supernatant and mitochondrial forms of malate dehydrogenase were chosen for investigation by this reagent. Supernatant malate dehydrogenase which has never been found to contain an “active center” cysteine is unaffected by this reagent. Mitochondrial malate dehydrogenase (L malate: NAD+ oxidoreductase, EC 1.1.1.37) from porcine heart can be irreversibly inactivated by a 20 fold M excess of the reagent. Chemical modification of two essential sulfhydryl residues is prevented by the presence of the coenzyme, NAD+, suggesting that the site of interaction is located at or near the coenzyme binding site and hence at or near the enzymatic center of this enzyme.  相似文献   

6.
Mitochondrial malate dehydrogenase (L-malate: NAD+ oxidoreductase, EC 1.1.1.37) from porcine heart exhibits a time dependent loss in enzymatic activity in the presence of the reagent butanedione. The inhibition occurs concomitant with the modification of 2.4 residues of arginine per molecular weight of 70,000. The presence of the reduced coenzyme, NADH, protects the enzyme from inhibition by butanedione and from modification of arginine residues, suggesting that the residues modified are located near the coenzyme binding site and hence at or near the enzymatic active center of this enzyme.  相似文献   

7.
NAD+ has been covalently attached to dextrans having different molecular weights to give various NAD+ densities (mol NAD+ per mol d-glucosyl residue). The effects of molecular weight of dextran and of NAD+ density on the coenzyme activity of the dextran-bound NAD+ derivatives were examined for the reactions catalysed by alcohol dehydrogenase (alcohol: NAD+ oxidoreductase, EC 1.1.1.1) and lactate dehydrogenase (l-lactate:NAD+ oxidoreductase, EC 1.1.1.27). The molecular weight of dextran had little effect on coenzyme activity in the range 10 000 to 500 000. At low NAD+ density (<0.05 mol NAD+/mol d-glucosyl residue), the coenzyme activities of the derivatives were relatively low, but higher densities had little effect on the activity. Dextran-bound NAD+ derivatives were twice as stable as free NAD+.  相似文献   

8.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

9.
Liver alcohol dehydrogenase (LADH; E.C. 1.1.1.1) provides an excellent system for probing the role of binding interactions with NAD+ and alcohols as well as with NADH and the corresponding aldehydes. The enzyme catalyzes the transfer of hydride ion from an alcohol substrate to the NAD+ cofactor, yielding the corresponding aldehyde and the reduced cofactor, NADH. The enzyme is also an excellent catalyst for the reverse reaction. X-ray crystallography has shown that the NAD+ binds in an extended conformation with a distance of 15 Å between the buried reacting carbon of the nicotinamide ring and the adenine ring near the surface of the horse liver enzyme. A major criticism of X-ray crystallographic studies of enzymes is that they do not provide dynamic information. Such data provide time-averaged and space-averaged models. Significantly, entries in the protein data bank contain both coordinates as well as temperature factors. However, enzyme function involves both dynamics and motion. The motions can be as large as a domain closure such as observed with liver alcohol dehydrogenase or as small as the vibrations of certain atoms in the active site where reactions take place. Ternary complexes produced during the reaction of the enzyme binary entity, E-NAD+, with retinol (vitamin A alcohol) lead to retinal (vitamin A aldehyde) release and the enzyme binary entity E-NADH. Retinal is further metabolized via the E-NAD+-retinal ternary complex to retinoic acid (vitamin A acid). To unravel the mechanistic aspects of these transformations, the kinetics and energetics of interconversion between various ternary complexes are characterized. Proton transfers along hydrogen bond bridges and NADH hydride transfers along hydrophobic entities are considered in some detail. Secondary kinetic isotope effects with retinol are not particularly large with the wild-type form of alcohol dehydrogenase from horse liver. We analyze alcohol dehydrogenase catalysis through a re-examination of the reaction coordinates. The ground states of the binary and ternary complexes are shown to be related to the corresponding transition states through topology and free energy acting along the reaction path.  相似文献   

10.
NAD+ and NADP+, chemically similar and with almost identical standard oxidation–reduction potentials, nevertheless have distinct roles, NAD+ serving catabolism and ATP generation whereas NADPH is the biosynthetic reductant. Separating these roles requires strict specificity for one or the other coenzyme for most dehydrogenases. In many organisms this holds also for glutamate dehydrogenases (GDH), NAD+-dependent for glutamate oxidation, NADP+-dependent for fixing ammonia. In higher animals, however, GDH has dual specificity. It has been suggested that GDH in mitochondria reacts only with NADP(H), the NAD+ reaction being an in vitro artefact. However, contrary evidence suggests mitochondrial GDH not only reacts with NAD+ but maintains equilibrium using the same pool as accessed by β-hydroxybutyrate dehydrogenase. Another complication is the presence of an energy-linked dehydrogenase driving NADP+ reduction by NADH, maintaining the coenzyme pools at different oxidation–reduction potentials. Its coexistence with GDH makes possible a futile cycle, control of which is not yet properly explained. Structural studies show NAD+-dependent, NADP+-dependent and dual-specificity GDHs are closely related and a few site-directed mutations can reverse specificity. Specificity for NAD+ or for NADP+ has probably emerged repeatedly during evolution, using different structural solutions on different occasions. In various GDHs the P7 position in the coenzyme-binding domain plays a key role. However, whereas in other dehydrogenases an acidic P7 residue usually hydrogen bonds to the 2′- and 3′-hydroxyls, dictating NAD+ specificity, among GDHs, depending on detailed conformation of surrounding residues, an acidic P7 may permit binding of NAD+ only, NADP+ only, or in higher animals both.  相似文献   

11.
The fluorescence decay mechanism of 1, N6-ethenoadenosine diphosphoribose bound to rabbit muscle glyceraldehyde 3-phosphate dehydrogenase markedly differs from that of the intact coenzyme analog (εNAD+) bound to the same enzyme. In the latter case the fluorescence is partially quenched by interactions between the ethenoadenine ring and amino acid residues in its binding site. Binding of the nicotinamide moiety of the coenzyme thus affects the relative orientation of the adenine ring within its binding site leading to the quenching interactions. The interactions of the adenine group with its binding site induce conformational changes in the enzyme which affect the binding of additional coenzyme molecules. The nicotinamide base thus determines, indirectly, the negative cooperativity found in NAD+ binding.  相似文献   

12.
Structural analysis of glucose dehydrogenase from Haloferax mediterranei revealed that the adenosine 2′-phosphate of NADP+ was stabilized by the side chains of Arg207 and Arg208. To investigate the structural determinants for coenzyme specificity, several mutants involving residues Gly206, Arg207 and Arg208 were engineered and kinetically characterized. The single mutants G206D and R207I were less efficient with NADP+ than the wild type, and the double and triple mutants G206D/R207I and G206D/R207I/R208N showed no activity with NADP+.In the single mutant G206D, the relation kcat/KNAD+ was 1.6 times higher than in the wild type, resulting in an enzyme that preferred NAD+ over NADP+. The single mutation was sufficient to modify coenzyme specificity, whereas other dehydrogenases usually required more than one or two mutations to change coenzyme specificity. However, the highest reaction rates were reached with the double mutant G206D/R207I and with coenzyme NAD+, where the kcat was 1.6 times higher than the kcat of the wild-type enzyme with NADP+. However, catalytic efficiency with NAD+ was lower, as the Km value for coenzyme was 77 times higher than the wild type with NADP+.  相似文献   

13.
A malic enzyme from a cell-free extract of Pseudomonas diminuta IFO-13182 was purified to electrophoretic homogeneity by DEAE-Sepharose, Sephacryl, and Blue-Sepharose chromatographies. The purified enzyme required either NAD+ or NADP+ as a coenzyme. From the results of coenzyme specificity, the enzyme should be classified as l-malate: NAD+ oxidoreductase (decarboxylating) [EC 1.1.1.39]. The purified enzyme was most active at pH 7.5 and 50°C and was stable in the pH range from 7.0 to 9.0. The isoelectric point was pH 4.3. Its molecular weight was 680,000 by COSMOSIL 5-Diol high performance liquid gel filtration on chromatography and 65,000 by SDS polyacrylamide gel electrophoresis. This indicates that the enzyme consisted of 10 subunits. The malic enzyme activity with NADP+ was about twice that measured with NAD+.  相似文献   

14.
  • 1.1. A simple, facile one-step method has been devised to measure the stereospecificity of NADP+-linked oxidoreductases. The procedure involves coupling the test enzymes to enzymes of known stereospecificity in the presence of deuterated substrates. The regenerated NADP+ in the coupled reactions is analyzed by PMR for its deuterium content at the carbon-4 position of the nicotinamide ring.
  • 2.2. It is found that malate dehydrogenase (EC 1.1.1.37). lactate dehydrogenase (EC 1.1.1.27) and glycerate dehydrogenase (EC 1.1.1.29) are A-side stereospecific whereas glutamate dehydrogenase (EC 1.4.1.3) and glycerol-3-phosphate dehydrogenase (EC 1.1.1.8) are B-side stereospecific.
  • 3.3. Enzymes which can utilize both NAD+ and NADP+ have the same stereospecificity with respect to the coenzyme.
  相似文献   

15.
The ethanologenic bacterium Zymomonas mobilis ZM4 is of special interest because it has a high ethanol yield. This is made possible by the two alcohol dehydrogenases (ADHs) present in Z. mobilis ZM4 (zmADHs), which shift the equilibrium of the reaction toward the synthesis of ethanol. They are metal-dependent enzymes: zinc for zmADH1 and iron for zmADH2. However, zmADH2 is inactivated by oxygen, thus implicating zmADH2 as the component of the cytosolic respiratory system in Z. mobilis. Here, we show crystal structures of zmADH2 in the form of an apo-enzyme and an NAD+-cofactor complex. The overall folding of the monomeric structure is very similar to those of other functionally related ADHs with structural variations around the probable substrate and NAD+ cofactor binding region. A dimeric structure is formed by the limited interactions between the two subunits with the bound NAD+ at the cleft formed along the domain interface. The catalytic iron ion binds near to the nicotinamide ring of NAD+, which is likely to restrict and locate the ethanol to the active site together with the oxidized Cys residue and several nonpolar bulky residues. The structures of the zmADH2 from the proficient ethanologenic bacterium Z. mobilis, with and without NAD+ cofactor, and modeling ethanol in the active site imply that there is a typical metal-dependent catalytic mechanism.  相似文献   

16.
The interactions of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase with NAD+ and with its fluorescent derivative 1, N6-etheno-adenine dinucleotide were investigated using a variety of spectroscopic methods. These techniques included: difference spectroscopy, circular dichroism, fluorescence and circular polarized luminescence. It was found that the greatest structural change in the protein tetramer occurs upon binding of the first mole of coenzyme. We have also demonstrated that progressive structural changes occur at the adenine subsite in the NAD+ binding site as a function of coenzyme saturation. These conformational changes are probably responsible for the progressive decrease in the affinity towards the coenzyme. It was also found that every NAD+ molecule induces the same conformational change of the nicotinamide subsite. These results offer a molecular explanation for the negative co-operativity in the binding of the coenzyme, without a change in the catalytic power of the NAD+ site as a function of coenzyme saturation. These results also offer a new explanation for the fact that enzyme exhibits half-of-the-sites reactivity towards certain ligands and full-site reactivity towards others. It is suggested that those ligands interacting at the adenine subsite of the NAD+ binding site induce the half-of-the-sites reactivity.Our results support the view that both the negative co-operativity in coenzyme binding and half-of-the-sites reactivity are due to ligand-induced conformational changes on an a priori symmetric glyceraldehyde-3-phosphate dehydrogenase molecule.  相似文献   

17.
Betaine aldehyde dehydrogenase (BADH EC 1.2.1.8) catalyzes the irreversible oxidation of betaine aldehyde to glycine betaine using NAD+ as a coenzyme. Porcine kidney BADH (pkBADH) follows a bi‐bi ordered mechanism in which NAD+ binds to the enzyme before the aldehyde. Previous studies showed that NAD+ induces complex and unusual conformational changes on pkBADH and that potassium is required to maintain its quaternary structure. The aim of this work was to analyze the structural changes in pkBADH caused by NAD+ binding and the role played by potassium in those changes. The pkBADH cDNA was cloned and overexpressed in Escherichia coli, and the protein was purified by affinity chromatography using a chitin matrix. The pkBADH/NAD+ interaction was analyzed by circular dichroism (CD) and by isothermal titration calorimetry (ITC) by titrating the enzyme with NAD+. The cDNA has an open reading frame of 1485 bp and encodes a protein of 494 amino acids, with a predicted molecular mass of 53.9 kDa. CD data showed that the binding of NAD+ to the enzyme caused changes in its secondary structure, whereas the presence of K+ helps maintain its α‐helix content. K+ increased the thermal stability of the pkBADH‐NAD+ complex by 5.3°C. ITC data showed that NAD+ binding occurs with different association constants for each active site between 37.5 and 8.6 μM. All the results support previous data in which the enzyme incubation with NAD+ provoked changes in reactivity, which is an indication of slow conformational rearrangements of the active site.  相似文献   

18.
With “reactive” arginine as a kinetic reporter, 2-hydroxyacid dehydrogenases are assessed in basis of their specialization as NAD+-reducing or NADH-oxidizing enzymes. Specifically, M4 and H4 lactate dehydrogenases (LDHs) and cytoplasmic and mitochondrial malate dehydrogenases (MDHs) are compared to assess if their coenzyme specificity may involve electrostatics of cationic or neutral nicotinamide structure as the basis. The enzymes from diverse eukaryote and prokaryote sources thus are assessed in “reactivity” of functionally-critical arginine as a function of salt concentration and pH. Electrostatic calculations were performed on “reactive” arginines and found good correspondence with experiment. The reductive and oxidative LDHs and MDHs are assessed in their count over ionizable residues and in placement details of the residues in their structures as proteins. The variants found to be high or low in ΔpKa of “reactive” arginine are found to be also strong or weak cations that preferentially oxidize NADH (neutral nicotinamide structure) or reduce NAD+ (cationic nicotinamide structure). The ionized groups of protein structure may thus be important to redox specificity of the enzyme on basis of electrostatic preference for the oxidized (cationic nicotinamide) or reduced (neutral nicotinamide) coenzyme. Detailed comparisons of isozymes establish that the residues contributing in their redox specificity are scrambled in structure of the reductive enzyme.  相似文献   

19.
A range of biomimetic analogues of the nicotinamide nucleotide coenzymes NAD(P)(H) have been developed based on the structure of a triazine dye template. These biomimetic redox coenzymes are relatively straightforward and inexpensive to synthesise and display NAD+-like activity with different dehydrogenases, despite their apparently minimal structural similarity to the native coenzyme NAD+. Horse liver alcohol dehydrogenase oxidises butan-1-ol, using the most active biomimetic coenzyme (Nap 1), with a k cat value an order of magnitude lower and a K m for the coenzyme two orders of magnitude higher than those using native NAD+. The enzymatically reduced biomimetic coenzymes may be reoxidised by phenazine methosulfate. We believe that these coenzymes may find applications in biotransformations and biosensors, and in the development of biomimetic catalysts where the redox enzyme itself is replaced by a synthetic binding site. Received: 26 October 1998 / Received revision: 25 January 1999 / Accepted: 31 January 1999  相似文献   

20.
It has been hypothesized that the bifunctional enzyme DmpFG channels its intermediate, acetaldehyde, from one active site to the next using a buried intermolecular channel identified in the crystal structure. This channel appears to switch between an open and a closed conformation depending on whether the coenzyme NAD+ is present or absent. Here, we applied molecular dynamics and metadynamics to investigate channeling within DmpFG in both the presence and absence of NAD+. We found that substrate channeling within this enzyme is energetically feasible in the presence of NAD+ but was less likely in its absence. Tyr-291, a proposed control point at the channel's entry, does not appear to function as a molecular gate. Instead, it is thought to orientate the substrate 4-hydroxy-2-ketovalerate in DmpG before reaction occurs, and may function as a proton shuttle for the DmpG reaction. Three hydrophobic residues at the channel's exit appear to have an important role in controlling the entry of acetaldehyde into the DmpF active 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