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 203 毫秒
1.
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   

2.
The F420-dependent alcohol dehydrogenase (ADH) of Methanogenium liminatans and the NADP(+)-dependent ADH of Methanobacterium palustre were purified to homogeneity. The native F420-dependent ADH of Mg. liminatans had a molecular mass of 150 kDa and consisted of four (presumably identical) subunits with a mass of 39 kDa. The temperature optimum was 42 degrees C, the optimum pH 6.0 and NaCl or KCl were inhibitory. The NADP(+)-dependent ADH of Mb. palustre had a molecular mass of 175 kDa and consisted also of four (presumably identical) subunits with a mass of 44 kDa. The temperature optimum was 60 degrees C, the optimum pH 8.0 and optimal activity was observed in the presence of 500 mM NaCl or KCl. The ADHs of both organisms catalysed the oxidation of various secondary and cyclic alcohols to the corresponding ketones and the reverse reaction. No primary alcohols were apparently oxidized. The NADP(+)-dependent ADH of Mb. palustre contained 4-8 mol atoms zinc/mol enzyme and was inhibited by low concentrations of iodoacetate and 4-hydroxymercuribenzoate, whereas the F420-dependent ADH of Mg. liminatans presumably contained no zinc ions and was inhibited by 1,10-phenanthroline or high concentrations (e.g. 100 microM) of 4-hydroxymercuribenzoate. Polyclonal antibodies against the NADP(+)-dependent ADH of Mb. palustre precipitated only the homologous ADH. A precipitation of the NADP(+)-dependent ADH of Methanocorpusculum parvum required a 10-fold higher antibody concentration, showing at least a distant relationship of both ADHs. Antibodies against the NADP(+)-dependent ADH of Mcp. parvum, however, formed precipitates with the homologous ADH of Mcp. parvum and with the NADP(+)-dependent ADH of Mb. palustre. They also formed precipitates with the ADH of Thermoanaerobium brockii, which is not related to methane bacteria. Antibodies against the F420-dependent ADH of Mg. liminatans reacted only with the homologous enzyme and did not form precipitates with NADP(+)-dependent ADHs. No immunological relation of the NADP(+)- or F420-dependent ADHs of methanogens with ADH of yeast or horse liver was found. In accordance with the immunological data, the N-terminal amino acid sequences of the NADP(+)-dependent ADHs of Mb. palustre and Mcp. parvum had a high degree of similarity, whereas the N-terminal amino acid sequence of the ADH of Mg. liminatans revealed no similarity with the two NADP(+)-dependent enzymes.  相似文献   

3.
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   

4.
The amphibian enzyme ADH8, previously named class IV-like, is the only known vertebrate alcohol dehydrogenase (ADH) with specificity towards NADP(H). The three-dimensional structures of ADH8 and of the binary complex ADH8-NADP(+) have been now determined and refined to resolutions of 2.2A and 1.8A, respectively. The coenzyme and substrate specificity of ADH8, that has 50-65% sequence identity with vertebrate NAD(H)-dependent ADHs, suggest a role in aldehyde reduction probably as a retinal reductase. The large volume of the substrate-binding pocket can explain both the high catalytic efficiency of ADH8 with retinoids and the high K(m) value for ethanol. Preference of NADP(H) appears to be achieved by the presence in ADH8 of the triad Gly223-Thr224-His225 and the recruitment of conserved Lys228, which define a binding pocket for the terminal phosphate group of the cofactor. NADP(H) binds to ADH8 in an extended conformation that superimposes well with the NAD(H) molecules found in NAD(H)-dependent ADH complexes. No additional reshaping of the dinucleotide-binding site is observed which explains why NAD(H) can also be used as a cofactor by ADH8. The structural features support the classification of ADH8 as an independent ADH class.  相似文献   

5.
Three alcohol dehydrogenases have been identified in Acinetobacter calcoaceticus sp. strain HO1-N: an NAD(+)-dependent enzyme and two NADP(+)-dependent enzymes. One of the NADP(+)-dependent alcohol dehydrogenases was partially purified and was specific for long-chain substrates. With tetradecanol as substrate an apparent Km value of 5.2 microM was calculated. This enzyme has a pI of 4.5 and a molecular mass of 144 kDa. All three alcohol dehydrogenases were constitutively expressed. Three aldehyde dehydrogenases were also identified: an NAD(+)-dependent enzyme, an NADP(+)-dependent enzyme and one which was nucleotide independent. The NAD(+)-dependent enzyme represented only 2% of the total activity and was not studied further. The NADP(+)-dependent enzyme was strongly induced by growth of cells on alkanes and was associated with hydrocarbon vesicles. With tetradecanal as substrate an apparent Km value of 0.2 microM was calculated. The nucleotide-independent aldehyde dehydrogenase could use either Würster's Blue or phenazine methosulphate (PMS) as an artificial electron acceptor. This enzyme represents approximately 80% of the total long-chain aldehyde oxidizing activity within the cell when the enzymes were induced by growing the cells on hexadecane. It is particulate but can be solubilized using Triton X-100. The enzyme has an apparent Km of 0.36 mM for decanal.  相似文献   

6.
An NADP(+)-dependent D-xylose dehydrogenase from pig liver cytosol was purified about 2000-fold to apparent homogeneity with a yield of 15% and specific activity of 6 units/mg of protein. An Mr value of 62,000 was obtained by gel filtration. PAGE in the presence of SDS gave an Mr value of 32,000, suggesting that the native enzyme is a dimer of similar or identical subunits. D-Xylose, D-ribose, L-arabinose, 2-deoxy-D-glucose, D-glucose and D-mannose were substrates in the presence of NADP+ but the specificity constant (ratio kcat./Km(app.)) is, by far, much higher for D-xylose than for the other sugars. The enzyme is specific for NADP+; NAD+ is not reduced in the presence of D-xylose or other sugars. Initial-velocity studies for the forward direction with xylose or NADP+ concentrations varied at fixed concentrations of the nucleotide or the sugar respectively revealed a pattern of parallel lines in double-reciprocal plots. Km values for D-xylose and NADP+ were 8.8 mM and 0.99 mM respectively. Dead-end inhibition studies to confirm a ping-pong mechanism showed that NAD+ acted as an uncompetitive inhibitor versus NADP+ (Ki 5.8 mM) and as a competitive inhibitor versus xylose. D-Lyxose was a competitive inhibitor versus xylose and uncompetitive versus NADP+. These results fit better to a sequential compulsory ordered mechanism with NADP+ as the first substrate, but a ping-pong mechanism with xylose as the first substrate has not been ruled out. The presence of D-xylose dehydrogenase suggests that in mammalian liver D-xylose is utilized by a pathway other than the pentose phosphate pathway.  相似文献   

7.
The R‐specific alcohol dehydrogenase from Lactobacillus brevis (Lb‐ADH) catalyzes the enantioselective reduction of prochiral ketones to the corresponding secondary alcohols. It is stable and has broad substrate specificity. These features make this enzyme an attractive candidate for biotechnological applications. A drawback is its preference for NADP(H) as a cofactor, which is more expensive and labile than NAD(H). Structure‐based computational protein engineering was used to predict mutations to alter the cofactor specificity of Lb‐ADH. Mutations were introduced into Lb‐ADH and tested against the substrate acetophenone, with either NAD(H) or NADP(H) as cofactor. The mutant Arg38Pro showed fourfold increased activity with acetophenone and NAD(H) relative to the wild type. Both Arg38Pro and wild type exhibit a pH optimum of 5.5 with NAD(H) as cofactor, significantly more acidic than with NADP(H). These and related Lb‐ADH mutants may prove useful for the green synthesis of pharmaceutical precursors.  相似文献   

8.
Formate dehydrogenase from Candida boidinii (CboFDH) catalyses the oxidation of formate anion to carbon dioxide with concomitant reduction of NAD(+) to NADH. CboFDH is highly specific to NAD(+) and virtually fails to catalyze the reaction with NADP(+). Based on structural information for CboFDH, the loop region between beta-sheet 7 and alpha-helix 10 in the dinucleotide-binding fold was predicted as a principal determinant of coenzyme specificity. Sequence alignment with other formate dehydrogenases revealed two residues (Asp195 and Tyr196) that could account for the observed coenzyme specificity. Positions 195 and 196 were subjected to two rounds of site-saturation mutagenesis and screening and enabled the identification of a double mutant Asp195Gln/Tyr196His, which showed a more than 2 x 10(7)-fold improvement in overall catalytic efficiency with NADP(+) and a more than 900-fold decrease in the efficiency with NAD(+) as cofactors. The results demonstrate that the combined polar interactions and steric factors comprise the main structural determinants responsible for coenzyme specificity. The double mutant Asp195Gln/Tyr196His was tested for practical applicability in a cofactor recycling system composed of cytochrome P450 monooxygenase from Bacillus subtilis, (CYP102A2), NADP(+), formic acid and omega-(p-nitrophenyl)dodecanoic acid (12-pNCA). Using a 1250-fold excess of 12-pNCA over NADP(+) the first order rate constant was determined to be equal to k(obs) = 0.059 +/- 0.004 min(-1).  相似文献   

9.
In the oxidation of methylglyoxal by 2-oxoaldehyde dehydrogenase, the apparent Km value for NADP+ was about 2.5 times lower than the corresponding Km for NAD+; the apparent Km values for methylglyoxal and for the amine activator L-2-aminopropan-1-ol, with NADP+ as cofactor, were also different from those obtained with NAD+. In the presence of NADP+, the enzyme was not activated by P1, in contrast with the activation of the enzyme when NAD+ was used. The significance of the results is discussed.  相似文献   

10.
In Sm. lipolytica one NAD+-dependent and three NADP+-dependent alcohol dehydrogenases are detectable by polyacrylamide gelelectrophoresis. The NAD+-dependent ADH (ADH I), with a molecular weight of 240,000 daltons, reacts more intensively with long-chain alcohols (octanol) than with short-chain alcohols (methanol, ethanol). The ADH I is not or only minimally subject to glucose repression. Besides the ADH I band no additional inducible NAD+-dependent ADH band is gel-electrophoretically detectable during growth of yeast cells in medium containing ethanol or paraffin. The ADH I band is very probably formed by two ADH enzymes with the same electrophoretic mobility. The NADP+-dependent alcohol dehydrogenases (ADH II--IV) react with methanol, ethanol and octanol with different intensity. In polyacrylamide gradients two bands of NADP+-dependent ADH are detectable: one with a molecular weight of 70,000 daltons and the other with 120,000 daltons. The occurrence of the three NADP+-dependent alcohol dehydrogenases is regulated by the carbon source of the medium. Sm. lipolytica shows a high tolerance against allylalcohol. Resistant mutants can be isolated only at concentrations of 1 M allylalcohol in the medium. All isolates of allylalcohol-resistant mutants show identical growth in medium containing ethanol as the wild type strain.  相似文献   

11.
The N-1 position of the adenine ring of NADP was selectively alkylated by the reaction of 2',3'-cyclic NADP with 3-propiolactone to yield 2',3'-cyclic 1-(2-carboxyethyl)-NADP (I). Derivative I was converted to a mixture of the isomers of N6-(2-carboxyethyl)-NADP with their phosphate groups at the 2' or 3' position (IIa and IIb) by chemical reduction, alkaline rearrangement and chemical reoxidation. Carbodiimide coupling of the mixture of IIa and IIb to alpha, omega-diaminopoly(ethylene glycol) gave the 2', 3'-cyclic derivative of poly(ethylene glycol)-bound NADP (III), which was enzymically hydrolyzed to yield poly(ethylene glycol)-bound NADP (PEG-NADP). PEG-NADP has good cofactor activity (16-100% of that of NADP) for NADP-specific and NAD(P)-specific dehydrogenases except isocitrate and glucose dehydrogenases. For NAD-specific enzymes, PEG-NADP has higher cofactor activity than NADP: for horse liver alcohol dehydrogenase, the cofactor activity of PEG-NADP is 40 times that of NADP and 14% of that of NAD. Kinetic studies show that for most of enzymes tested, Km values for PEG-NADP are larger than those for NADP and V values for PEG-NADP are similar to those for NADP. PEG-NADP proved to be applicable in a continuous enzyme reactor, in which reactions of glutamate dehydrogenase and glucose-6-phosphate dehydrogenase were coupled by the recycling of PEG-NADP.  相似文献   

12.
Malic enzymes are widely distributed in nature, and have important biological functions. They catalyze the oxidative decarboxylation of malate to produce pyruvate and CO(2) in the presence of divalent cations (Mg(2+), Mn(2+)). Most malic enzymes have a clear selectivity for the dinucleotide cofactor, being able to use either NAD(+) or NADP(+), but not both. Structural studies of the human mitochondrial NAD(+)-dependent malic enzyme established that malic enzymes belong to a new class of oxidative decarboxylases. Here we report the crystal structure of the pigeon cytosolic NADP(+)-dependent malic enzyme, in a closed form, in a quaternary complex with NADP(+), Mn(2+), and oxalate. This represents the first structural information on an NADP(+)-dependent malic enzyme. Despite the sequence conservation, there are large differences in several regions of the pigeon enzyme structure compared to the human enzyme. One region of such differences is at the binding site for the 2'-phosphate group of the NADP(+) cofactor, which helps define the cofactor selectivity of the enzymes. Specifically, the structural information suggests Lys362 may have an important role in the NADP(+) selectivity of the pigeon enzyme, confirming our earlier kinetic observations on the K362A mutant. Our structural studies also revealed differences in the organization of the tetramer between the pigeon and the human enzymes, although the pigeon enzyme still obeys 222 symmetry.  相似文献   

13.
Coenzymic activities of the following NADP derivatives were investigated: 2'-O-(2-carboxyethyl)phosphono-NAD (I), N6-(2-carboxyethyl)-NADP (II), 2'-O-(2-carboxyethyl)phosphono-N6-(2-carboxyethyl)-NAD (III), 2'-O-[N-(2-aminoethyl)carbamoylethyl]phosphono-NAD (IV), N6-[N-(2-aminoethyl)carbamoylethyl]-NADP (Va), 2',3'-cyclic NADP, and 3'-NADP. Derivatives I and IV show the effects of modification at the 2'-phosphate group, and derivatives II and Va show those at the 6-amino group of NADP. As for enzymes, alcohol, isocitrate, 6-phosphogluconate, glucose, glucose-6-phosphate, and glutamate dehydrogenases were used. These enzymes were grouped on the basis of the ratio of the activities for NAD and NADP into NADP-specific enzymes (ratio less than 0.01), NAD(P)-specific enzymes (0.01 less than ratio less than 100), and NAD-specific enzymes (ratio greater than 100). For NADP-specific enzymes, modifications at the 2'-phosphate group of NADP caused great loss of cofactor activity. The relative cofactor activities (NADP = 100%) of derivatives I and IV for these enzymes were 0.5-20 and 0.01-0.5%, respectively. On the other hand, NAD(P)-specific enzymes showed several types of responses to the NADP derivatives. The relative cofactor activities of I and IV for Leuconostoc mesenteroides and Bacillus stearothermophilus glucose-6-phosphate dehydrogenases and beef liver glutamate dehydrogenase were 60-200%; whereas, for B. megaterium glucose dehydrogenase and L. mesenteroides alcohol dehydrogenase, the values were 0.8-8%. For NAD-specific enzymes, these values were 20-50%. The relative cofactor activities of 2',3'-cyclic NADP and 3'-NADP were very low (less than 0.2%) except for beef liver glutamate dehydrogenase, B. stearothermophilus glucose-6-phosphate dehydrogenase, and horse liver alcohol dehydrogenase. Kinetic studies showed that the losses of the cofactor activity of NADP by these modifications were mainly due to the increase of the Km value. The mechanisms of coenzyme specificity of dehydrogenases are discussed. Unlike the 2'-phosphate group, the 6-amino group is common to NAD and NADP, and the effects of modification at the 6-amino group were independent of the coenzyme specificity of enzymes used for the assay. Derivatives II and Va had high relative cofactor activities (65-130%) for most of the enzymes except for isocitrate and glucose dehydrogenases (less than 1%) and L. mesenteroides alcohol dehydrogenase (20-60%). The cofactor activity of derivative III was generally lower than those of I and II.  相似文献   

14.
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   

15.
The crystal structure of NAD(+)-dependent alcohol dehydrogenase from Bacillus stearothermophilus strain LLD-R (htADH) was determined using X-ray diffraction data at a resolution of 2.35 A. The structure of homotetrameric htADH is highly homologous to those of bacterial and archaeal homotetrameric alcohol dehydrogenases (ADHs) and also to the mammalian dimeric ADHs. There is one catalytic zinc atom and one structural zinc atom per enzyme subunit. The enzyme was crystallized as a binary complex lacking the nicotinamide adenine dinucleotide (NAD(+)) cofactor but including a zinc-coordinated substrate analogue trifluoroethanol. The binary complex structure is in an open conformation similar to ADH structures without the bound cofactor. Features important for the thermostability of htADH are suggested by a comparison with a homologous mesophilic enzyme (55% identity), NAD(+)-dependent alcohol dehydrogenase from Escherichia coli. To gain insight into the conformational change triggered by NAD(+) binding, amide hydrogen-deuterium exchange of htADH, in the presence and absence of NAD(+), was studied by HPLC-coupled electrospray mass spectrometry. When the deuteron incorporation of the protein-derived peptides was analyzed, it was found that 9 of 21 peptides show some decrease in the level of deuteron incorporation upon NAD(+) binding, and another 4 peptides display slower exchange rates. With one exception (peptide number 8), none of the peptides that are altered by bound NAD(+) are in contact with the alcohol-substrate-binding pocket. Furthermore, peptides 5 and 8, which are located outside the NAD(+)-binding pocket, are notable by displaying changes upon NAD(+) binding. This suggests that the transition from the open to the closed conformation caused by cofactor binding has some long-range effects on the protein structure and dynamics.  相似文献   

16.
The amino acid sequence is reported for CNBr and tryptic peptide fragments of the NAD(+)-dependent glutamate dehydrogenase of Clostridium symbiosum. Together with the N-terminal sequence, these make up about 75% of the total sequence. The sequence shows extensive similarity with that of the NADP(+)-dependent glutamate dehydrogenase of Escherichia coli (52% identical residues out of the 332 compared) allowing confident placing of the peptide fragments within the overall sequence. This demonstrated sequence similarity with the E. coli enzyme, despite different coenzyme specificity, is much greater than the similarity (31% identities) between the GDH's of C. symbiosum and Peptostreptococcus asaccharolyticus, both NAD(+)-linked. The evolutionary implications are discussed. In the 'fingerprint' region of the nucleotide binding fold the sequence Gly X Gly X X Ala is found, rather than Gly X Gly X X Gly. The sequence found here has previously been associated with NADP+ specificity and its finding in a strictly NAD(+)-dependent enzyme requires closer examination of the function of this structural motif.  相似文献   

17.
The Bacillus methanolicus methanol dehydrogenase (MDH) is a decameric nicotinoprotein alcohol dehydrogenase (family III) with one Zn(2+) ion, one or two Mg(2+) ions, and a tightly bound cofactor NAD(H) per subunit. The Mg(2+) ions are essential for binding of cofactor NAD(H) in MDH. A B. methanolicus activator protein strongly stimulates the relatively low coenzyme NAD(+)-dependent MDH activity, involving hydrolytic removal of the NMN(H) moiety of cofactor NAD(H) (Kloosterman, H., Vrijbloed, J. W., and Dijkhuizen, L. (2002) J. Biol. Chem. 277, 34785-34792). Members of family III of NAD(P)-dependent alcohol dehydrogenases contain three unique, conserved sequence motifs (domains A, B, and C). Domain C is thought to be involved in metal binding, whereas the functions of domains A and B are still unknown. This paper provides evidence that domain A constitutes (part of) a new magnesium-dependent NAD(P)(H)-binding domain. Site-directed mutants D100N and K103R lacked (most of the) bound cofactor NAD(H) and had lost all coenzyme NAD(+)-dependent MDH activity. Also mutants G95A and S97G were both impaired in cofactor NAD(H) binding but retained coenzyme NAD(+)-dependent MDH activity. Mutant G95A displayed a rather low MDH activity, whereas mutant S97G was insensitive to activator protein but displayed "fully activated" MDH reaction rates. The various roles of these amino acid residues in coenzyme and/or cofactor NAD(H) binding in MDH are discussed.  相似文献   

18.
Geraniol dehydrogenase (GeDH), which plays an important role in the biosynthesis of neral, an alarm pheromone, was purified from the astigmatid mite Carpoglyphus lactis. The enzyme was obtained in an apparently homogeneous and active form after 1879-fold purification through seven steps of chromatography. Car. lactis GeDH was determined to be a monomer in its active form with a relative molecular mass of 42 800, which is a unique subunit structure in comparison with already established alcohol dehydrogenases. Car. lactis GeDH oxidized geraniol into geranial in the presence of NAD+. NADP+ was ineffective as a cofactor, suggesting that Car. lactis GeDH is an NAD+-dependent alcohol dehydrogenase. The optimal pH and temperature for geraniol oxidation were determined to be pH 9.0 and 25 degrees C, respectively. The Km values for geraniol and NAD+ were 51.0 microm and 59.5 microm, respectively. Car. lactis GeDH was shown to selectively oxidize geraniol, whereas its geometrical isomer, nerol, was inert as a substrate. The high specificity for geraniol suggests that Car. lactis GeDH specializes in the alarm pheromone biosynthesis of Car. lactis. Car. lactis GeDH is composed of 378 amino acids. Structurally, Car. lactis GeDH showed homology with zinc-dependent alcohol dehydrogenases found in mammals and a mosquito (36.6-37.6% identical), and the enzyme was considered to be a member of the medium-chain dehydrogenase/reductase family, in view of the highly conserved sequences of zinc-binding and NAD+-binding sites. Phylogenetic analyses indicate that Car. lactis GeDH could be categorized as a new class, different from other established alcohol dehydrogenases.  相似文献   

19.
L-threonine dehydrogenase (TDH) is an enzyme that catalyzes the oxidation of L-threonine to 2-amino-3-ketobutyrate. We solved the first crystal structure of a medium chain L-threonine dehydrogenase from a hyperthermophilic archaeon, Pyrococcus horikoshii (PhTDH), by the single wavelength anomalous diffraction method using a selenomethionine-substituted enzyme. This recombinant PhTDH is a homo-tetramer in solution. Three monomers of PhTDHs were located in the crystallographic asymmetric unit, however, the crystal structure exhibits a homo-tetramer structure with crystallographic and non-crystallographic 222 symmetry in the cell. Despite the low level of sequence identity to a medium-chain NAD(H)-dependent alcohol dehydrogenase (ADH) and the different substrate specificity, the overall folds of the PhTDH monomer and tetramer are similar to those of the other ADH. Each subunit is composed of two domains: a nicotinamide cofactor (NAD(H))-binding domain and a catalytic domain. The NAD(H)-binding domain contains the alpha/beta Rossmann fold motif, characteristic of the NAD(H)-binding protein. One molecule of PhTDH contains one zinc ion playing a structural role. This metal ion exhibits coordination with four cysteine ligands and some of the ligands are conserved throughout the structural zinc-containing ADHs and TDHs. However, the catalytic zinc ion that is coordinated at the bottom of the cleft in the case of ADH was not observed in the crystal of PhTDH. There is a significant difference in the orientation of the catalytic domain relative to the coenzyme-binding domain that results in a larger interdomain cleft.  相似文献   

20.
There is growing evidence that metabolic enzymes may act as multifunctional proteins performing diverse roles in cellular metabolism. Among these functions are the RNA-binding activities of NAD(+)-dependent dehydrogenases. Previously, we have characterized the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) as an RNA-binding protein with preference to adenine-uracil-rich sequences. In this study, we used GST-GAPDH fusion proteins generated by deletion mutagenesis to search for the RNA binding domain. We established that the N-terminal 43 amino acid residues of GAPDH, which correspond to the first mononucleotide-binding domain of the NAD(+)-binding fold is sufficient to confer RNA-binding. We also provide evidence that this single domain, although it retains most of the RNA-binding activity, loses sequence specificity. Our results suggest a molecular basis for RNA-recognition by NAD(+)-dependent dehydrogenases and (di)nucleotide-binding metabolic enzymes that had been reported to have RNA-binding activity with different specificity. To support this prediction we also identified other members of the family of NAD(+)-dependent dehydrogenases with no previous history of nucleic acid binding as RNA binding proteins in vitro. Based on our findings we propose the addition of the NAD(+)-binding domain to the list of RNA binding domains/motifs.  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