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

2.
Chin JK  Klinman JP 《Biochemistry》2000,39(6):1278-1284
A tunneling contribution to hydride transfer has been demonstrated previously in the oxidation of benzyl alcohol catalyzed by an active-site mutant (F93W) of horse liver alcohol dehydrogenase (LADH) [Bahnson, B. J., et al. (1993) Biochemistry 32, 5503-5507]. Mutation of a residue that lies directly behind the nicotinamide ring of the bound cofactor has further shown that side-chain bulk can contribute to catalytic efficiency and tunneling in a correlated fashion [Bahnson, B. J., et al. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 12797-12802]. Second site mutations of F93W have now been made at positions more remote from the active site. In particular, we have focused on an isoleucine residue that interacts with the adenine moiety of the NAD(+) cofactor, 20 A from the nicotinamide ring. Replacement of this remote residue with glycine (F93W:I224G), alanine (F93W:I224A), valine (F93W:I224V), and leucine (F93W:I224L) is concluded to destabilize the binding of NAD(+). All double mutants exhibited a K(M) for NAD(+) that is 2-25 times higher than that for the F93W enzyme. However, neither the catalytic efficiency for turnover of benzyl alcohol [k(cat)/K(M(benzyl alcohol))] nor the relationship between the secondary k(H)/k(T) and k(D)/k(T) isotope effects for benzyl alcohol oxidation was significantly affected. The lack of differences observed in the isotope effects indicates that these mutations have little effect on the extent of hydrogen tunneling in the reaction. The complete removal of the side chain at position 224 in the F93W:I224G enzyme resulted in a less than 5% decrease in the ratio of the secondary isotope effects, maintaining the ratio above the semiclassical limit for the indication of tunneling in the reaction. By contrast, K(i) for NAD(+) increased 60-fold for this mutant. The results obtained with F93W:I224G are consistent with remote interactions that affect the association and binding of cofactor in a reactive conformation. However, once this conformation is achieved, hydride transfer and its tunneling component proceed as with the single F93W mutant enzyme, uninfluenced by the remote mutation. Replacement of other side chains, with alpha-carbon positions from about 8 to over 20 A from the C4 position of the nicotinamide ring, demonstrated a similar insensitivity of k(cat)/K(M(benzyl alcohol)) to protein modification. Comparison to earlier studies with active-site mutants of LADH implicates a role for proximal, but not distal, side chains in the modulation of hydrogen tunneling for this enzyme.  相似文献   

3.
The roles of particular amino acids in substrate and coenzyme binding and catalysis of glucose-6-phosphate dehydrogenase of Leuconostoc mesenteroides have been investigated by site-directed mutagenesis, kinetic analysis, and determination of binding constants. The enzyme from this species has functional dual NADP(+)/NAD(+) specificity. Previous investigations in our laboratories determined the three-dimensional structure. Kinetic studies showed an ordered mechanism for the NADP-linked reaction while the NAD-linked reaction is random. His-240 was identified as the catalytic base, and Arg-46 was identified as important for NADP(+) but not NAD(+) binding. Mutations have been selected on the basis of the three-dimensional structure. Kinetic studies of 14 mutant enzymes are reported and kinetic mechanisms are reported for 5 mutant enzymes. Fourteen substrate or coenzyme dissociation constants have been measured for 11 mutant enzymes. Roles of particular residues are inferred from k(cat), K(m), k(cat)/K(m), K(d), and changes in kinetic mechanism. Results for enzymes K182R, K182Q, K343R, and K343Q establish Lys-182 and Lys-343 as important in binding substrate both to free enzyme and during catalysis. Studies of mutant enzymes Y415F and Y179F showed no significant contribution for Tyr-415 to substrate binding and only a small contribution for Tyr-179. Changes in kinetics for T14A, Q47E, and R46A enzymes implicate these residues, to differing extents, in coenzyme binding and discrimination between NADP(+) and NAD(+). By the same measure, Lys-343 is also involved in defining coenzyme specificity. Decrease in k(cat) and k(cat)/K(m) for the D374Q mutant enzyme defines the way Asp-374, unique to L. mesenteroides G6PD, modulates stabilization of the enzyme during catalysis by its interaction with Lys-182. The greatly reduced k(cat) values of enzymes P149V and P149G indicate the importance of the cis conformation of Pro-149 in accessing the correct transition state.  相似文献   

4.
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   

5.
Dihydrofolate reductase (EC 1.5.1.3) is a key enzyme in the folate biosynthetic pathway. Information regarding key residues in the dihydrofolate-binding site of Mycobacterium avium dihydrofolate reductase is lacking. On the basis of previous information, Asp31 and Leu32 were selected as residues that are potentially important in interactions with dihydrofolate and antifolates (e.g. trimethoprim), respectively. Asp31 and Leu32 were modified by site-directed mutagenesis, giving the mutants D31A, D31E, D31Q, D31N and D31L, and L32A, L32F and L32D. Mutated proteins were expressed in Escherichia coli BL21(DE3)pLysS and purified using His-Bind resin; functionality was assessed in comparison with the recombinant wild type by a standard enzyme assay, and growth complementation and kinetic parameters were evaluated. All Asp31 substitutions affected enzyme function; D31E, D31Q and D31N reduced activity by 80-90%, and D31A and D31L by > 90%. All D31 mutants had modified kinetics, ranging from three-fold (D31N) to 283-fold (D31L) increases in K(m) for dihydrofolate, and 12-fold (D31N) to 223 077-fold (D31L) decreases in k(cat)/K(m). Of the Leu32 substitutions, only L32D caused reduced enzyme activity (67%) and kinetic differences from the wild type (seven-fold increase in K(m); 21-fold decrease in k(cat)/K(m)). Only minor variations in the K(m) for NADPH were observed for all substitutions. Whereas the L32F mutant retained similar trimethoprim affinity as the wild type, the L32A mutation resulted in a 12-fold decrease in affinity and the L32D mutation resulted in a seven-fold increase in affinity for trimethoprim. These findings support the hypotheses that Asp31 plays a functional role in binding of the substrate and Leu32 plays a functional role in binding of trimethoprim.  相似文献   

6.
Elrod P  Zhang J  Yang X  Yin D  Hu Y  Borchardt RT  Schowen RL 《Biochemistry》2002,41(25):8134-8142
Residues glutamate 156 (E156), aspartate 190 (D190), asparagine 181 (N181), lysine 186 (K186), and asparagine 191 (N191) in the active site of S-adenosylhomocysteine (AdoHcy) hydrolase have been mutated to alanine (A). AdoHcy hydrolase achieves catalysis of AdoHcy hydrolysis to adenosine (Ado) and homocysteine (Hcy) by means of a redox partial reaction (3'-oxidation of AdoHcy at the beginning and 3'-reduction of Ado at the end of the catalytic cycle) spanning an elimination/addition partial reaction (elimination of Hcy from the oxidized substrate and addition of water to generate the oxidized product), with the enzyme in an open NAD(+) form in the ligand-free state and in a closed NADH form during the elimination/addition partial reaction. Mutation K186A reduces the rate of a model enzymatic reaction for the redox partial reaction by a factor of 280000 and the rate of a model reaction for the elimination/addition partial reaction by a factor of 24000, consistent with a primary catalytic role in both partial reactions as a proton donor/acceptor at the 3'-OH/3'-keto center. Secondary roles for N181 and N191 in localizing the flexible side chain of K186 in a catalytically effective position are supported by rate reduction factors for N181A of 2500 (redox) and 240 (elimination/addition) and for N191A of 730 (redox) and 340 (elimination/addition). A role of D190 in orienting the substrate for effective transition-state stabilization is consistent with rate reduction factors of 1300 (redox) and 30 (elimination/addition) for D190A. Residue E156 may act to maintain K186 in the desired protonation state: rate deduction factors are 1100 (redox) and 70 (elimination/addition). The mutational increases in free energy barriers for k(cat)/K(M) are described by a linear combination of the effects for the partial reactions with the coefficients equal to the fractional degree that each partial reaction determines the rate for k(cat)/K(M). A similar linear equation for k(cat) overestimates the barrier increase by a uniform 5 kJ/mol, probably reflecting reactant-state stabilization by the wild-type enzyme that is abolished by the mutations.  相似文献   

7.
The deacetoxycephalosporin C synthase (DAOCS) from Streptomyces clavuligerus was engineered with the aim of enhancing the conversion of penicillin G into phenylacetyl-7-aminodeacetoxycephalosporanic acid, a precursor of 7-aminodeacetoxycephalosporanic acid, for industrial application. A single round of random mutagenesis followed by the screening of 5,500 clones identified three mutants, G79E, V275I, and C281Y, that showed a two- to sixfold increase in the k(cat)/K(m) ratio compared to the wild-type enzyme. Site-directed mutagenesis to modify residues surrounding the substrate resulted in three mutants, N304K, I305L, and I305M, with 6- to 14-fold-increased k(cat)/K(m) values. When mutants containing all possible combinations of these six sites were generated to optimize the ring expansion activity for penicillin G, the double mutant, YS67 (V275I, I305M), showed a significant 32-fold increase in the k(cat)/K(m) ratio and a 5-fold increase in relative activity for penicillin G, while the triple mutant, YS81 (V275I, C281Y, I305M), showed an even greater 13-fold increase in relative activity toward penicillin G. Our results demonstrate that this is a robust approach to the modification of DAOCS for an optimized DAOCS-penicillin G reaction.  相似文献   

8.
ALDH5 (aka succinic semialdehyde dehydrogenase) is a NAD(+)-dependent aldehyde dehydrogenase crucial for the proper removal of the GABA metabolite succinic semialdehyde (SSA). All known ALDH5 family members contain the conserved amino acid sequence "MITRK". Our studies of rat ALDH5A indicate that residue R166 in this sequence may play a role in the substrate specificity of ALDH5A for the gamma-carboxylated succinic semialdehyde versus other aliphatic and aromatic aldehydes including acetaldehyde and benzaldehyde. We tested the hypothesis that the R166 residue regulates aldehyde specificity by utilizing rat ALDH5A wild-type (R166wt) and R166K, R166H, R166A, and R166E mutants. The V(MAX) using SSA fell whereas the K(M) for SSA increased for all mutants analyzed yielding k(cat)/K(M) (s(-1)/microM) ratios of 52.3 (R166wt), 5.5 (R166K), 0.01 (R166H), 0.008 (R166E), and 0.004 (R166A). Utilization of acetaldehyde by the R166H mutant was similar to R166wt with k(cat)/K(M)'s of 0.003 and 0.002, respectively. Almost no activity towards acetaldehyde was noted for the R166E and R166A mutants. Unexpectedly, the K(M) for NAD(+) changed: 21 microM (R166wt), 81 microM (R166K), 63 microM (R166H), 35 microM (R166E) and 44 microM (R166A). As release of NADH can be a rate-limiting step for ALDH activity, NADH binding was evaluated for R166wt and R166H enzymes. The K(D) of NADH for R166H (0.9 microM) was 11-fold less than that of ALDH5A wt (10.3 microM) and possibly explains the increase in the K(M) for NAD(+). Furthermore, data using R166K and R166H mutants demonstrate that inhibition of enzyme activity by low pH is regulated in part by the R166 residue. Our data indicate that the R166 residue of ALDH5A regulates multiple enzymatic functions.  相似文献   

9.
Based on recent directed evolution of P450 2B1, six P450 2B11 mutants at three positions were created in an N-terminal modified construct termed P450 2B11dH and characterized for enzyme catalysis using five substrates. Mutant I209A demonstrated a 3.2-fold enhanced k(cat)/K(m) for 7-ethoxy-4-trifluoromethylcourmarin O-deethylation, largely due to a dramatic decrease in K(m) (0.72 microM vs. 18 microM). I209A also demonstrated enhanced selectivity for testosterone 16beta-hydroxylation over 16alpha-hydroxylation. In contrast, V183L showed a 4-fold increased k(cat) for 7-benzyloxyresorufin debenzylation and a 4.7-fold increased k(cat)/K(m) for testosterone 16alpha-hydroxylation. V183L also displayed a 1.7-fold higher k(cat)/K(m) than P450 2B11dH with the anti-cancer prodrugs cyclophosphamide and ifosfamide, resulting from a approximately 4-fold decrease in K(m). Introduction of the V183L mutation into full-length P450 2B11 did not enhance the k(cat)/K(m). Overall, the re-engineered P450 2B11dH enzymes exhibited enhanced catalytic efficiency with several substrates including the anti-cancer prodrugs.  相似文献   

10.
The six mutations, referred to as the Hex mutations, that together have been shown to convert Escherichia coli aspartate aminotransferase (AATase) specificity to be substantially like that of E. coli tyrosine aminotransferase (TATase) are dissected into two groups, (T109S/N297S) and (V39L/K41Y/T47I/N69L). The letters on the left and right of the numbers designate AATase and TATase residues, respectively. The T109S/N297S pair has been investigated previously. The latter group, the "Grease" set, is now placed in the AATase framework, and the retroGrease set (L39V/Y41K/I47T/L69N) is substituted into TATase. The Grease mutations in the AATase framework were found primarily to lower K(M)s for both aromatic and dicarboxylic substrates. In contrast, retroGrease TATase exhibits lowered k(cat)s for both substrates. The six retroHex mutations, combining retroGrease and S109T/S297N, were found to invert the substrate specificity of TATase, creating an enzyme with a nearly ninefold preference (k(cat)/K(M)) for aspartate over phenylalanine. The retroHex mutations perturb the electrostatic environment of the pyridoxal phosphate cofactor, as evidenced by a spectrophotometric titration of the internal aldimine, which uniquely shows two pK(a)s, 6.1 and 9.1. RetroHex was also found to have impaired dimer stability, with a K(D) for dimer dissociation of 350 nM compared with the wild type K(D) of 4 nM. Context dependence and additivity analyses demonstrate the importance of interactions of the Grease residues with the surrounding protein framework in both the AATase and TATase contexts, and with residues 109 and 297 in particular. Context dependence and cooperativity are particularly evident in the effects of mutations on k(cat)/K(M)(Asp). Effects on k(cat)/K(M)(Phe) are more nearly additive and context independent.  相似文献   

11.
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   

12.
Phosphite dehydrogenase (PTDH) catalyzes the NAD-dependent oxidation of phosphite to phosphate, a reaction that is 15 kcal/mol exergonic. The enzyme belongs to the family of D-hydroxy acid dehydrogenases. Five other family members that were analyzed do not catalyze the oxidation of phosphite, ruling out the possibility that this is a ubiquitous activity of these proteins. PTDH does not accept any alternative substrates such as thiophosphite, hydrated aldehydes, and methylphosphinate, and potential small nucleophiles such as hydroxylamine, fluoride, methanol, and trifluoromethanol do not compete with water in the displacement of the hydride from phosphite. The pH dependence of k(cat)/K(m,phosphite) is bell-shaped with a pK(a) of 6.8 for the acidic limb and a pK(a) of 7.8 for the basic limb. The pK(a) of 6.8 is assigned to the second deprotonation of phosphite. However, whether the dianionic form of phosphite is the true substrate is not clear since a reverse protonation mechanism is also consistent with the available data. Unlike k(cat)/K(m,phosphite), k(cat) and k(cat)/K(m,NAD) are pH-independent. Sulfite is a strong inhibitor of PTDH that is competitive with respect to phosphite and uncompetitive with respect to NAD(+). Incubation of the enzyme with NAD(+) and low concentrations of sulfite results in a covalent adduct between NAD(+) and sulfite in the active site of the enzyme that binds very tightly. Fluorescent titration studies provided the apparent dissociation constants for NAD(+), NADH, sulfite, and the sulfite-NAD(+) adduct. Substrate isotope effect studies with deuterium-labeled phosphite resulted in small normal isotope effects (1.4-2.1) on both k(cat) and k(cat)/K(m,phosphite) at pH 7.25 and 8.0. Solvent isotope effects (SIEs) on k(cat) are similar in size; however, the SIE of k(cat)/K(m,phosphite) at pH 7.25 is significantly larger (4.4), whereas at pH 8.0, it is the inverse (0.6). The pH-rate profile of k(cat)/K(m,phosphite), which predicts that the observed SIEs will have a significant thermodynamic origin, can account for these effects.  相似文献   

13.
The meta-cleavage product hydrolase from Pseudomonas fluorescens IP01 (CumD) hydrolyzes 2-hydroxy-6-oxo-7-methylocta-2,4-dienoate (6-isopropyl HODA) in the cumene (isopropylbenzene) degradation pathway. To modulate the substrate specificity and catalytic efficiency of CumD toward substrates derived from monocyclic aromatic compounds, we constructed the CumD mutants, A129V, I199V, and V227I, as well as four types of double and triple mutants. Toward substrates with smaller side chains (e.g. 2-hydroxy-6-oxohepta-2,4-dienoate; 6-ethyl-HODA), the k(cat)/K(m) values of the single mutants were 4.2-11 fold higher than that of the wild type enzyme and 1.8-4.7 fold higher than that of the meta-cleavage product hydrolase from Pseudomonas putida F1 (TodF). The A129V mutant showed the highest k(cat)/K(m) value for 2-hydroxy-6-oxohepta-2,4-dienoate (6-ethyl-HODA). The crystal structure of the A129V mutant was determined at 1.65 A resolution, enabling location of the Ogamma atom of the Ser103 side chain. A chloride ion was bound to the oxyanion hole of the active site, and mutant enzymes at the residues forming this site were also examined. The k(cat) values of Ser34 mutants were decreased 2.9-65 fold, suggesting that the side chain of Ser34 supports catalysis by stabilizing the anionic oxygen of the proposed intermediate state (gem-diolate). This is the first crystal structure determination of CumD in an active form, with the Ser103 residue, one of the catalytically essential "triad", being intact.  相似文献   

14.
The factors that govern the substrate reactivity and stereoselectivity of phosphotriesterase (PTE) toward organophosphotriesters containing various combinations of methyl, ethyl, isopropyl, and phenyl substituents at the phosphorus center were determined by systematic alterations in the dimensions of the active site. The wild type PTE prefers the S(P)-enantiomers over the corresponding R(P)-enantiomers by factors ranging from 10 to 90. Enlargement of the small subsite of PTE with the substitution of glycine and alanine residues for Ile-106, Phe-132, and/or Ser-308 resulted in significant improvements in k(cat)/K(a) for the R(P)-enantiomers of up to 2700-fold but had little effect on k(cat)/K(a) for the corresponding S(P)-enantiomers. The kinetic preferences for the S(P)-enantiomers were thus relaxed without sacrificing the inherent catalytic activity of the wild type enzyme. A reduction in the size of the large subsite with the mutant H257Y resulted in a reduction in k(cat)/K(a) for the S(P)-enantiomers, while the values of k(cat)/K(a) for the R(P)-enantiomers were essentially unchanged. The initial stereoselectivity observed with the wild type enzyme toward the chiral substrate library was significantly reduced with the H257Y mutant. Simultaneous alternations in the sizes of the large and small subsites resulted in the complete reversal of the chiral specificity. With this series of mutants, the R(P)-enantiomers were preferred as substrates over the corresponding S(P)-enantiomers by up to 500-fold. These results have demonstrated that the stereochemical determinants for substrate hydrolysis by PTE can be systematically altered through a rational reconstruction of the dimensions of the active site.  相似文献   

15.
Based on the X-ray crystal structures of 4-(4-chlorophenyl)imidazole (4-CPI)- and bifonazole (BIF)-bound P450 2B4, eight active site mutants at six positions were created in an N-terminal modified construct termed 2B4dH and characterized for enzyme inhibition and catalysis. I363A showed a >4-fold decrease in differential inhibition by BIF and 4-CPI (IC(50,BIF)/IC(50,4-CPI)). F296A, T302A, I363A, V367A, and V477A showed a 2-fold decreased k(cat) for 7-ethoxy-4-trifluoromethylcoumarin O-deethylation, whereas V367A and V477F showed an altered K(m). T302A, V367L, and V477A showed >4-fold decrease in total testosterone hydroxylation, whereas I363A, V367A, and V477F showed altered stereo- and regioselectivity. Interestingly, I363A showed a 150-fold enhanced k(cat)/K(m) with testosterone, and yielded a new metabolite. Furthermore, testosterone docking into three-dimensional models of selected mutants based on the 4-CPI-bound structure suggested a re-positioning of residues 363 and 477 to yield products. In conclusion, our results suggest that the 4-CPI-bound 2B4dH/H226Y crystal structure is an appropriate model for predicting enzyme catalysis.  相似文献   

16.
Mouse thymidylate synthase R209K (a mutation corresponding to R218K in Lactobacillus casei), overexpressed in thymidylate synthase-deficient Escherichia coli strain, was poorly soluble and with only feeble enzyme activity. The mutated protein, incubated with FdUMP and N(5,10)-methylenetetrahydrofolate, did not form a complex stable under conditions of SDS/polyacrylamide gel electrophoresis. The reaction catalyzed by the R209K enzyme (studied in a crude extract), compared to that catalyzed by purified wild-type recombinant mouse thymidylate synthase, showed the K(m) value for dUMP 571-fold higher and V(max) value over 50-fold (assuming that the mutated enzyme constituted 20% of total crude extract protein) lower. Thus the ratios k(cat, R209K)/k(cat, 'wild') and (k(cat, R209K)/K(m, R209K)(dUMP))/( k(cat, 'wild')/K(m, 'wild')(dUMP)) were 0.019 and 0.000032, respectively, documenting that mouse thymidylate synthase R209, similar to the corresponding L. casei R218, is essential for both dUMP binding and enzyme reaction.  相似文献   

17.
Klimacek M  Nidetzky B 《Biochemistry》2002,41(31):10158-10165
Mannitol dehydrogenases (MDH) are a family of Zn(2+)-independent long-chain alcohol dehydrogenases that catalyze the regiospecific NAD(+)-dependent oxidation of a secondary alcohol group in polyol substrates. pH and primary deuterium kinetic isotope effects on kinetic parameters for reaction of recombinant MDH from Pseudomonas fluorescens with D-mannitol have been measured in H(2)O and D(2)O at 25 degrees C and used to determine the relative timing of C-H and O-H bond cleavage steps during alcohol conversion. The enzymatic rates decreased at low pH; apparent pK values for log(k(cat)/K(mannitol)) and log k(cat) were 9.2 and 7.7 in H(2)O, respectively, and both were shifted by +0.4 pH units in D(2)O. Proton inventory plots for k(cat) and k(cat)/K(mannitol) were determined at pL 10.0 using protio or deuterio alcohol and were linear at the 95% confidence level. They revealed the independence of primary deuterium isotope effects on the atom fraction of deuterium in a mixed H(2)O-D(2)O solvent and yielded single-site transition-state fractionation factors of 0.43 +/- 0.05 and 0.47 +/- 0.01 for k(cat)/K(mannitol) and k(cat), respectively. (D)(k(cat)/K(mannitol)) was constant (1.80 +/- 0.20) in the pH range 6.0-9.5 and decreased at high pH to a limiting value of approximately 1. Measurement of (D)(k(cat)/K(fructose)) at pH 10.0 and 10.5 using NADH deuterium-labeled in the 4-pro-S position gave a value of 0.83, the equilibrium isotope effect on carbonyl group reduction. A mechanism of D-mannitol oxidation by MDH is supported by the data in which the partly rate-limiting transition state of hydride transfer is stabilized by a single solvation catalytic proton bridge. The chemical reaction involves a pH-dependent internal equilibrium which takes place prior to C-H bond cleavage and in which proton transfer from the reactive OH to the enzyme catalytic base may occur. Loss of a proton from the enzyme at high pH irreversibly locks the ternary complex with either alcohol or alkoxide bound in a conformation committed of undergoing NAD(+) reduction at a rate about 2.3-fold slower than the corresponding reaction rate of the protonated complex. Transient kinetic studies for D-mannitol oxidation at pH(D) 10.0 showed that the solvent isotope effect on steady-state turnover originates from a net rate constant of NADH release that is approximately 85% rate-limiting for k(cat) and 2-fold smaller in D(2)O than in H(2)O.  相似文献   

18.
Cytochrome b(5) reductase (cb5r) catalyzes the transfer of reducing equivalents from NADH to cytochrome b(5). Utilizing an efficient heterologous expression system that produces a histidine-tagged form of the hydrophilic, diaphorase domain of the enzyme, site-directed mutagenesis has been used to generate cb5r mutants with substitutions at position 91 in the primary sequence. Arginine 91 is an important residue in binding the FAD prosthetic group and part of a conserved "RxY(T)(S)xx(S)(N)" sequence motif that is omnipresent in the "ferredoxin:NADP(+) reductase" family of flavoproteins. Arginine 91 was replaced with K, L, A, P, D, Q, and H residues, respectively, and all the mutant proteins purified to homogeneity. Individual mutants were expressed with variable efficiency and all exhibited molecular masses of approximately 32 kDa. With the exception of R91H, all the mutants retained visible absorption spectra typical of a flavoprotein, the former being produced as an apoprotein. Visible absorption spectra of R91A, L, and P were red shifted with maxima at 458 nm, while CD spectra indicated an altered FAD environment for all the mutants except R91K. Fluorescence spectra showed a reduced degree of intrinsic flavin fluorescence quenching for the R91K, A, and P, mutants, while thermal stability studies suggested all the mutants, except R91K, were somewhat less stable than the wild-type domain. Initial-rate kinetic measurements demonstrated that the mutants exhibited decreased NADH:ferricyanide reductase activity with the R91P mutant retaining the lowest activity, corresponding to a k(cat) of 283 s(-1) and a K(NADH)(m) of 105 microM, when compared to the wild-type domain (k(cat) = 800 s(-1) K(NADH)(m) = 6 microM). These results demonstrate that R91 is not essential for FAD binding in cb5r; however, mutation of R91 perturbs the flavin environment and alters both diaphorase substrate recognition and utilization.  相似文献   

19.
The Michaelis constant (K(m)) and V(mas) (E0k(cat)) values for two mutant sets of enzymes were studied from the viewpoint of their definition in a rapid equilibrium reaction model and in a steady state reaction model. The "AMP set enzyme" had a mutation at the AMP-binding site (Y95F, V67I, and V67I/L76V), and the "ATP set enzyme" had a mutation at a possible ATP-binding region (Y32F, Y34F, and Y32A/Y34A). Reaction rate constants obtained using steady state model analysis explained discrepancies found by the rapid equilibrium model analysis. (i) The unchanged number of bound AMPs for Y95F and the wild type despite the markedly increased K(m) values for AMP of the AMP set of enzymes was explained by alteration of the rate constants of the AMP step (k(+2), k(-2)) to retain the ratio k(+2)/k(-2). (ii) A 100 times weakened selectivity of ATP for Y34F in contrast to no marked changes in K(m) values for both ATP and AMP for the ATP set of enzymes was explained by the alteration of the rate constants of the ATP steps. A similar alteration of the K(m) and k(cat) values of these enzymes resulted from distinctive alterations of their rate constants. The pattern of alteration was highly suggestive. The most interesting finding was that the rate constants that decided the K(m) and k(cat) values were replaced by the mutation, and the simple relationships between K(m), k(cat), and the rate constants of K(m)1 = k(+1)/k(-1) and k(cat) = k(f) were not valid. The nature of the K(m) and k(cat) alterations was discussed.  相似文献   

20.
Yeast glyceraldehyde-3-phosphate dehydrogenase as a typical SH enzyme is inactivated by the antipodes of a-iodopropionic acid and its amide at different rates. The apoenzyme reacts faster with the D(+) antipode of the free a-iodopropionic acid (k(D)/k(L) = 6.8) and the L(-) antipode of the amide (k(L)/k(D) = 3). On addition of NAD(+) the stereoselectivity of the SH group towards a-iodopropionic acid is inverted, that towards the amide is enlarged, the rate relationships depending on the NAD(+) concentration.The results were interpreted by the assumption, that the allosteric T state of the enzyme reacts most rapidly with the D(+) antipodes, whereas the R state favours the L(-) antipodes of the alkylation reagents. The dependence of the reaction rates on the NAD(+) concentration could be fitted to the allosteric function of state R.  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