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.
F Fan  J A Lorenzen  B V Plapp 《Biochemistry》1991,30(26):6397-6401
In the three-dimensional structures of enzymes that bind NAD or FAD, there is an acidic residue that interacts with the 2'- and 3'-hydroxyl groups of the adenosine ribose of the coenzyme. The size and charge of the carboxylate might repel the binding of the 2'-phosphate group of NADP and explain the specificity for NAD. In the NAD-dependent alcohol dehydrogenases, Asp-223 (horse liver alcohol dehydrogenase sequence) appears to have this role. The homologous residue in yeast alcohol dehydrogenase I (residue 201 in the protein sequence) was substituted with Gly, and the D223G enzyme was expressed in yeast, purified, and characterized. The wild-type enzyme is specific for NAD. In contrast, the D223G enzyme bound and reduced NAD+ and NADP+ equally well, but, relative to wild-type enzyme, the dissociation constant for NAD+ was increased 17-fold, and the reactivity (V/K) on ethanol was decreased to 1%. Even though catalytic efficiency was reduced, yeast expressing the altered or wild-type enzyme grew at comparable rates, suggesting that equilibration of NAD and NADP pools is not lethal. Asp-223 participates in binding NAD and in excluding NADP, but it is not the only residue important for determining specificity for coenzyme.  相似文献   

2.
NAD kinase is a ubiquitous enzyme that catalyzes the phosphorylation of NAD to NADP using ATP or inorganic polyphosphate (poly(P)) as phosphate donor, and is regarded as the only enzyme responsible for the synthesis of NADP. We present here the crystal structures of an NAD kinase from the archaeal organism Archaeoglobus fulgidus in complex with its phosphate donor ATP at 1.7 A resolution, with its substrate NAD at 3.05 A resolution, and with the product NADP in two different crystal forms at 2.45 A and 2.0 A resolution, respectively. In the ATP bound structure, the AMP portion of the ATP molecule is found to use the same binding site as the nicotinamide ribose portion of NAD/NADP in the NAD/NADP bound structures. A magnesium ion is found to be coordinated to the phosphate tail of ATP as well as to a pyrophosphate group. The conserved GGDG loop forms hydrogen bonds with the pyrophosphate group in the ATP-bound structure and the 2' phosphate group of the NADP in the NADP-bound structures. A possible phosphate transfer mechanism is proposed on the basis of the structures presented.  相似文献   

3.
The 2',3'-dialdehyde nicotinamide ribose derivatives of NAD (oNAD) and NADH (oNADH) have been prepared enzymatically from the corresponding 2',3'-dialdehyde analogs of NADP and NADPH. Pig heart NAD-dependent isocitrate dehydrogenase requires NAD as coenzyme but binds NADPH, as well as NADH, ADP, and ATP, at regulatory sites. Incubation of 1-3 mM oNAD or oNADH with this isocitrate dehydrogenase causes a time-dependent decrease in activity to a limiting value 40% that of the initial enzyme, suggesting that reaction does not occur at the catalytic coenzyme site. Upon varying the concentration of oNAD or oNADH from 0.2 to 3 mM, the inactivation rate constants increase in a nonlinear manner, consistent with reversible binding of oNAD and oNADH to the enzyme prior to covalent reaction. Inactivation is accompanied by incorporation of radioactive reagent with extrapolation to 0.54 mol [14C]oNAD or 0.45 mol [14C]oNADH/mol average enzyme subunit (or about 2 mol reagent/mol enzyme tetramer) when the enzyme is maximally inactivated; this value corresponds to the number of reversible binding sites for each of the natural ligands of isocitrate dehydrogenase. The protection against oNAD or oNADH inactivation by NADH, NADPH, and ADP (but not by isocitrate, NAD, or NADP) indicates that reaction occurs in the region of a nucleotide regulatory site. In contrast to the effects of oNAD and oNADH, oNADP and oNADPH cause total inactivation of the NAD-dependent isocitrate dehydrogenase, concomitant with incorporation, respectively, of about 3.5 mol [14C]oNADP or 1.3 mol [14C]oNADPH/mol average subunit. Reaction rates exhibit a linear dependence on [oNADP] or [oNADPH] and protection by natural ligands against inactivation is not striking. These results imply that oNADP and oNADPH are acting in this case as general chemical modifiers and indicate the importance of the free adenosine 2'-OH of oNAD and oNADH for specific labeling of the NAD-dependent isocitrate dehydrogenase. The new availability of 2',3'-dialdehyde nicotinamide ribose derivatives of NAD, NADH, NADP, and NADPH may allow selection of the appropriate reactive coenzyme analog for affinity labeling of a variety of dehydrogenases.  相似文献   

4.
Woodyer R  van der Donk WA  Zhao H 《Biochemistry》2003,42(40):11604-11614
Homology modeling was used to identify two particular residues, Glu175 and Ala176, in Pseudomonas stutzeri phosphite dehydrogenase (PTDH) as the principal determinants of nicotinamide cofactor (NAD(+) and NADP(+)) specificity. Replacement of these two residues by site-directed mutagenesis with Ala175 and Arg176 both separately and in combination resulted in PTDH mutants with relaxed cofactor specificity. All three mutants exhibited significantly better catalytic efficiency for both cofactors, with the best kinetic parameters displayed by the double mutant, which had a 3.6-fold higher catalytic efficiency for NAD(+) and a 1000-fold higher efficiency for NADP(+). The cofactor specificity was changed from 100-fold in favor of NAD(+) for the wild-type enzyme to 3-fold in favor of NADP(+) for the double mutant. Isoelectric focusing of the proteins in a nondenaturing gel showed that the replacement with more basic residues indeed changed the effective pI of the protein. HPLC analysis of the enzymatic products of the double mutant verified that the reaction proceeded to completion using either substrate and produced only the corresponding reduced cofactor and phosphate. Thermal inactivation studies showed that the double mutant was protected from thermal inactivation by both cofactors, while the wild-type enzyme was protected by only NAD(+). The combined results provide clear evidence that Glu175 and Ala176 are both critical for nicotinamide cofactor specificity. The rationally designed double mutant might be useful for the development of an efficient in vitro NAD(P)H regeneration system for reductive biocatalysis.  相似文献   

5.
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   

6.
The kinetic mechanism of the oxidative decarboxylation of 2R,3S-isopropylmalate by the NAD-dependent isopropylmalate dehydrogenase of Thermus thermophilus was investigated. Initial rate results typical of random or steady-state ordered sequential mechanisms are obtained for both the wild-type and two mutant enzymes (E87G and E87Q) regardless of whether natural or alternative substrates (2R-malate, 2R,3S-tartrate and/or NADP) are utilized. Initial rate data fail to converge on a rapid equilibrium-ordered pattern despite marked reductions in specificity (kcat/Km) caused by the mutations and alternative substrates. Although the inhibition studies alone might suggest an ordered kinetic mechanism with cofactor binding first, a detailed analysis reveals that the expected noncompetitive patterns appear uncompetitive because the dissociation constants from the ternary complexes are far smaller than those from the binary complexes. Equilibrium fluorescence studies both confirm the random binding of substrates and the kinetic estimates of the dissociation constants of the substrates from the binary complexes. The latter are not distributed markedly by the mutations at site 87. Mutations at site 87 do not affect the dissociation constants from the binary complexes, but do greatly increase the Michaelis constants, indicating that E87 helps stabilize the Michaelis complex of the wild-type enzyme. The available structural data, the patterns of the kinetics results, and the structure of a pseudo-Michaelis complex of the homologous isocitrate dehydrogenase of Escherichia coli suggest that E87 interacts with the nicotinamide ring.  相似文献   

7.
Nitrogenase activity in the photosynthetic bacterium Rhodospirillum rubrum is reversibly regulated by ADP-ribosylation of a specific arginine residue of dinitrogenase reductase based on the cellular nitrogen or energy status. In this paper, we have investigated the ability of nicotinamide adenine dinucleotide, NAD (the physiological ADP-ribose donor), and its analogs to support covalent modification of dinitrogenase reductase in vitro. R. rubrum dinitrogenase reductase can be modified by DRAT in the presence of 2 mM NAD, but not with 2 mM nicotinamide mononucleotide (NMN) or nicotinamide adenine dinucleotide phosphate (NADP). We also found that the apo- and the all-ferrous forms of R. rubrum dinitrogenase reductase are not substrates for covalent modification. In contrast, Azotobacter vinelandii dinitrogenase reductase can be modified by the dinitrogenase reductase ADP-ribosyl transferase (DRAT) in vitro in the presence of either 2 mM NAD, NMN or NADP as nucleotide donors. We found that: (1) a simple ribose sugar in the modification site of the A. vinelandii dinitrogenase reductase is sufficient to inactivate the enzyme, (2) phosphoADP-ribose is the modifying unit in the NADP-modified enzyme, and (3) the NMN-modified enzyme carries two ribose-phosphate units in one modification site. This is the first report of NADP- or NMN-dependent modification of a target protein by an ADP-ribosyl transferase.  相似文献   

8.
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   

9.
The oxidation of IMP to XMP is the rate-limiting step in the de novo synthesis of guanine ribonucleotides. This NAD-dependent reaction is catalyzed by the enzyme inosine monophosphate dehydrogenase (IMPDH). Based upon the recent structural determination of IMPDH complexed to oxidized IMP (XMP*) and the potent uncompetitive inhibitor mycophenolic acid (MPA), we have selected active site residues and prepared mutants of human type II IMPDH. The catalytic parameters of these mutants were determined. Mutations G326A, D364A, and the active site nucleophile C331A all abolish enzyme activity to less than 0.1% of wild type. These residues line the IMP binding pocket and are necessary for correct positioning of the substrate, Asp364 serving to anchor the ribose ring of the nucleotide. In the MPA/NAD binding site, significant loss of activity was seen by mutation of any residue of the triad Arg322, Asn303, Asp274 which form a hydrogen bonding network lining one side of this pocket. From a model of NAD bound to the active site consistent with the mutational data, we propose that these resides are important in binding the ribose ring of the nicotinamide substrate. Additionally, mutations in the pair Thr333, Gln441, which lies close to the xanthine ring, cause a significant drop in the catalytic activity of IMPDH. It is proposed that these residues serve to deliver the catalytic water molecule required for hydrolysis of the cysteine-bound XMP* intermediate formed after oxidation by NAD.  相似文献   

10.
The structure of crystalline porcine mitochondrial NADP-dependent isocitrate dehydrogenase (IDH) has been determined in complex with Mn2+-isocitrate. Based on structural alignment between this porcine enzyme and seven determined crystal structures of complexes of NADP with bacterial IDHs, Arg83, Thr311, and Asn328 were chosen as targets for site-directed mutagenesis of porcine IDH. The circular dichroism spectra of purified wild-type and mutant enzymes are similar. The mutant enzymes exhibit little change in Km for isocitrate or Mn2+, showing that these residues are not involved in substrate binding. In contrast, the Arg83 mutants, Asn328 mutants, and T311A exhibit 3-20-fold increase in the Km(NADP). We propose that Arg83 enhances NADP affinity by hydrogen bonding with the 3'-OH of the nicotinamide ribose, whereas Asn328 hydrogen bonds with N1 of adenine. The pH dependence of Vmax for Arg83 and Asn328 mutants is similar to that of wild-type enzyme, but for all the Thr311 mutants, pK(es) is increased from 5.2 in the wild type to approximately 6.0. We have previously attributed the pH dependence of Vmax to the deprotonation of the metal-bound hydroxyl of isocitrate in the enzyme-substrate complex, prior to the transfer of a hydride from isocitrate to NADP's nicotinamide moiety. Thr311 interacts with the nicotinamide ribose and is the closest of the target amino acids to the nicotinamide ring. Distortion of the nicotinamide by Thr311 mutation will likely be transmitted to Mn2+-isocitrate resulting in an altered pK(es). Because porcine and human mitochondrial NADP-IDH have 95% sequence identity, these results should be applicable to the human enzyme.  相似文献   

11.
Gastric tissues from amphibian Rana perezi express the only vertebrate alcohol dehydrogenase (ADH8) that is specific for NADP(H) instead of NAD(H). In the crystallographic ADH8-NADP+ complex, a binding pocket for the extra phosphate group of coenzyme is formed by ADH8-specific residues Gly223-Thr224-His225, and the highly conserved Leu200 and Lys228. To investigate the minimal structural determinants for coenzyme specificity, several ADH8 mutants involving residues 223 to 225 were engineered and kinetically characterized. Computer-assisted modeling of the docked coenzymes was also performed with the mutant enzymes and compared with the wild-type crystallographic binary complex. The G223D mutant, having a negative charge in the phosphate-binding site, still preferred NADP(H) over NAD(H), as did the T224I and H225N mutants. Catalytic efficiency with NADP(H) dropped dramatically in the double mutants, G223D/T224I and T224I/H225N, and in the triple mutant, G223D/T224I/H225N (kcat/KmNADPH = 760 mm-1 min-1), as compared with the wild-type enzyme (kcat/KmNADPH = 133330 mm-1 min-1). This was associated with a lower binding affinity for NADP+ and a change in the rate-limiting step. Conversely, in the triple mutant, catalytic efficiency with NAD(H) increased, reaching values (kcat/KmNADH = 155000 mm-1 min-1) similar to those of the wild-type enzyme with NADP(H). The complete reversal of ADH8 coenzyme specificity was therefore attained by the substitution of only three consecutive residues in the phosphate-binding site, an unprecedented achievement within the ADH family.  相似文献   

12.
The human mitochondrial NAD(P)+-dependent malic enzyme (m-NAD-ME) is a malic enzyme isoform with dual cofactor specificity and substrate binding cooperativity. Previous kinetic studies have suggested that Lys362 in the pigeon cytosolic NADP+-dependent malic enzyme has remarkable effects on the binding of NADP+ to the enzyme and on the catalytic power of the enzyme (Kuo, C. C., Tsai, L. C., Chin, T. Y., Chang, G.-G., and Chou, W. Y. (2000) Biochem. Biophys. Res. Commun. 270, 821-825). In this study, we investigate the important role of Gln362 in the transformation of cofactor specificity from NAD+ to NADP+ in human m-NAD-ME. Our kinetic data clearly indicate that the Q362K mutant shifted its cofactor preference from NAD+ to NADP+. The Km(NADP) and kcat(NADP) values for this mutant were reduced by 4-6-fold and increased by 5-10-fold, respectively, compared with those for the wild-type enzyme. Furthermore, up to a 2-fold reduction in Km(NADP)/Km(NAD) and elevation of kcat(NADP)/kcat(NAD) were observed for the Q362K enzyme. Mutation of Gln362 to Ala or Asn did not shift its cofactor preference. The Km(NADP)/Km(NAD) and kcat(NADP)/kcat(NAD) values for Q362A and Q362N were comparable with those for the wild-type enzyme. The DeltaG values for Q362A and Q362N with either NAD+ or NADP+ were positive, indicating that substitution of Gln with Ala or Asn at position 362 brings about unfavorable cofactor binding at the active site and thus significantly reduces the catalytic efficiency. Our data also indicate that the cooperative binding of malate became insignificant in human m-NAD-ME upon mutation of Gln362 to Lys because the sigmoidal phenomenon appearing in the wild-type enzyme was much less obvious that that in Q362K. Therefore, mutation of Gln362 to Lys in human m-NAD-ME alters its kinetic properties of cofactor preference, malate binding cooperativity, and allosteric regulation by fumarate. However, the other Gln362 mutants, Q362A and Q362N, have conserved malate binding cooperativity and NAD+ specificity. In this study, we provide clear evidence that the single mutation of Gln362 to Lys in human m-NAD-ME changes it to an NADP+-dependent enzyme, which is characteristic because it is non-allosteric, non-cooperative, and NADP+-specific.  相似文献   

13.
The aldehyde dehydrogenases (ALDHs) are a superfamily of multimeric enzymes which catalyse the oxidation of a broad range of aldehydes into their corresponding carboxylic acids with the reduction of their cofactor, NAD or NADP, into NADH or NADPH. At present, the only known structures concern NAD-dependent ALDHs. Three structures are available in the Protein Data Bank: two are tetrameric and the other is a dimer. We solved by molecular replacement the first structure of an NADP-dependent ALDH isolated from Streptococcus mutans, in its apo form and holo form in complex with NADP, at 1.8 and 2.6 A resolution, respectively. Although the protein sequence shares only approximately 30 % identity with the other solved tetrameric ALDHs, the structures are very similar. However, a large local conformational change in the region surrounding the 2' phosphate group of the adenosine moiety is observed when the enzyme binds NADP, in contrast to the NAD-dependent ALDHs.Structure and sequence analyses reveal several properties. A small number of residues seem to determine the oligomeric state. Likewise, the nature (charge and volume) of the residue at position 180 (Thr in ALDH from S. mutans) determines the cofactor specificity in comparison with the structures of NAD-dependent ALDHs. The presence of a hydrogen bond network around the cofactor not only allows it to bind to the enzyme but also directs the side-chains in a correct orientation for the catalytic reaction to take place. Moreover, a specific part of this network appears to be important in substrate binding. Since the enzyme oxidises the same substrate, glyceraldehyde-3-phosphate (G3P), as NAD-dependent phosphorylating glyceraldehyde-3-phosphate dehydrogenases (GAPDH), the active site of GAPDH was compared with that of the S. mutans ALDH. It was found that Arg103, Arg283 and Asp440 might be key residues for substrate binding.  相似文献   

14.
For better understanding of the coenzyme specificity in NAD-dependent MDH (tMDH) from Thermus flavus AT-62, we determined the crystal structures of tMDH-NADP(H) complex at maximally 1.65 A resolution. The overall structure is almost the same as that of the tMDH-NADH complex. However, NADP(H) binds to tMDH in the reverse orientation, where adenine occupies the position near the catalytic center and nicotinamide is positioned at the adenine binding site of the tMDH-NADH complex. Consistent with this, kinetic analysis of the malate-oxidizing reaction revealed that NADP(+) inhibited tMDH at high concentrations. This has provided the first evidence for the alternative binding mode of the nicotinamide coenzyme, that has pseudo-symmetry in its structure, in a single enzyme.  相似文献   

15.
Using the purified NADP(H)-binding domain of proton-translocating Escherichia coli transhydrogenase (ecIII) overexpressed in (15)N- and (2)H-labeled medium, together with the purified NAD(H)-binding domain from E. coli (ecI), the interface between ecIII and ecI, the NADP(H)-binding site and the influence on the interface by NAD(P)(H) was investigated in solution by NMR chemical shift mapping. Mapping of the NADP(H)-binding site showed that the NADP(H) substrate is bound to ecIII in an extended conformation at the C-terminal end of the parallel beta-sheet. The distribution of chemical shift perturbations in the NADP(H)-binding site, and the nature of the interaction between ecI and ecIII, indicated that the nicotinamide moiety of NADP(H) is located near the loop comprising residues P346-G353, in agreement with the recently determined crystal structures of bovine [Prasad, G. S., et al. (1999) Nat. Struct. Biol. 6, 1126-1131] and human heart [White, A. W., et al. (2000) Structure 8, 1-12] transhydrogenases. Further chemical shift perturbation analysis also identified regions comprising residues G389-I406 and G430-V434 at the C-terminal end of ecIII's beta-sheet as part of the ecI-ecIII interface, which were regulated by the redox state of the NAD(P)(H) substrates. To investigate the role of these loop regions in the interaction with domain I, the single cysteine mutants T393C, R425C, G430C, and A432C were generated in ecIII and the transhydrogenase activities of the resulting mutant proteins characterized using the NAD(H)-binding domain I from Rhodospirillum rubrum (rrI). All mutants except R425C showed altered NADP(H) binding and domain interaction properties. In contrast, the R425C mutant showed almost exclusively changes in the NADP(H)-binding properties, without changing the affinity for rrI. Finally, by combining the above conclusions with information obtained by a further characterization of previously constructed mutants, the implications of the findings were considered in a mechanistic context.  相似文献   

16.
Sequence alignment of pig mitochondrial NADP-dependent isocitrate dehydrogenase with eukaryotic (human, rat, and yeast) and Escherichia coli isocitrate dehydrogenases reveals that Tyr316 is completely conserved and is equivalent to the E. coli Tyr345, which interacts with the 2'-phosphate of NADP in the crystal structure [Hurley et al., Biochemistry 30 (1991) 8671-8678]. Lys321 is also completely conserved in the five isocitrate dehydrogenases. Either an arginine or lysine residue is found among the enzymes from other species at the position corresponding to the pig enzyme Arg314. While Arg323 is not conserved among all species, its proximity to the coenzyme site makes it a good candidate for investigation. The importance of these four amino acids to the function of pig mitochondrial NADP-isocitrate dehydrogenase was studied by site-directed mutagenesis. Mutants (R314Q, Y316F, Y316L, K321Q, and R323Q) were generated by a megaprimer polymerase chain reaction method. Wild-type and mutant enzymes were expressed in E. coli and purified to homogeneity. All mutant and wild-type enzymes exhibited comparable molecular weights indicative of the dimeric enzyme. Mutations do not cause an appreciable change in enzyme secondary structure as revealed by circular dichroism measurements. The kinetic parameters (V(max) and K(M) values) of K321Q and R323Q are similar to those of wild-type, indicating that Lys321 and Arg323 are not involved in enzyme function. R314Q exhibits a 10-fold increase in K(M) for NADP as compared to that of wild-type, while they have comparable V(max) values. These results suggest that Arg314 contributes to the affinity between the enzyme and NADP. The hydroxyl group of Tyr316 is not required for enzyme function since Y316F exhibits similar kinetic parameters to those of wild-type. Y316L shows a 4-fold increase in K(M) for NADP and a decrease in V(max) as compared to wild-type, suggesting that the aromatic ring of the Tyr of isocitrate dehydrogenase contributes to the affinity for coenzyme, as well as to catalysis. The K(i) for NAD of R314Q, Y316F, and Y316L is comparable to that of wild-type, indicating that the Arg314 and Tyr316 may be located near the 2'-phosphate of enzyme-bound NADP.  相似文献   

17.
The 2.1 A resolution crystal structure of flavin reductase P with the inhibitor nicotinamide adenine dinucleotide (NAD) bound in the active site has been determined. NAD adopts a novel, folded conformation in which the nicotinamide and adenine rings stack in parallel with an inter-ring distance of 3.6 A. The pyrophosphate binds next to the flavin cofactor isoalloxazine, while the stacked nicotinamide/adenine moiety faces away from the flavin. The observed NAD conformation is quite different from the extended conformations observed in other enzyme/NAD(P) structures; however, it resembles the conformation proposed for NAD in solution. The flavin reductase P/NAD structure provides new information about the conformational diversity of NAD, which is important for understanding catalysis. This structure offers the first crystallographic evidence of a folded NAD with ring stacking, and it is the first enzyme structure containing an FMN cofactor interacting with NAD(P). Analysis of the structure suggests a possible dynamic mechanism underlying NADPH substrate specificity and product release that involves unfolding and folding of NADP(H).  相似文献   

18.
The ubiquitous redox cofactors nicotinamide adenine dinucleotides [NAD and NADP] are very similar molecules, despite their participation in substantially different biochemical processes. NADP differs from NAD in only the presence of an additional phosphate group esterified to the 2′-hydroxyl group of the ribose at the adenine end and yet NADP is confined with few exceptions to the reactions of reductive biosynthesis, whereas NAD is used almost exclusively in oxidative degradations. The discrimination between NAD and NADP is therefore an impressive example of the power of molecular recognition by proteins. The many known tertiary structures of NADP complexes affords the possibility for an analysis of their discrimination. A systematic analysis of several crystal structures of NAD(P)-protein complexes show that: 1) the NADP coenzymes are more flexible in conformation than those of NAD; 2) although the protein-cofactor interactions are largely conserved in the NAD complexes, they are quite variable in those of NADP; and 3) in both cases the pocket around the nicotinamide moiety is substrate dependent. The conserved and variable interactions between protein and cofactors in the respective binding pockets are reported in detail. Discrimination between NAD and NADP is essentially a consequence of the overall pocket and not of a few residues. A clear fingerprint in NAD complexes is a carboxylate side chain that chelates the diol group at the ribose near the adenine, whereas in NADP complexes an arginine side chain faces the adenine plane and interacts with the phosphomonoester. The latter type of interaction might be a general feature of recognition of nucleotides by proteins. Other features such as strand-like hydrogen bonding between the NADP diphosphate moeties and the protein are also significant. The NADP binding pocket properties should prove useful in protein engineering and design. © 1997 Wiley-Liss Inc.  相似文献   

19.
Two of the three metabolic subtypes of species utilizing C4-pathway photosynthesis are defined by high activities of either NADP malic enzyme (NADP malic enzyme type) or a coenzyme A (CoA)- and acetyl-CoA-activated NAD malic enzyme (NAD malic enzyme type). These enzymes function to decarboxylate malate as an integral part of the photosynthetic process. Leaves of NADP malic enzyme-type species also contain significant NAD-dependent malic enzyme activity. The purpose of the present study was to examine the nature and photosynthetic role of this activity. With Zea mays, this NAD-dependent activity was found to vary widely in fresh leaf extracts. Incubating extracts at 25 °C resulted in a disproportionate increase in NAD activity so that the final ratio of NADP to NAD activity was always about 5. Strong evidence was provided that the NADP and NAD malic enzyme activities in Z. mays extracts were catalyzed by the same enzyme. These activities remained associated during purification and were coincident after polyacrylamide gel electrophoresis. The pH optimum for NAD-dependent activity was about 7.1, compared with 8.3 for NADP malic enzyme activity. Other properties of the NAD-dependent activity are described, a particularly notable feature being the inhibition of this activity by less than 1 μm NADP and NADPH. Evidence is provided that the NADP malic enzyme of several other NADP malic enzyme-type C4 species also has associated activity toward NAD. We concluded that the NAD-dependent malic enzyme activity would have no significant function in photosynthesis.  相似文献   

20.
Nicotinamide adenine dinucleotide (NAD) and nicotinamide adenine dinucleotide phosphate (NADP), which is derived from NAD, have important roles as a redox carriers in metabolism. A combination of de novo and salvage pathways contribute to the biosynthesis of NAD in all organisms. The pathways and enzymes of the NAD salvage pathway in yeast and animals, which diverge at nicotinamide, have been extensively studied. Yeast cells convert nicotinamide to nicotinic acid, while mammals lack the enzyme nicotinamidase and instead convert nicotinamide to nicotinamide mononucleotide. Here we show that Arabidopsis thaliana gene At2g22570 encodes a nicotinamidase, which is expressed in all tissues, with the highest levels observed in roots and stems. The 244-residue protein, designated AtNIC1, converts nicotinamide to nicotinic acid and has a Km value of 118 +/- 17 microM and a Kcat value of 0.93 +/- 0.13 sec(-1). Plants homozygous for a null AtNIC1 allele, nic1-1, have lower levels of NAD and NADP under normal growth conditions, indicating that AtNIC1 participates in a yeast-type NAD salvage pathway. Mutant plants also exhibit hypersensitivity to treatments of abscisic acid and NaCl, which is correlated with their inability to increase the cellular levels of NAD(H) under these growth conditions, as occurs in wild-type plants. We also show that the growth of the roots of wild-type but not nic1-1 mutant plants is inhibited and distorted by nicotinamide.  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