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.
Lehoux IE  Mitra B 《Biochemistry》2000,39(33):10055-10065
(S)-Mandelate dehydrogenase from Pseudomonas putida is an FMN-dependent alpha-hydroxy acid dehydrogenase. Structural studies of two homologous enzymes, glycolate oxidase and flavocytochrome b(2), indicated that a conserved arginine residue (R277 in MDH) interacts with the product carboxylate group [Lindqvist, Y., Branden, C.-I., Mathews, F. S., and Lederer, F. (1991) J. Biol. Chem. 266, 3198-3207]. The catalytic role of R277 was investigated by site-specific mutagenesis together with chemical rescue experiments. The R277K, R277G, R277H, and R277L proteins were generated and purified in active forms. The k(cat) for the charge-conserved mutation, R277K, was only 4-fold lower than wt-MDH, but its K(m) value was 40-fold lower; in contrast, k(cat)s for R277G, R277H, and R277L were 400-1000-fold lower than for wt-MDH and K(m) values were 5-15-fold lower compared to R277K. The K(d)s for negatively charged competitive inhibitors were relatively unaffected in all four R277 mutants. The k(cat) for R277G could be enhanced by the addition of exogenous guanidines or imidazoles; the maximum rescued k(cat) was approximately 70% of the wt-MDH value. Only reagents that were positively charged and could function as hydrogen bond donors were effective rescue agents. Our results indicate that R277 plays a major role in transition state stabilization through its positive charge-consistent with a mechanism involving a carbanion intermediate. The positive charge has a relatively small contribution toward substrate binding. R277 also forms a specific hydrogen bond with both the substrate and the transition state; this interaction contributes significantly to the low K(m) for (S)-mandelate.  相似文献   

2.
人双专一性磷酸酶活性位点Cys^124附近精氨酸突变及功能   总被引:1,自引:0,他引:1  
为研究人双专一性磷酸酶活性位点Cys12 4 附近 3个带正电的精氨酸对酶催化功能的影响 ,用QuikChange定点突变方法获得 6个突变体 :R12 5L、R130 L、R130 K、R130 L/S131A、R158K和R158L。将含突变基因的重组质粒转化大肠杆菌菌株BL2 1(DE3) ,经IPTG诱导表达获得的目的蛋白质均以可溶形式存在。通过镍离子亲和层析纯化得到纯度大于 90 %的蛋白质。对人痘苗病毒H1相关磷酸酶 (VHR)及其突变体进行稳态动力学参数和竞争性抑制常数Ki 的测定 ,结果显示上述Arg130 和Arg158突变体的kcat/Km 值都较野生型有大幅度下降 ,而Ki 值有明显上升 ,表明 130和 15 8位的精氨酸是VHR活性所必需 ,而且可能与底物上带负电的磷酸基团结合有关。另外 ,单突变体R130 L和双突变体R130 L/S131A之间的kcat值相差很小 ,提示Arg130 单点突变后可能破坏了Ser131与Cys12 4 间的氢键。再者 ,R12 5L、R130 L和R158L突变体都降低了砷酸盐结合亲和性 ,暗示这 3个精氨酸残基侧链上的正电荷可能有助于底物与酶的结合。  相似文献   

3.
Dewanti AR  Xu Y  Mitra B 《Biochemistry》2004,43(7):1883-1890
(S)-Mandelate dehydrogenase (MDH) from Pseudomonas putida is a flavin mononucleotide (FMN)-dependent enzyme that oxidizes (S)-mandelate to benzoylformate. In this work, we show that the ethyl and methyl esters of (S)-mandelic acid are substrates for MDH. Although the binding affinity of the neutral esters is 25-50-fold lower relative to the negatively charged (S)-mandelate, they are oxidized with comparable k(cat)s. Substrate analogues in which the carbonyl group on the C-1 carbon is replaced by other electron-withdrawing groups were not substrates. The requirement of a carbonyl group on the C-1 carbon in a substrate suggests that the negative charge developed during the reaction is stabilized by delocalization to the carbonyl oxygen. Arg277, a residue that is important in both binding and transition state stabilization for the activity with (S)-mandelate, is also critical for transition state stabilization for the esters, but not for their binding affinity. We previously showed that the substrate oxidation half-reaction with (S)-mandelate has two rate-limiting steps of similar activation energies and proceeds through the formation of a charge-transfer complex of an electron-rich donor and oxidized FMN [Dewanti, A. R., and Mitra, B. (2003) Biochemistry 42, 12893-12901]. This charge-transfer intermediate was observed with the neutral esters as well. The observation of this electron-rich intermediate for the oxidation of an uncharged substrate to an uncharged product, as well as the critical role of Arg277 in the reaction with the esters, provides further evidence that the MDH reaction mechanism is not a concerted transfer of a hydride ion from the substrate to the FMN, but involves the transient formation of a carbanion/ene(di)olate intermediate.  相似文献   

4.
Dewanti AR  Mitra B 《Biochemistry》2003,42(44):12893-12901
(S)-Mandelate dehydrogenase from Pseudomonas putida is a member of a FMN-dependent enzyme family that oxidizes (S)-alpha-hydroxyacids to alpha-ketoacids. The reductive half-reaction consists of the steps involved in substrate oxidation and FMN reduction. In this study, we investigated the mechanism of this half-reaction in detail. At low temperatures, a transient intermediate was formed in the course of the FMN reduction reaction. This intermediate is characteristic of a charge-transfer complex of oxidized FMN and an electron-rich donor and is formed prior to full reduction of the flavin. The intermediate was not due to binding of anionic substrates or inhibitors. It was only observed with efficient substrates that have high k(cat) values. At higher temperatures, it was formed within the dead time of the stopped-flow instrument. The rate of formation of the intermediate was 3-4-fold faster than its rate of disappearance; the former had a larger isotope effect. This suggests that the charge-transfer donor is an electron-rich carbanion/enolate intermediate that is generated by the base-catalyzed abstraction of the substrate alpha-proton. This is consistent with the observation that the intermediate was not observed with the R277K and R277G mutants, which have been shown to destabilize the carbanion intermediate (Lehoux, I. E., and Mitra, B. (2000) Biochemistry 39, 10055-10065). Thus, the MDH reaction has two rate-limiting steps of similar activation energies: the formation and breakdown of a distinct intermediate, with the latter step being slightly more rate limiting. We also show that MDH is capable of catalyzing the reverse reaction, the reoxidation of reduced MDH by the product ketoacid, benzoylformate. The transient intermediate was observed during the reverse reaction as well, confirming that it is indeed a true intermediate in the MDH reaction pathway.  相似文献   

5.
IMP dehydrogenase (IMPDH) catalyzes the oxidation of inosine 5'-monophosphate (IMP) to xanthosine 5'-monophosphate (XMP) and the reduction of NAD(+). The reaction involves formation of an E-XMP covalent intermediate; hydrolysis of the E-XMP intermediate is rate-limiting and requires the enzyme to adopt a closed conformation. Arg418 appears to act as the base that activates water for the hydrolysis reaction [Guillen-Schlippe, Y. V., and Hedstrom, L. (2005) Biochemistry 44, 11700-11707]. Deprotonation of Arg418 also stabilizes the closed conformation. Here we show that guanidine derivatives rescue the activity of the Arg418Ala variant. Amines and imidazole do not rescue. The rescue reaction appears to be saturable, with the values of K(R) ranging from 40 to 400 mM. The value of k(rescue) for the best rescue agents approaches the value of k(cat) for the reaction of the wild-type enzyme. Guanidine derivatives also rescue the activity of the Arg418Ala/Tyr419Phe variant. Multiple-inhibitor experiments suggest that the guanidine derivatives do not restore the equilibrium between open and closed conformations. Therefore, rescue agents must accelerate the hydrolysis of the E-XMP intermediate. The rate of the rescue reaction increases with an increase in pH, consistent with the hypothesis that the reaction involves neutral guanidine. A solvent D(2)O isotope effect is observed at low concentrations of the rescue agent, consistent with rate-limiting transfer of a proton from water. The value of k(cat) (rescue)/K(R)(base) correlates with the pK(a) of the guanidine derivative (Bronsted coefficient beta approximately 1). These results suggest that proton transfer from water to guanidine is almost complete in the transition state.  相似文献   

6.
Phosphotransacetylase (EC 2.3.1.8) catalyzes the reversible transfer of the acetyl group from acetyl phosphate to coenzyme A (CoA): CH(3)COOPO(3)(2-) + CoASH <==> CH(3)COSCoA + HPO(4)(2-). The role of arginine residues was investigated for the phosphotransacetylase from Methanosarcina thermophila. Kinetic analysis of a suite of variants indicated that Arg 87 and Arg 133 interact with the substrate CoA. Arg 87 variants were reduced in the ability to discriminate between CoA and the CoA analog 3'-dephospho-CoA, indicating that Arg 87 forms a salt bridge with the 3'-phosphate of CoA. Arg 133 is postulated to interact with the 5'-phosphate of CoA. Large decreases in k(cat) and k(cat)/K(m) for all of the Arg 87 and Arg 133 variants indicated that these residues are also important, although not essential, for catalysis. Large decreases in k(cat) and k(cat)/K(m) were also observed for the variants in which lysine replaced Arg 87 and Arg 133, suggesting that the bidentate interaction of these residues with CoA or their greater bulk is important for optimal activity. Desulfo-CoA is a strong competitive inhibitor of the enzyme, suggesting that the sulfhydryl group of CoA is important for the optimization of CoA-binding energy but not for tight substrate binding. Chemical modification of the wild-type enzyme by 2,3-butanedione and substrate protection by CoA indicated that at least one reactive arginine is in the active site and is important for activity. The inhibition pattern of the R87Q variant indicated that Arg 87 is modified, which contributes to the inactivation; however, at least one additional active-site arginine is modified leading to enzyme inactivation, albeit at a lower rate.  相似文献   

7.
To investigate the function of aspartic acid residue 101 and arginine residue 166 in the active site of Escherichia coli alkaline phosphatase (EAP), two single mutants D101S (Asp 101 →Ser) and R166K (Arg 166 →Lys) and a double mutant D101S/R166K of EAP were generated through site-directed mutagenesis based on over-lap PCR method. Their enzymatic kinetic properties, thermal stabilities and possible reaction mechanism were explored. In the presence of inorganic phosphate acceptor, 1 M diethanolamine buffer, the k cat for D101S mutant enzyme increased 10-fold compared to that of wild-type EAP. The mutant R166K has a 2-fold decrease of k cat relative to the wild-type EAP, but the double mutant D101S/R166K was in the middle of them, indicative of an additive effect of these two mutations. On the other hand, the catalytic efficiencies of mutant enzymes are all reduced because of a substantial increase of K m values. All three mutants were more resistant to phosphate inhibitor than the wild-type enzyme. The analysis of the kinetic data suggests that (1) the D101S mutant enzyme obtains a higher catalytic activity by allowing a faster release of the product; (2) the R166K mutant enzyme can reduce the binding of the substrate and phosphate competitive inhibitor; (3) the double mutant enzyme has characteristics of both quicker catalytic turnover number and decreased affinity for competitive inhibitor. Additionally, pre-steady-state kinetics of D101S and D101S/R166K mutants revealed a transient burst followed by a linear steady state phase, obviously different from that of wild-type EAP, suggesting that the rate-limiting step has partially change from the release of phosphate from non-covalent E-Pi complex to the hydrolysis of covalent E-Pi complex for these two mutants.  相似文献   

8.
To investigate the function of aspartic acid residue 101 and arginine residue 166 in the active site of Escherichia coli alkaline phosphatase (EAP), two single mutants D101S (Asp 101 &#77 Ser) and R166K (Arg 166 &#77 Lys) and a double mutant D101S/R166K of EAP were generated through site-directed mutagenesis based on over-lap PCR method. Their enzymatic kinetic properties, thermal stabilities and possible reaction mechanism were explored. In the presence of inorganic phosphate acceptor, 1 M diethanolamine buffer, the k cat for D101S mutant enzyme increased 10-fold compared to that of wild-type EAP. The mutant R166K has a 2-fold decrease of k cat relative to the wild-type EAP, but the double mutant D101S/R166K was in the middle of them, indicative of an additive effect of these two mutations. On the other hand, the catalytic efficiencies of mutant enzymes are all reduced because of a substantial increase of K m values. All three mutants were more resistant to phosphate inhibitor than the wild-type enzyme. The analysis of the kinetic data suggests that (1) the D101S mutant enzyme obtains a higher catalytic activity by allowing a faster release of the product; (2) the R166K mutant enzyme can reduce the binding of the substrate and phosphate competitive inhibitor; (3) the double mutant enzyme has characteristics of both quicker catalytic turnover number and decreased affinity for competitive inhibitor. Additionally, pre-steady-state kinetics of D101S and D101S/R166K mutants revealed a transient burst followed by a linear steady state phase, obviously different from that of wild-type EAP, suggesting that the rate-limiting step has partially change from the release of phosphate from non-covalent E-Pi complex to the hydrolysis of covalent E-Pi complex for these two mutants.  相似文献   

9.
Previous studies using chemical modification approach have shown the importance of arginine residues in bile salt activation of carboxyl ester lipase (CEL) activity. However, the x-ray crystal structure of CEL failed to show the involvement of arginine residues in CEL-bile salt interaction. The current study used a site-specific mutagenesis approach to determine the role of arginine residues 63 and 423 in bile salt-dependent and bile salt-independent hydrolytic activities of rat CEL. Mutations of Arg(63) to Ala(63) (R63A) and Arg(423) to Gly(423) (R423G) resulted in enzymes with increased bile salt-independent hydrolytic activity against lysophosphatidylcholine, having 6.5- and 2-fold higher k(cat) values, respectively, in comparison to wild type CEL. In contrast, the R63A and R423A mutant enzymes displayed 5- and 11-fold decreases in k(cat), in comparison with wild type CEL, for bile salt-dependent cholesteryl ester hydrolysis. Although taurocholate induced similar changes in circular dichroism spectra for wild type, R63A, and R423G proteins, this bile salt was less efficient in protecting the mutant enzymes against thermal inactivation in comparison with control CEL. Lipid binding studies revealed less R63A and R423G mutant CEL were bound to 1,2-diolein monolayer at saturation compared with wild type CEL. These results, along with computer modeling of the CEL protein, indicated that Arg(63) and Arg(423) are not involved directly with monomeric bile salt binding. However, these residues participate in micellar bile salt modulation of CEL enzymatic activity through intramolecular hydrogen bonding with the C-terminal domain. These residues are also important, probably through similar intramolecular hydrogen bond formation, in stabilizing the enzyme in solution and at the lipid-water interface.  相似文献   

10.
Segraves EN  Holman TR 《Biochemistry》2003,42(18):5236-5243
Mammalian lipoxygenases have been implicated in several inflammatory disorders; however, the details of the kinetic mechanism are still not well understood. In this paper, human platelet 12-lipoxygenase (12-hLO) and human reticulocyte 15-lipoxygenase-1 (15-hLO) were tested with arachidonic acid (AA) and linoleic acid (LA), respectively, under a variety of changing experimental conditions, such as temperature, dissolved oxygen concentration, and viscosity. The data that are presented show that 12-hLO and 15-hLO have slower rates of product release (k(cat)) than soybean lipoxygenase-1 (sLO-1), but similar or better rates of substrate capture for the fatty acid (k(cat)/K(M)) or molecular oxygen [k(cat)/K(M(O)2)]. The primary, kinetic isotope effect (KIE) for 15-hLO with LA was determined to be temperature-independent and large ((D)k(cat) = 40 +/- 8), over the range of 10-35 degrees C, indicating that C-H bond cleavage is the sole rate-limiting step and proceeds through a tunneling mechanism. The (D)k(cat)/K(M) for 15-hLO, however, was temperature-dependent, consistent with our previous results [Lewis, E. R., Johansen, E., and Holman, T. R. (1999) J. Am. Chem. Soc. 121, 1395-1396], indicating multiple rate-limiting steps. This was confirmed by a temperature-dependent, k(cat)/K(M) solvent isotope effect (SIE), which indicated a hydrogen bond rearrangement step at low temperatures, similar to that of sLO-1 [Glickman, M. H., and Klinman, J. P. (1995) Biochemistry 34, 14077-14092]. The KIE could not be determined for 12-hLO due to its inability to efficiently catalyze LA, but the k(cat)/K(M) SIE was temperature-independent, indicating distinct rate-limiting steps from both 15-hLO and sLO-1.  相似文献   

11.
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   

12.
Arginines R23, R178, R179 and R218 in thymidylate synthase (TS, EC 2. 1.1.45) are hydrogen bond donors to the phosphate moiety of the substrate, dUMP. In order to investigate how these arginines contribute to enzyme function, we prepared complete replacement sets of mutants at each of the four sites in Lactobacillus casei TS. Mutations of R23 increase K:(m) for dUMP 2-20-fold, increase K:(m) for cofactor 8-40-fold and decrease k(cat) 9-20-fold, reflecting the direct role of the R23 side chain in binding and orienting the cofactor in ternary complexes of the enzyme. Mutations of R178 increase K:(m) for dUMP 40-2000-fold, increase K:(m) for cofactor 3-20-fold and do not significantly affect k(cat). These results are consistent with the fact that this residue is an integral part of the dUMP-binding wall and contributes to the orientation and ordering of several other dUMP binding residues. Kinetic parameters for all R179 mutations except R179P were not significantly different from wild-type values, reflecting the fact that this external arginine does not directly contact the cofactor or other ligand-binding residues. R218 is essential for the structure of the catalytic site and all mutations of this arginine except R218K were inactive.  相似文献   

13.
The conjoint substitution of three active-site residues in aspartate aminotransferase (AspAT) of Escherichia coli (Y225R/R292K/R386A) increases the ratio of L-aspartate beta-decarboxylase activity to transaminase activity >25 million-fold. This result was achieved by combining an arginine shift mutation (Y225R/R386A) with a conservative substitution of a substrate-binding residue (R292K). In the wild-type enzyme, Arg(386) interacts with the alpha-carboxylate group of the substrate and is one of the four residues that are invariant in all aminotransferases; Tyr(225) is in its vicinity, forming a hydrogen bond with O-3' of the cofactor; and Arg(292) interacts with the distal carboxylate group of the substrate. In the triple-mutant enzyme, k(cat)' for beta-decarboxylation of L-aspartate was 0.08 s(-1), whereas k(cat)' for transamination was decreased to 0.01 s(-1). AspAT was thus converted into an L-aspartate beta-decarboxylase that catalyzes transamination as a side reaction. The major pathway of beta-decarboxylation directly produces L-alanine without intermediary formation of pyruvate. The various single- or double-mutant AspATs corresponding to the triple-mutant enzyme showed, with the exception of AspAT Y225R/R386A, no measurable or only very low beta-decarboxylase activity. The arginine shift mutation Y225R/R386A elicits beta-decarboxylase activity, whereas the R292K substitution suppresses transaminase activity. The reaction specificity of the triple-mutant enzyme is thus achieved in the same way as that of wild-type pyridoxal 5'-phosphate-dependent enzymes in general and possibly of many other enzymes, i.e. by accelerating the specific reaction and suppressing potential side reactions.  相似文献   

14.
Phosphite dehydrogenase (PTDH) from Pseudomonas stutzeri catalyzes the nicotinamide adenine dinucleotide-dependent oxidation of phosphite to phosphate. The enzyme belongs to the family of D-hydroxy acid dehydrogenases (DHDHs). A search of the protein databases uncovered many additional putative phosphite dehydrogenases. The genes encoding four diverse candidates were cloned and expressed, and the enzymes were purified and characterized. All oxidized phosphite to phosphate and had similar kinetic parameters despite a low level of pairwise sequence identity (39-72%). A recent crystal structure identified Arg301 as a residue in the active site that has not been investigated previously. Arg301 is fully conserved in the enzymes shown here to be PTDHs, but the residue is not conserved in other DHDHs. Kinetic analysis of site-directed mutants of this residue shows that it is important for efficient catalysis, with an ~100-fold decrease in k(cat) and an almost 700-fold increase in K(m,phosphite) for the R301A mutant. Interestingly, the R301K mutant displayed a slightly higher k(cat) than the parent PTDH, and a more modest increase in K(m) for phosphite (nearly 40-fold). Given these results, Arg301 may be involved in the binding and orientation of the phosphite substrate and/or play a catalytic role via electrostatic interactions. Three other residues in the active site region that are conserved in the PTDH orthologs but not DHDHs were identified (Trp134, Tyr139, and Ser295). The importance of these residues was also investigated by site-directed mutagenesis. All of the mutants had k(cat) values similar to that of the wild-type enzyme, indicating these residues are not important for catalysis.  相似文献   

15.
Mutagenesis of the conserved glutamic acid of influenza type A (E277) and Micromonospora viridifaciens (E260) sialidases was performed to probe the contribution of this strictly conserved residue to catalysis. Kinetic studies of the E260D and E260C M. viridifaciens mutant enzymes reveal that the overall mechanism of action has not changed. That is, the mutants are retaining sialidases in which glycosylation and deglycosylation are rate-limiting for k(cat)/K(m) and k(cat), respectively. The solvent kinetic isotope effect and proton inventory on k(cat) for the E260C mutant sialidase provide strong evidence that the newly installed cysteine residue provides little catalytic acceleration. The results are consistent with the conserved aspartic acid residue (D92) becoming the key general acid/base residue in the catalytic cycle. In addition, the E277D mutant influenza type A sialidase is catalytically active toward 4-nitrophenyl α-D-sialoside, although no measurable hydrolysis of natural substrates was observed. Thus, mutating the glutamate residue (E277) to an aspartate increases the activation free energy of hydrolysis for natural substrates by >22 kJ/mol.  相似文献   

16.
M Slatner  B Nidetzky  K D Kulbe 《Biochemistry》1999,38(32):10489-10498
To characterize catalysis by NAD-dependent long-chain mannitol 2-dehydrogenases (MDHs), the recombinant wild-type MDH from Pseudomonas fluorescens was overexpressed in Escherichia coli and purified. The enzyme is a functional monomer of 54 kDa, which does not contain Zn(2+) and has B-type stereospecificity with respect to hydride transfer from NADH. Analysis of initial velocity patterns together with product and substrate inhibition patterns and comparison of primary deuterium isotope effects on the apparent kinetic parameters, (D)k(cat), (D)(k(cat)/K(NADH)), and (D)(k(cat)/K(fructose)), show that MDH has an ordered kinetic mechanism at pH 8.2 in which NADH adds before D-fructose, and D-mannitol and NAD are released in that order. Isomerization of E-NAD to a form which interacts with D-mannitol nonproductively or dissociation of NAD from the binary complex after isomerization is the slowest step (>/=110 s(-)(1)) in D-fructose reduction at pH 8.2. Release of NADH from E-NADH (32 s(-)(1)) is the major rate-limiting step in mannitol oxidation at this pH. At the pH optimum for D-fructose reduction (pH 7.0), the rate of hydride transfer contributes significantly to rate limitation of the catalytic cascade and the overall reaction. (D)(k(cat)/K(fructose)) decreases from 2.57 at pH 7.0 to a value of 相似文献   

17.
Jao SC  Huang LF  Hwang SM  Li WS 《Biochemistry》2006,45(6):1547-1553
Analysis of the pH-rate profile for catalysis of bradykinin cleavage by aminopeptidase P (AMPP), a manganese-containing hydrolase from Escherichia coli, was carried out to show that optimal catalytic function is obtained at neutral pH. On the basis of information derived from the crystal structure, peptidase sequence alignments, and the hydrolysis of organophosphate triesters, active site residues Arg153, Arg370, Trp88, Tyr387, and Arg404 were identified as potential catalytic residues. Site-directed mutagenesis was used to substitute these residues with Leu, Ala, Trp, Lys, or Phe. The kcat values for the Arg153, Arg370, and Trp88 mutants were nearly the same as that for the wild-type enzyme. The kcat values of the R404K, R404A, and Y387A mutants were lower by factors of 285, 400, and 16, respectively. Inductively coupled plasma mass spectrometry and circular dichroism spectroscopy showed that Arg404 is not required for metal chelation or stabilization of protein secondary structure. The hydrogen bond network observed between the side chains of conserved residues Asp260, Arg404, and Tyr387 indicated that Arg404 participates in proton relay. This was further evidenced by the return of activity in the R404A mutant by the addition of guanidine. Also, reduced catalytic efficiency in the R404K mutant, which conserves the positive charge at the bridge site, shows that only the arginine group of Arg404 (not the ammonium group of Lys404) can participate in the hydrogen bond network. The hydrogen bond interaction between the Arg404 and the Tyr387 ring hydroxyl group is suggested by the reduced catalytic efficiency of the Y387F mutant.  相似文献   

18.
Clark DD  Boyd JM  Ensign SA 《Biochemistry》2004,43(21):6763-6771
2-[(R)-2-Hydroxypropylthio]ethanesulfonate (R-HPC) dehydrogenase (DH) catalyzes the reversible oxidation of R-HPC to 2-(2-ketopropylthio)ethanesulfonate (2-KPC) in a key reaction in the bacterial conversion of chiral epoxides to beta-keto acids. R-HPCDH is highly specific for the R-enantiomer of HPC, while a separate enzyme, S-HPCDH, catalyzes the oxidation of the corresponding S-enantiomer. In the present study, the features of substrate and enzyme imparting stereospecificity have been investigated for R-HPCDH. S-HPC was a substrate for R-HPCDH with a K(m) identical to that for R-HPC but with a k(cat) 600 times lower. Achiral 2-propanol and short-chain (R)- and (S)-2-alkanols were substrates for R-HPCDH. For (R)-alkanols, as the carbon chain length increased, K(m) decreased, with the K(m) for (R)-2-octanol being 1700 times lower than for 2-propanol. At the same time, k(cat) changed very little and was at least 90% lower than k(cat) for R-HPC and at least 22 times higher than k(cat) for S-HPC. (S)-2-Butanol and (S)-2-pentanol were substrates for R-HPCDH. The K(m) for (S)-2-butanol was identical to that for (R)-2-butanol, while the K(m) for (S)-2-pentanol was 7.5 times higher than for (R)-2-pentanol. Longer chain (S)-2-alkanols were sufficiently poor substrates for R-HPCDH that kinetic parameters could not be determined. Mutagenesis of C-terminal arginine residues of R-HPCDH revealed that R152 and R196 are essential for effective catalysis with the natural substrates R-HPC and 2-KPC but not for catalysis with 2-alkanols or ketones as substrates. Short-chain alkylsulfonates and coenzyme M (2-mercaptoethanesulfonate) were found to modify the kinetic parameters for 2-butanone reduction by R-HPCDH in a saturable fashion, with the general effect of increasing k(cat), decreasing K(m), and increasing the enantioselectivity of 2-butanone reduction to a theoretical value of 100% (S)-2-butanol. The modulating effects of ethanesulfonate and propanesulfonate provided thermodynamic binding constants close to K(m) for the natural substrates R-HPC and 2-KPC. The effects of alkylsulfonates on modulating the enantioselectivity and kinetic properties of R-HPCDH were abolished in R152A and R196A mutants but not in mutants of other C-terminal arginine residues. Collectively, the results suggest that interactions between the sulfonate of CoM and specific arginine residues are key to the enantioselectivity and catalytic efficiency of R-HPCDH. A model is proposed wherein sulfonate-arginine interactions within an alkylsulfonate binding pocket control the catalytic properties of R-HPCDH.  相似文献   

19.
GDP-mannose hydrolase (GDPMH) catalyzes the hydrolysis of GDP-alpha-d-sugars by nucleophilic substitution with inversion at the anomeric C1 atom of the sugar, with general base catalysis by H124. Three lines of evidence indicate a mechanism with dissociative character. First, in the 1.3 A X-ray structure of the GDPMH-Mg(2+)-GDP.Tris(+) complex [Gabelli, S. B., et al. (2004) Structure 12, 927-935], the GDP leaving group interacts with five catalytic components: R37, Y103, R52, R65, and the essential Mg(2+). As determined by the effects of site-specific mutants on k(cat), these components contribute factors of 24-, 100-, 309-, 24-, and >/=10(5)-fold, respectively, to catalysis. Both R37 and Y103 bind the beta-phosphate of GDP and are only 5.0 A apart. Accordingly, the R37Q/Y103F double mutant exhibits partially additive effects of the two single mutants on k(cat), indicating cooperativity of R37 and Y103 in promoting catalysis, and antagonistic effects on K(m). Second, the conserved residue, D22, is positioned to accept a hydrogen bond from the C2-OH group of the sugar undergoing substitution at C1, as was shown by modeling an alpha-d-mannosyl group into the sugar binding site. The D22A and D22N mutations decreased k(cat) by factors of 10(2.1) and 10(2.6), respectively, for the hydrolysis of GDP-alpha-d-mannose, and showed smaller effects on K(m), suggesting that the D22 anion stabilizes a cationic oxocarbenium transition state. Third, the fluorinated substrate, GDP-2F-alpha-d-mannose, for which a cationic oxocarbenium transition state would be destabilized by electron withdrawal, exhibited a 16-fold decrease in k(cat) and a smaller, 2.5-fold increase in K(m). The D22A and D22N mutations further decreased the k(cat) with GDP-2F-alpha-d-mannose to values similar to those found with GDP-alpha-d-mannose, and decreased the K(m) of the fluorinated substrate. The choice of histidine as the general base over glutamate, the preferred base in other Nudix enzymes, is not due to the greater basicity of histidine, since the pK(a) of E124 in the active complex (7.7) exceeded that of H124 (6.7), and the H124E mutation showed a 10(2.2)-fold decrease in k(cat) and a 4.0-fold increase in K(m) at pH 9.3. Similarly, the catalytic triad detected in the X-ray structure (H124- - -Y127- - -P120) is unnecessary for orienting H124, since the Y127F mutation had only 2-fold effects on k(cat) and K(m) with either H124 or E124 as the general base. Hence, a neutral histidine rather than an anionic glutamate may be necessary to preserve electroneutrality in the active complex.  相似文献   

20.
The functional role of the highly conserved active site Arg 59 in the prototype of the gamma-class carbonic anhydrase Cam (carbonic anhydrase from Methanosarcina thermophila) was investigated. Variants (R59A, -C, -E, -H, -K, -M, and -Q) were prepared by site-directed mutagenesis and characterized by size exclusion chromatography (SEC), circular dichroism (CD) spectroscopy, and stopped-flow kinetic analyses. CD spectra indicated similar secondary structures for the wild type and the R59A and -K variants, independent of nondenaturing concentrations of guanidine hydrochloride (GdnHCl). SEC indicated that all variants purified as homotrimers like the wild type. SEC also revealed that the R59A and -K variants unfolded at > or = 1.5 M GdnHCl, compared to 3.0 M GdnHCl for the wild type. These results indicate that Arg 59 contributes to the thermodynamic stability of the Cam trimer. The R59K variant had k(cat) and k(cat)/K(m) values that were 8 and 5% of the wild-type values, respectively, while all other variants had k(cat) and k(cat)/K(m) values 10-100-fold lower than those of the wild type. The R59A, -C, -E, -M, and -Q variants exhibited 4-63-fold increases in k(cat) and 9-120-fold increases in k(cat)/K(m) upon addition of 100 mM GdnHCl, with the largest increases observed for the R59A variant, which was comparable to the R59K variant. The kinetic results indicate that a positive charge at position 59 is essential for the CO(2) hydration step of the overall catalytic mechanism.  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