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 62 毫秒
1.
L-Gulonate 3-dehydrogenase (GDH) catalyzes the NAD(+)-linked dehydrogenation of L-gulonate into dehydro-L-gulonate in the uronate cycle. In this study, we isolated the enzyme and its cDNA from rabbit liver, and found that the cDNA is identical to that for rabbit lens lambda-crystallin except for lacking a codon for Glu(309). The same cDNA species, but not the lambda-crystallin cDNA with the codon for Glu(309), was detected in the lens, which showed the highest GDH activity among rabbit tissues. In addition, recombinant human lambda-crystallin that lacks Glu(309) displays enzymatic properties similar to rabbit GDH. These data indicate that GDH is recruited as lambda-crystallin without gene duplication. An outstanding feature of GDH is modulation of its activity by low concentrations of P(i), which decreases the catalytic efficiency in a dose dependent manner. P(i) also protects the enzyme against both thermal and urea denaturation. Kinetic analysis suggests that P(i) binds to both the free enzyme and its NAD(H)-complex in the sequential ordered mechanism. Furthermore, we examined the roles of Asp(36), Ser(124), His(145), Glu(157 )and Asn(196) in the catalytic function of rabbit GDH by site-directed mutagenesis. The D36R mutation leads to a switch in favor of NADP(H) specificity, suggesting an important role of Asp(36) in the coenzyme specificity. The S124A mutation decreases the catalytic efficiency 500-fold, and the H145Q, N196Q and N195D mutations result in inactive enzyme forms, although the E157Q mutation produces no large kinetic alteration. Thus, Ser(124), His(145) and Asn(196) may be critical for the catalytic function of GDH.  相似文献   

2.
Clostridial glutamate dehydrogenase mutants, designed to accommodate the 2'-phosphate of disfavoured NADPH, showed the expected large specificity shifts with NAD(P)H. Puzzlingly, similar assays with oxidized cofactors initially revealed little improvement with NADP(+) , although rates with NAD(+) were markedly diminished. This article reveals that the enzyme's discrimination in favour of NAD(+) and against NADP(+) had been greatly underestimated and has indeed been abated by a factor of >?16,000 by the mutagenesis. Initially, stopped-flow studies of the wild-type enzyme showed a burst increase of A(340) with NADP(+) but not NAD(+), with amplitude depending on the concentration of the coenzyme, rather than enzyme. Amplitude also varied with the commercial source of the NADP(+). FPLC, HPLC and mass spectrometry identified NAD(+) contamination ranging from 0.04 to 0.37% in different commercial samples. It is now clear that apparent rates of NADP(+) utilization mainly reflected the reduction of contaminating NAD(+), creating an entirely false view of the initial coenzyme specificity and also of the effects of mutagenesis. Purification of the NADP(+) eliminated the burst. With freshly purified NADP(+), the NAD(+) : NADP(+) activity ratio under standard conditions, previously estimated as 300 : 1, is 11,000. The catalytic efficiency ratio is even higher at 80,000. Retested with pure cofactor, mutants showed marked specificity shifts in the expected direction, for example, 16 200 fold change in catalytic efficiency ratio for the mutant F238S/P262S, confirming that the key structural determinants of specificity have been successfully identified. Of wider significance, these results underline that, without purification, even the best commercial coenzyme preparations are inadequate for such studies.  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.
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.
Sharkey MA  Gori A  Capone M  Engel PC 《The FEBS journal》2012,279(17):3003-3009
Active-site mutants of glutamate dehydrogenase from Clostridium?symbiosum have been designed and constructed and the effects on coenzyme preference evaluated by detailed kinetic measurements. The triple mutant F238S/P262S/D263K shows complete reversal in coenzyme selectivity from NAD(H) to NADP(H) with retention of high levels of catalytic activity for the new coenzyme. For oxidized coenzymes, k(cat) /K(m) ratios of the wild-type and triple mutant enzyme indicate a shift in preference of approximately 1.6?×?10(7) -fold, from ~?80?000-fold in favour of NAD(+) to ~?200-fold in favour of NADP(+) . For reduced coenzymes the corresponding figure is 1.7?×?10(4) -fold, from ~?1000-fold in favour of NADH to ~?17-fold in favour of NADPH. A fourth mutation (N290G), previously identified as having a potential bearing on coenzyme specificity, did not engender any further shift in preference when incorporated into the triple mutant, despite having a significant effect when expressed as a single mutant.  相似文献   

6.
Ferredoxin-NADP(+) reductase catalyses NADP(+) reduction, being specific for NADP(+)/H. To understand coenzyme specificity determinants and coenzyme specificity reversion, mutations at the NADP(+)/H pyrophosphate binding and of the C-terminal regions have been simultaneously introduced in Anabaena FNR. The T155G/A160T/L263P/Y303S mutant was produced. The mutated enzyme presents similar k(cat) values for NADPH and NADH, around 2.5 times slower than that reported for WT FNR with NADPH. Its K(m) value for NADH decreased 20-fold with regard to WT FNR, whereas the K(m) for NADPH remains similar. The combined effect is a much higher catalytic efficiency for NAD(+)/H, with a minor decrease of that for NADP(+)/H. In the mutated enzyme, the specificity for NADPH versus NADH has been decreased from 67,500 times to only 12 times, being unable to discriminate between both coenzymes. Additionally, giving the role stated for the C-terminal Tyr in FNR, its role in the energetics of the FAD binding has been analysed.  相似文献   

7.
The primary structure of the aldose xylose reductase from Candida tenuis (CtAR) is shown to be 39% identical to that of human aldose reductase (hAR). The catalytic tetrad of hAR is completely conserved in CtAR (Tyr51, Lys80, Asp46, His113). The amino acid residues involved in binding of NADPH by hAR (D.K. Wilson, et al., Science 257 (1992) 81-84) are 64% identical in CtAR. Like hAR the yeast enzyme is specific for transferring the 4-pro-R hydrogen of the coenzyme. These properties suggest that CtAR is a member of the aldo/keto reductase superfamily. Unlike hAR the enzyme from C. tenuis has a dual coenzyme specificity and shows similar specificity constants for NADPH and NADH. It binds NADP(+) approximately 250 times less tightly than hAR. Typical turnover numbers for aldehyde reduction by CtAR (15-20 s(-1)) are up to 100-fold higher than corresponding values for hAR, probably reflecting an overall faster dissociation of NAD(P)(+) in the reaction catalyzed by the yeast enzyme.  相似文献   

8.
Reduction of 17-ketosteroids is a biocatalytic process of economic significance for the production of steroid drugs. This reaction can be catalyzed by different microbial 17beta-hydroxysteroid dehydrogenases (17beta-HSD), like the 17beta-HSD activity of Saccharomyces cerevisiae, Pichia faranosa and Mycobacterium sp., and by purified 3beta,17beta-HSD from Pseudomonas testosteroni. In addition to the bacterial 3beta,17beta-HSD the 17beta-HSD of the filamentous fungus Cochliobolus lunatus is the only microbial 17beta-HSD that has been expressed as a recombinant protein and fully characterized. On the basis of its modeled 3D structure, we selected several positions for the replacement of amino acids by site-directed mutagenesis to change substrate specificity, alter coenzyme requirements, and improve overall catalytic activity. Replacement of Val161 and Tyr212 in the substrate-binding region by Gly and Ala, respectively, increased the initial rates for the conversion of androstenedione to testosterone. Replacement of Tyr49 within the coenzyme binding site by Asp changed the coenzyme specificity of the enzyme. This latter mutant can convert the steroids not only in the presence of NADP(+) and NADPH, but also in the presence of NADH and NAD(+). The replacement of His164, located in the non-flexible part of the 'lid' covering the active center resulted in a conformation of the enzyme that possessed a higher catalytic activity.  相似文献   

9.
Drosophila alcohol dehydrogenase (ADH), an NAD(+)-dependent dehydrogenase, shares little sequence similarity with horse liver ADH. However, these two enzymes do have substantial similarity in their secondary structure at the NAD(+)-binding domain [Benyajati, C., Place, A. P., Powers, D. A. & Sofer, W. (1981) Proc. Natl Acad. Sci. USA 78, 2717-2721]. Asp38, a conserved residue between Drosophila and horse liver ADH, appears to interact with the hydroxyl groups of the ribose moiety in the AMP portion of NAD+. A secondary-structure comparison between the nucleotide-binding domain of NAD(+)-dependent enzymes and that of NADP(+)-dependent enzymes also suggests that Asp38 could play an important role in cofactor specificity. Mutating Asp38 of Drosophila ADH into Asn38 decreases Km(app)NADP 62-fold and increases kcat/Km(app)NADP 590-fold at pH 9.8, when compared with wild-type ADH. These results suggest that Asp38 is in the NAD(+)-binding domain and its substituent, Asn38, allows Drosophila ADH to use both NAD+ and NADP+ as its cofactor. The observations from the experiments of thermal denaturation and kinetic measurement with pH also confirm that the repulsion between the negative charges of Asp38 and 2'-phosphate of NADP+ is the major energy barrier for NADP+ to serve as a cofactor for Drosophila ADH.  相似文献   

10.
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   

11.
Previously we found that replacement of seven amino acid residues in a loop region markedly shifted the coenzyme specificity of malate dehydrogenase from NAD(H) toward NADP(H). In the present study, we replaced the seven amino acid residues in the corresponding region of an NAD(H)-dependent lactate dehydrogenase with those of NADP(H)-dependent malate dehydrogenase, and examined the coenzyme specificity of the resulting mutant enzyme. Coenzyme specificity was significantly shifted by 399-fold toward NADPH when k cat/Km(coenzyme) was used as the measure of coenzyme specificity. The effect of the replacements on coenzyme specificity is discussed based on in silico simulation of the three-dimensional structure of the lactate dehydrogenase mutant.  相似文献   

12.
The role of Asp-328 and Ile-329 as a cofactor discrimination site of the NAD-dependent isocitrate dehydrognase (NAD-IDH) from Pyrococcus furiosus has been verified by replacing these residues with Lys and Tyr, respectively, which are the corresponding residues in NADP-IDH from Escherichia coli. The Asp-328-Lys mutant showed dual coenzyme specificity, whereas introduction of the double mutation, Asp-328-Lys/Ile-329-Tyr shifted the cofactor preference from NAD to NADP. NADP-dependent P. furiosus IDH retained thermostability and thermoactivity compared with NAD-IDH.  相似文献   

13.
L A LeBrun  B V Plapp 《Biochemistry》1999,38(38):12387-12393
The rate of association of NAD(+) with wild-type horse liver alcohol dehydrogenase (ADH) is maximal at pH values between pK values of about 7 and 9, and the rate of NADH association is maximal at a pH below a pK of 9. The catalytic zinc-bound water, His-51 (which interacts with the 2'- and 3'-hydroxyl groups of the nicotinamide ribose of the coenzyme in the proton relay system), and Lys-228 (which interacts with the adenosine 3'-hydroxyl group and the pyrophosphate of the coenzyme) may be responsible for the observed pK values. In this study, the Lys228Arg, His51Gln, and Lys228Arg/His51Gln (to isolate the effect of the catalytic zinc-bound water) mutations were used to test the roles of the residues in coenzyme binding. The steady state kinetic constants at pH 8 for the His51Gln enzyme are similar to those for wild-type ADH. The Lys228Arg and Lys228Arg/His51Gln substitutions decrease the affinity for the coenzymes up to 16-fold, probably due to altered interactions with the arginine at position 228. As determined by transient kinetics, the rate constant for association of NAD(+) with the mutated enzymes no longer decreases at high pH. The pH profile for the Lys228Arg enzyme retains the pK value near 7. The His51Gln and Lys228Arg/His51Gln substitutions significantly decrease the rate constants for NAD(+) association, and the pH dependencies show that these enzymes bind NAD(+) most rapidly at a pH above pK values of 8. 0 and 9.0, respectively. It appears that the pK of 7 in the wild-type enzyme is shifted up by the H51Q substitutions, and the resulting pH dependence is due to the deprotonation of the catalytic zinc-bound water. Kinetic simulations suggest that isomerization of the enzyme-NAD(+) complex is substantially altered by the mutations. In contrast, the pH dependencies for NADH association with His51Gln, Lys228Arg, and Lys228Arg/His51Gln enzymes were the same as for wild-type ADH, suggesting that the binding of NAD(+) and the binding of NADH are controlled differently.  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.
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   

16.
Previously we found that replacement of seven amino acid residues in a loop region markedly shifted the coenzyme specificity of malate dehydrogenase from NAD(H) toward NADP(H). In the present study, we replaced the seven amino acid residues in the corresponding region of an NAD(H)-dependent lactate dehydrogenase with those of NADP(H)-dependent malate dehydrogenase, and examined the coenzyme specificity of the resulting mutant enzyme. Coenzyme specificity was significantly shifted by 399-fold toward NADPH when k cat?K m coenzyme was used as the measure of coenzyme specificity. The effect of the replacements on coenzyme specificity is discussed based on in silico simulation of the three-dimensional structure of the lactate dehydrogenase mutant.  相似文献   

17.
Influences on coenzyme preference are explored. Lysine 137 (192 in class 1/2 ALDH) lies close to the adenine ribose, directly interacting with the adenine ribose in NAD-specific ALDHs and the 2'-phosphate of NADP in NADP-specific ALDHs. Lys-137 in class 3 ALDH interacts with the adenine ribose indirectly through an intervening water molecule. However, this residue is present in all ALDHs and, as a result, is unlikely to directly influence coenzyme specificity. Glutamate 140 (195) coordinates the 2'- and 3'-hydroxyls of the adenine ribose of NAD in the class 3 tertiary structure. Thus, it appeared that this residue would influence coenzyme specificity. Mutation to aspartate, asparagine, glutamine or threonine shifts the coenzyme specificity towards NADP, but did not completely change the specificity. Still, the mutants show the 2'-phosphate of NADP is repelled by Glu-140 (195). Although Glu-140 (195) has a major influence on coenzyme specificity, it is not the only influence since class 3 ALDHs, can use both coenzymes, and class 2 ALDHs, which are NAD-specific, have a glutamate at this position. One explanation may be that the larger space between Lys-137 (192) and the adenine ribose hydroxyls in the class 3 ALDH:NAD binary structure may provide space to accommodate the 2'-phosphate of NADP. Also, a structural shift upon binding NADP may also occur in class 3 ALDHs to help accommodate the 2'-phosphate of NADP.  相似文献   

18.
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   

19.
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   

20.
Influences on coenzyme preference are explored. Lysine 137 (192 in class 1/2 ALDH) lies close to the adenine ribose, directly interacting with the adenine ribose in NAD-specific ALDHs and the 2′-phosphate of NADP in NADP-specific ALDHs. Lys-137 in class 3 ALDH interacts with the adenine ribose indirectly through an intervening water molecule. However, this residue is present in all ALDHs and, as a result, is unlikely to directly influence coenzyme specificity. Glutamate 140 (195) coordinates the 2′- and 3′-hydroxyls of the adenine ribose of NAD in the class 3 tertiary structure. Thus, it appeared that this residue would influence coenzyme specificity. Mutation to aspartate, asparagine, glutamine or threonine shifts the coenzyme specificity towards NADP, but did not completely change the specificity. Still, the mutants show the 2′-phosphate of NADP is repelled by Glu-140 (195). Although Glu-140 (195) has a major influence on coenzyme specificity, it is not the only influence since class 3 ALDHs, can use both coenzymes, and class 2 ALDHs, which are NAD-specific, have a glutamate at this position. One explanation may be that the larger space between Lys-137 (192) and the adenine ribose hydroxyls in the class 3 ALDH:NAD binary structure may provide space to accommodate the 2′-phosphate of NADP. Also, a structural shift upon binding NADP may also occur in class 3 ALDHs to help accommodate the 2′-phosphate of NADP.  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