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 468 毫秒
1.
Pichia stipitis NAD(+)-dependent xylitol dehydrogenase (XDH), a medium-chain dehydrogenase/reductase, is one of the key enzymes in ethanol fermentation from xylose. For the construction of an efficient biomass-ethanol conversion system, we focused on the two areas of XDH, 1) change of coenzyme specificity from NAD(+) to NADP(+) and 2) thermostabilization by introducing an additional zinc atom. Site-directed mutagenesis was used to examine the roles of Asp(207), Ile(208), Phe(209), and Asn(211) in the discrimination between NAD(+) and NADP(+). Single mutants (D207A, I208R, F209S, and N211R) improved 5 approximately 48-fold in catalytic efficiency (k(cat)/K(m)) with NADP(+) compared with the wild type but retained substantial activity with NAD(+). The double mutants (D207A/I208R and D207A/F209S) improved by 3 orders of magnitude in k(cat)/K(m) with NADP(+), but they still preferred NAD(+) to NADP(+). The triple mutant (D207A/I208R/F209S) and quadruple mutant (D207A/I208R/F209S/N211R) showed more than 4500-fold higher values in k(cat)/K(m) with NADP(+) than the wild-type enzyme, reaching values comparable with k(cat)/K(m) with NAD(+) of the wild-type enzyme. Because most NADP(+)-dependent XDH mutants constructed in this study decreased the thermostability compared with the wild-type enzyme, we attempted to improve the thermostability of XDH mutants by the introduction of an additional zinc atom. The introduction of three cysteine residues in wild-type XDH gave an additional zinc-binding site and improved the thermostability. The introduction of this mutation in D207A/I208R/F209S and D207A/I208R/F209S/N211R mutants increased the thermostability and further increased the catalytic activity with NADP(+).  相似文献   

2.
Hydroxysteroid dehydrogenases (HSDs) are essential for the biosynthesis and mechanism of action of all steroid hormones. We report the complete kinetic mechanism of a mammalian HSD using rat 3alpha-HSD of the aldo-keto reductase superfamily (AKR1C9) with the substrate pairs androstane-3,17-dione and NADPH (reduction) and androsterone and NADP(+) (oxidation). Steady-state, transient state kinetics, and kinetic isotope effects reconciled the ordered bi-bi mechanism, which contained 9 enzyme forms and permitted the estimation of 16 kinetic constants. In both reactions, loose association of the NADP(H) was followed by two conformational changes, which increased cofactor affinity by >86-fold. For androstane-3,17-dione reduction, the release of NADP(+) controlled k(cat), whereas the chemical event also contributed to this term. k(cat) was insensitive to [(2)H]NADPH, whereas (D)k(cat)/K(m) and the (D)k(lim) (ratio of the maximum rates of single turnover) were 1.06 and 2.06, respectively. Under multiple turnover conditions partial burst kinetics were observed. For androsterone oxidation, the rate of NADPH release dominated k(cat), whereas the rates of the chemical event and the release of androstane-3,17-dione were 50-fold greater. Under multiple turnover conditions full burst kinetics were observed. Although the internal equilibrium constant favored oxidation, the overall K(eq) favored reduction. The kinetic Haldane and free energy diagram confirmed that K(eq) was governed by ligand binding terms that favored the reduction reactants. Thus, HSDs in the aldo-keto reductase superfamily thermodynamically favor ketosteroid reduction.  相似文献   

3.
Heredia VV  Penning TM 《Biochemistry》2004,43(38):12028-12037
3Alpha-hydroxysteroid dehydrogenases (3alpha-HSDs) catalyze the interconversion between 5alpha-dihydrotestosterone (5alpha-DHT), the most potent androgen, and 3alpha-androstanediol (3alpha-diol), a weak androgen metabolite. To identify the rate-determining step in this physiologically important reaction, rat liver 3alpha-HSD (AKR1C9) was used as the protein model for the human homologues in fluorescence stopped-flow transient kinetic and kinetic isotope effect studies. Using single and multiple turnover experiments to monitor the NADPH-dependent reduction of 5alpha-DHT, it was found that k(lim) and k(max) values were identical to k(cat), indicating that chemistry is rate-limiting overall. Kinetic isotope effect measurements, which gave (D)k(cat) = 2.4 and (D)2(O)k(cat) = 3.0 at pL 6.0, suggest that the slow chemical transformation is significantly rate-limiting. When the NADP(+)-dependent oxidation of 3alpha-diol was monitored, single and multiple turnover experiments showed a k(lim) and burst kinetics consistent with product release as being rate-limiting overall. When NAD(+) was substituted for NADP(+), burst phase kinetics was eliminated, and k(max) was identical to k(cat). Thus with the physiologically relevant substrates 5alpha-DHT plus NADPH and 3alpha-diol plus NAD(+), the slowest event is chemistry. R276 forms a salt-linkage with the phosphate of 2'-AMP, and when it is mutated, tight binding of NAD(P)H is no longer observed [Ratnam, K., et al. (1999) Biochemistry 38, 7856-7864]. The R276M mutant also eliminated the burst phase kinetics observed for the NADP(+)-dependent oxidation of 3alpha-diol. The data with the R276M mutant confirms that the release of the NADPH product is the slow event; and in its absence, chemistry becomes rate-limiting. W227 is a critical hydrophobic residue at the steroid binding site, and when it is mutated to alanine, k(cat)/K(m) for oxidation is significantly depressed. Burst phase kinetics for the NADP(+)-dependent turnover of 3alpha-diol by W227A was also abolished. In the W227A mutant, the slow release of NADPH is no longer observed since the chemical transformation is now even slower. Thus, residues in the cofactor and steroid-binding site can alter the rate-determining step in the NADP(+)-dependent oxidation of 3alpha-diol to make chemistry rate-limiting overall.  相似文献   

4.
The crystal structure of human type III 3alpha-hydroxysteroid dehydrogenase (HSD)/bile acid binding protein (AKR1C2) complexed with NADP(+) and 3alpha,7beta-dihydroxy-5beta-cholanic acid (ursodeoxycholate) at 3.0 A resolution is presented. Thus, the three-dimensional structure has now been solved for a human HSD member of the aldo-keto reductase superfamily. AKR1C2 is implicated in the prostatic production of the potent androgen 5alpha-dihydrotestosterone and the hepatic transport of bile acids. It also catalyzes the formation of the neurosteroid 3alpha-hydroxy-5alpha-pregnan-20-one in the central nervous system, and its allosteric modulation by fluoxetine has been linked to the use of this drug for premenstrual dsyphoria. Like other members of the superfamily, AKR1C2 folds into an alpha/beta-barrel and binds NADP(+) in an extended conformation. The carboxylate of ursodeoxycholate binds to AKR1C2 in the oxyanion hole at the active site. More interestingly, the orientation of ursodeoxycholate is essentially "backwards" and "upside-down" from that observed for testosterone in the related rat 3alpha-HSD.NADP(+).testosterone ternary complex, where testosterone assumes the position of a 3-ketosteroid substrate. The orientation of ursodeoxycholate is thus similar to that expected of a 17beta-HSD substrate. The ternary structure explains the ability of AKR1C2 to catalyze 3alpha-, 17beta-, and 20alpha-HSD reactions. Comparison of the steroid binding pocket of AKR1C2 with that of rat 3alpha-HSD reveals significant differences in the positions of conserved and nonconserved loop residues, providing insights into the structural basis for the functional flexibility that is observed in all the human 3alpha-HSD isoforms but not in the rat isoform.  相似文献   

5.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

6.
The mouse 17alpha-hydroxysteroid dehydrogenase (m17alpha-HSD) is the unique known member of the aldo-keto reductase (AKR) superfamily able to catalyze efficiently and in a stereospecific manner the conversion of androstenedione (Delta4) into epi-testosterone (epi-T), the 17alpha-epimer of testosterone. Structural and mutagenic studies had already identified one of the residues delineating the steroid-binding cavity, A24, as the major molecular determinant for the stereospecificity of m17alpha-HSD. We report here a ternary complex crystal structure (m17alpha-HSD:NADP(+):epi-T) determined at 1.85 A resolution that confirms this and reveals a unique steroid-binding mode for an AKR enzyme. Indeed, in addition to the interactions found in all other AKRs (van der Waals contacts stabilizing the core of the steroid and the hydrogen bonds established at the catalytic site by the Y55 and H117 residues with the oxygen atom of the ketone group to be reduced), m17alpha-HSD establishes with the other extremity of the steroid nucleus an additional interaction involving K31. By combining direct mutagenesis and kinetic studies, we found that the elimination of this hydrogen bond did not affect the affinity of the enzyme for its steroid substrate but led to a slight but significant increase of its catalytic efficiency (k(cat)/K(m)), suggesting a role for K31 in the release of the steroidal product at the end of the reaction. This previously unobserved steroid-binding mode for an AKR is similar to that adopted by other steroid-binding proteins, the hydroxysteroid dehydrogenases of the short-chain dehydrogenases/reductases (SDR) family and the steroid hormone nuclear receptors. Mutagenesis and structural studies made on the human type 3 3alpha-HSD, a closely related enzyme that shares 73% amino acids identity with the m17alpha-HSD, also revealed that the residue at position 24 of these two enzymes directly affects the binding and/or the release of NADPH, in addition to its role in their 17alpha/17beta stereospecificity.  相似文献   

7.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

8.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

9.
Ma H  Ratnam K  Penning TM 《Biochemistry》2000,39(1):102-109
Rat liver 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD), an aldo-keto reductase, binds NADP(+) in an extended anti-conformation across an (alpha/beta)(8)-barrel. The orientation of the nicotinamide ring, which permits stereospecific transfer of the 4-pro-R hydride from NAD(P)H to substrate, is achieved by hydrogen bonds formed between the C3-carboxamide of the nicotinamide ring and Ser 166, Asn 167, and Gln 190 and by pi-stacking between this ring and Tyr 216. These residues were mutated to yield S166A, N167A, Q190A, and Y216S. In these mutants, K(d)(NADP(H)) increased by 2-11-fold but without a significant change in K(d)(NAD(H)). Steady-state kinetic parameters showed that K(m)(NADP)()+ increased 13-151-fold, and this was accompanied by comparable decreases in k(cat)/K(m)(NADP)()+. By contrast, K(m)(NAD)()+ increased 4-8-fold, but changes in k(cat)/K(m)(NAD)()+ were more dramatic and ranged from 23- to 930-fold. Corresponding changes in binding energies indicated that each residue contributed equally to the binding of NADP(H) in the ground and transition states. However, the same residues stabilized the binding of NAD(H) only in the transition state. These observations suggest that different modes of binding exist for NADP(H) and NAD(H). Importantly, these modes were revealed by mutating residues in the nicotinamide pocket indicating that direct interactions with the 2'-phosphate in the adenine mononucleotide is not the sole determinant of cofactor preference. The single mutations were unable to invert or racemize the stereochemistry of hydride transfer even though the nicotinamide pocket can accommodate both anti- and syn-conformers once the necessary hydrogen bonds are eliminated. When 4-pro-R-[(3)H]NADH was used to monitor incorporation into [(14)C]-5alpha-dihydrotestosterone, a decrease in the (3)H:(14)C ratio was observed in the mutants relative to wild-type enzyme reflecting a pronounced primary kinetic isotope effect. This observation coupled with the change in the binding energy for NAD(P)(H) in the transition state suggests that these mutants have altered the reaction trajectory for hydride transfer.  相似文献   

10.
Chloroplast ferredoxin-NADP(+) reductase has a 32,000-fold preference for NADPH over NADH, consistent with its main physiological role of NADP(+) photoreduction for de novo carbohydrate biosynthesis. Although it is distant from the 2'-phosphoryl group of NADP(+), replacement of the C-terminal tyrosine (Tyr(308) in the pea enzyme) by Trp, Phe, Gly, and Ser produced enzyme forms in which the preference for NADPH over NADH was decreased about 2-, 10-, 300-, and 400-fold, respectively. Remarkably, in the case of the Y308S mutant, the k(cat) value for the NADH-dependent activity approached that of the NADPH-dependent activity of the wild-type enzyme. Furthermore, difference spectra of the NAD(+) complexes revealed that the nicotinamide ring of NAD(+) binds at nearly full occupancy in the active site of both the Y308G and Y308S mutants. These results correlate well with the k(cat) values obtained with these mutants in the NADH-ferricyanide reaction. The data presented support the hypothesis that specific recognition of the 2'-phosphate group of NADP(H) is required but not sufficient to ensure a high degree of discrimination against NAD(H) in ferredoxin-NADP(+) reductase. Thus, the C-terminal tyrosine enhances the specificity of the reductase for NADP(H) by destabilizing the interaction of a moiety common to both coenzymes, i.e. the nicotinamide.  相似文献   

11.
L Zhang  B Ahvazi  R Szittner  A Vrielink  E Meighen 《Biochemistry》1999,38(35):11440-11447
The fatty aldehyde dehydrogenase from the luminescent bacterium, Vibrio harveyi (Vh-ALDH), is unique with respect to its high specificity for NADP(+) over NAD(+). By mutation of a single threonine residue (Thr175) immediately downstream of the beta(B) strand in the Rossmann fold, the nucleotide specificity of Vh-ALDH has been changed from NADP(+) to NAD(+). Replacement of Thr175 by a negatively charged residue (Asp or Glu) resulted in an increase in k(cat)/K(m) for NAD(+) relative to that for NADP(+) of up to 5000-fold due to a decrease for NAD(+) and an increase for NADP(+) in their respective Michaelis constants (K(a)). Differential protection by NAD(+) and NADP(+) against thermal inactivation and comparison of the dissociation constants of NMN, 2'-AMP, 2'5'-ADP, and 5'-AMP for these mutants and the wild-type enzyme clearly support the change in nucleotide specificity. Moreover, replacement of Thr175 with polar residues (N, S, or Q) demonstrated that a more efficient NAD(+)-dependent enzyme T175Q could be created without loss of NADP(+)-dependent activity. Analysis of the three-dimensional structure of Vh-ALDH with bound NADP(+) showed that the hydroxyl group of Thr175 forms a hydrogen bond to the 2'-phosphate of NADP(+). Replacement with glutamic acid or glutamine strengthened interactions with NAD(+) and indicated why threonine would be the preferred polar residue at the nucleotide recognition site in NADP(+)-specific aldehyde dehydrogenases. These results have shown that the size and the structure of the residue at the nucleotide recognition site play the key roles in differentiating between NAD(+) and NADP(+) interactions while the presence of a negative charge is responsible for the decrease in interactions with NADP(+) in Vh-ALDH.  相似文献   

12.
Uronate dehydrogenase from Agrobacterium tumefaciens (AtUdh) belongs to the short-chain dehydrogenase/reductase superfamily and catalyzes the oxidation of D-galacturonic acid and D-glucuronic acid with NAD(+) as a cofactor. We have determined the crystal structures of an apo-form of AtUdh, a ternary form in complex with NADH and product (substrate-soaked structure), and an inactive Y136A mutant in complex with NAD(+). The crystal structures suggest AtUdh to be a homohexamer, which has also been observed to be the major form in solution. The monomer contains a Rossmann fold, essential for nucleotide binding and a common feature of the short-chain dehydrogenase/reductase family enzymes. The ternary complex structure reveals a product, D-galactaro-1,5-lactone, which is bound above the nicotinamide ring. This product rearranges in solution to D-galactaro-1,4-lactone as verified by mass spectrometry analysis, which agrees with our previous NMR study. The crystal structure of the mutant with the catalytic residue Tyr-136 substituted with alanine shows changes in the position of Ile-74 and Ser-75. This probably altered the binding of the nicotinamide end of NAD(+), which was not visible in the electron density map. The structures presented provide novel insights into cofactor and substrate binding and the reaction mechanism of AtUdh. This information can be applied to the design of efficient microbial conversion of D-galacturonic acid-based waste materials.  相似文献   

13.
Random sequencing of a peppermint essential oil gland secretory cell cDNA library revealed a large number of clones that specified redox-type enzymes. Full-length acquisitions of each type were screened by functional expression in Escherichia coli using a newly developed in situ assay. cDNA clones encoding the monoterpene double-bond reductases (-)-isopiperitenone reductase and (+)-pulegone reductase were isolated, representing two central steps in the biosynthesis of (-)-menthol, the principal component of peppermint essential oil, and the first reductase genes of terpenoid metabolism to be described. The (-)-isopiperitenone reductase cDNA has an open reading frame of 942 nucleotides that encodes a 314 residue protein with a calculated molecular weight of 34,409. The recombinant reductase has an optimum pH of 5.5, and K(m) values of 1.0 and 2.2 microM for (-)-isopiperitenone and NADPH, respectively, with k(cat) of 1.3s(-1) for the formation of the product (+)-cis-isopulegone. The (+)-pulegone reductase cDNA has an open reading frame of 1026 nucleotides and encodes a 342 residue protein with a calculated molecular weight of 37,914. This recombinant reductase catalyzes the reduction of the 4(8)-double bond of (+)-pulegone to produce both (-)-menthone and (+)-isomenthone in a 55:45 ratio, has an optimum pH of 5.0, and K(m) values of 2.3 and 6.9 microM for (+)-pulegone and NADPH, respectively, with k(cat) of 1.8s(-1). Deduced sequence comparison revealed that these two highly substrate specific double-bond reductases show less than 12% identity. (-)-Isopiperitenone reductase is a member of the short-chain dehydrogenase/reductase superfamily and (+)-pulegone reductase is a member of the medium-chain dehydrogenase/reductase superfamily, implying very different evolutionary origins in spite of the similarity in substrates utilized and reactions catalyzed.  相似文献   

14.
Pyridine Nucleotide Transhydrogenase from Azotobacter vinelandii   总被引:5,自引:0,他引:5       下载免费PDF全文
A method is described for the partial purification of pyridine nucleotide transhydrogenase from Azotobacter vinelandii (ATCC 9104) cells. The most highly purified preparation catalyzes the reduction of 300 mumoles of nicotinamide adenine dinucleotide (NAD(+)) per min per mg of protein under the assay conditions employed. The enzyme catalyzes the reduction of NAD(+), deamino-NAD(+), and thio-NAD(+) with reduced nicotinamide adenine dinucleotide phosphate (NADPH) as hydrogen donor, and the reduction of nicotinamide adenine dinucleotide phosphate (NADP(+)) and thio-NAD(+) with reduced NAD (NADH) as hydrogen donor. The reduction of acetylpyridine AD(+), pyridinealdehyde AD(+), acetylpyridine deamino AD(+), and pyridinealdehydedeamino AD(+) with NADPH as hydrogen donor was not catalyzed. The enzyme catalyzes the transfer of hydrogen more readily from NADPH than from NADH with different hydrogen acceptors. The transfer of hydrogen from NADH to NADP(+) and thio-NAD(+) was markedly stimulated by 2'-adenosine monophosphate (2'-AMP) and inhibited by adenosine diphosphate (ADP), adenosine triphosphate (ATP), and phosphate ions. The transfer of hydrogen from NADPH to NAD(+) was only slightly affected by phosphate ions and 2'-AMP, except at very high concentrations of the latter reagent. In addition, the transfer of hydrogen from NADPH to thio-NAD(+) was only slightly influenced by 2'-AMP, ADP, ATP, and other nucleotides. The kinetics of the transhydrogenase reactions which utilized thio-NAD(+) as hydrogen acceptor and NADH or NADPH as hydrogen donor were studied in some detail. The results suggest that there are distinct binding sites for NADH and NAD(+) and perhaps a third regulator site for NADP(+) or 2'-AMP. The heats of activation for the transhydrogenase reactions were determined. The properties of this enzyme are compared with those of other partially purified transhydrogenases with respect to the regulatory functions of 2'-AMP and other nucleotides on the direction of flow of hydrogen between NAD(+) and NADP(+).  相似文献   

15.
The enzyme 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD) has an important role in androgen metabolism, catalyzing the interconversion of dihydrotestosterone (DHT) and 5alpha-androstane-3alpha,17beta-diol (3alpha-DIOL). The net direction of this interconversion will affect the amount of biologically active ligand available for androgen receptor binding. We hypothesize that in Leydig cells, differential expression of 3alpha-HSD enzymes favoring one of the two directions is a mechanism by which DHT levels are controlled. In order to characterize 3alpha-HSD in rat Leydig cells, the following properties were analyzed: rates of oxidation (3alpha-DIOL to DHT) and reduction (DHT to 3alpha-DIOL) and preference for the cofactors NADP(H) and NAD(H) (i.e., the oxidized and reduced forms of both pyridine nucleotides) in Leydig cells isolated on Days 21, 35, and 90 postpartum. Levels of 3alpha-HSD protein were measured by immunoblotting using an antibody directed against the liver type of the enzyme. Levels of 3alpha-HSD protein and rates of reduction were highest on Day 21 and lowest on Day 90. The opposite was true for the rate of 3alpha-HSD oxidation, which was barely detectable on Day 21 and highest on Day 90 (59.08 +/- 6.35 pmol/min per 10(6) cells, mean +/- SE). Therefore, the level of 3alpha-HSD protein detectable by liver enzyme was consistent with reduction but not with oxidation. There was a clear partitioning of NADP(H)-dependent activity into the cytosolic fraction of Leydig cells, whereas on Days 35 and 90, Leydig cells also contained a microsomal NAD(H)-activated 3alpha-HSD. We conclude that 1) the cytosolic 3alpha-HSD in Leydig cells on Day 21 behaves as a unidirectional NADPH-dependent reductase; 2) by Day 35, a microsomal NAD(H)-dependent enzyme activity is present and may account for predominance of 3alpha-HSD oxidation over reduction and the resultant high capacity of Leydig cells on Day 90 to synthesize DHT from 3alpha-DIOL.  相似文献   

16.
ADP-L-glycero-D-mannoheptose 6-epimerase is required for lipopolysaccharide inner core biosynthesis in several genera of Gram-negative bacteria. The enzyme contains both fingerprint sequences Gly-X-Gly-X-X-Gly and Gly-X-X-Gly-X-X-Gly near its N terminus, which is indicative of an ADP binding fold. Previous studies of this ADP-l-glycero-D-mannoheptose 6-epimerase (ADP-hep 6-epimerase) were consistent with an NAD(+) cofactor. However, the crystal structure of this ADP-hep 6-epimerase showed bound NADP (Deacon, A. M., Ni, Y. S., Coleman, W. G., Jr., and Ealick, S. E. (2000) Structure 5, 453-462). In present studies, apo-ADP-hep 6-epimerase was reconstituted with NAD(+), NADP(+), and FAD. In this report we provide data that shows NAD(+) and NADP(+) both restored enzymatic activity, but FAD could not. Furthermore, ADP-hep 6-epimerase exhibited a preference for binding of NADP(+) over NAD(+). The K(d) value for NADP(+) was 26 microm whereas that for NAD(+) was 45 microm. Ultraviolet circular dichroism spectra showed that apo-ADP-hep 6-epimerase reconstituted with NADP(+) had more secondary structure than apo-ADP-hep 6-epimerase reconstituted with NAD(+). Perchloric acid extracts of the purified enzyme were assayed with NAD(+)-specific alcohol dehydrogenase and NADP(+)-specific isocitric dehydrogenase. A sample of the same perchloric acid extract was analyzed in chromatographic studies, which demonstrated that ADP-hep 6-epimerase binds NADP(+) in vivo. A structural comparison of ADP-hep 6-epimerase with UDP-galactose 4-epimerase, which utilizes an NAD(+) cofactor, has identified the regions of ADP-hep 6-epimerase, which defines its specificity for NADP(+).  相似文献   

17.
We have established a simple kinetic model applicable to the enzyme cycling reaction for the determination of 3alpha-hydroxysteroids. This reaction was conducted under the reversible catalytic function of a single 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD) with nucleotide cofactors, thio-NAD(+) (one of the NAD(+) analogues) for the oxidation of 3alpha-hydroxysteroids and NADH for the reduction of 3-oxosteroids. This model was constructed based on the reaction mechanism of 3alpha-HSD, following an ordered bi-bi mechanism with cofactor binding first, under the assumption that the respective enzyme-cofactor complexes were distributed according to the initial ratio of thio-NAD(+) to NADH by the rapid equilibrium of both enzyme-cofactor complexes. The cycling rate in the new kinetic model could be expressed with the dissociation constants of enzyme-cofactor complexes and the initial concentrations of cofactors and enzyme. The cycling rate was verified by a comparison with the experimental data using 3alpha-HSD from Pseudomonas sp. B-0831. The results showed that the experimental data corresponded well with the results obtained from the kinetic model.  相似文献   

18.
An isocitrate dehydrogenase from Zymomonas mobilis was overexpressed in Escherichia coli as a fused protein (ZmIDH). The molecular mass of recombinant ZmIDH, together with its 6× His partner, was estimated to be 74 kDa by gel filtration chromatography, suggesting a homodimeric structure. The purified recombinant ZmIDH displayed maximal activity at 55 °C, pH 8.0 with Mn(2+) and pH 8.5 with Mg(2+). Heat inactivation studies showed that the recombinant ZmIDH was rapidly inactivated above 40 °C. In addition, the recombinant ZmIDH activity was completely dependent on the divalent cation and Mn(2+) was the most effective cation. The recombinant ZmIDH displayed a 165-fold (k(cat)/K(m)) preference for NAD(+) over NADP(+) with Mg(2+), and a 142-fold greater specificity for NAD(+) than NADP(+) with Mn(2+). Therefore, the recombinant ZmIDH has remarkably high coenzyme preference for NAD(+). The catalytic efficiency (k(cat)/K(m)) of the recombinant ZmIDH was found to be much lower than that of its NADP(+)-dependent counterparts. The poor performance of the recombinant ZmIDH in decarboxylating might be improved by protein engineering techniques, thus making ZmIDH a potential genetic modification target for the development of optimized Z. mobilis strains.  相似文献   

19.
Ribitol dehydrogenase from Zymomonas mobilis (ZmRDH) catalyzes the conversion of ribitol to d-ribulose and concomitantly reduces NAD(P)(+) to NAD(P)H. A systematic approach involving an initial sequence alignment-based residue screening, followed by a homology model-based screening and site-directed mutagenesis of the screened residues, was used to study the molecular determinants of the cofactor specificity of ZmRDH. A homologous conserved amino acid, Ser156, in the substrate-binding pocket of the wild-type ZmRDH was identified as an important residue affecting the cofactor specificity of ZmRDH. Further insights into the function of the Ser156 residue were obtained by substituting it with other hydrophobic nonpolar or polar amino acids. Substituting Ser156 with the negatively charged amino acids (Asp and Glu) altered the cofactor specificity of ZmRDH toward NAD(+) (S156D, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 10.9, where K(m)(,NAD) is the K(m) for NAD(+) and K(m)(,NADP) is the K(m) for NADP(+)). In contrast, the mutants containing positively charged amino acids (His, Lys, or Arg) at position 156 showed a higher efficiency with NADP(+) as the cofactor (S156H, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 0.11). These data, in addition to those of molecular dynamics and isothermal titration calorimetry studies, suggest that the cofactor specificity of ZmRDH can be modulated by manipulating the amino acid residue at position 156.  相似文献   

20.
1. The disappearance of pyridine nucleotides during incubation with mosquito homogenates proceeds through the hydrolysis of the pyrophosphate linkage of these compounds as demonstrated by the formation of NMN and AMP from NAD(+). This reaction was also demonstrated by the loss in the coenzyme functioning property of NAD(+) (yeast alcohol dehydrogenase reaction) without a concomitant loss in reactivity towards cyanide. Transglycosidase activity was not observed in the mosquito homogenates, and low concentrations of nicotinamide did not inhibit the NAD(+) splitting activity of these homogenates. These observations are all in accord with the presence in these homogenates of a NAD(+) pyrophosphatase rather than a NADase. 2. The NAD(+) pyrophosphatase is destroyed by boiling, is not heat-activated, and has a pH optimum at pH8.75. In addition to NAD(+), other dinucleotides such as NADP(+), the 3-acetylpyridine and thionicotinamide analogues of NAD(+) and the thionicotinamide analogue of NADP(+), function as substrates in the hydrolysis catalysed by the pyrophosphatase. 3. A decrease in the specific activity of NAD(+) pyrophosphatase was observed during larval development, and a barely detectable activity was found in the pupa and adult. 4. Enzyme activity per organism increased in the larva but decreased to a very low value in the pupa and adult. These results indicate that the decrease in specific activity was due to a decrease in enzyme concentration rather than an increase in amounts of protein.  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