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 17 毫秒
1.
The basidiomycete Phanerochaete chrysosporium produces two glycoside hydrolase family 1 intracellular beta-glucosidases, BGL1A and BGL1B, during the course of cellulose degradation. In order to clarify the catalytic difference between two enzymes, in spite of their high similarity in amino acid sequences (65%), five amino acids around the catalytic site of BGL1A were individually mutated to those of BGL1B (V173C, M177L, D229N, H231D, and K253A), and the effects of the mutations on cellobiose hydrolysis were evaluated. When the kinetic parameters (K(m) and k(cat)) were compared at the optimum pH for the wild-type enzyme, the kinetic efficiency was decreased in the cases of D229N, H231D, and K253A, but not V173C or M177L. The pH dependence of cellobiose hydrolysis showed a significantly more acidic pH profile for the D229N mutant, compared with the wild-type enzyme. Since D229 is located between K253 and the putative acid/base catalyst E170, we prepared the double mutant D229N/K253A, and found that its hydrolytic activity at neutral pH was restored to that of the wild-type enzyme. Our results indicate that the interaction between D229 and K253 is critical for the pH dependence and catalytic activity of BGL1A. Biotechnol. Bioeng.  相似文献   

2.
Dreier J  Li Q  Khosla C 《Biochemistry》2001,40(41):12407-12411
Fatty acids and polyketides are synthesized by mechanistically and evolutionarily related multienzyme systems. Their carbon chain backbones are synthesized via repeated decarboxylative condensations of alpha-carboxylated building blocks onto a growing acyl chain. These alpha-carboxylated building blocks are transferred from the corresponding coenzyme A thioesters onto the phosphopantetheine arm of an acyl carrier protein (ACP) by acyl transferases, which operate by a ping-pong mechanism involving an acyl-O-serine intermediate. In the course of our studies on the malonyl-CoA:ACP transacylase (MAT) from Streptomyces coelicolor, we observed that an active-site Ser (97) --> Ala mutant retains activity as well as the ability to be covalently labeled by (14)C malonyl-CoA. Here we demonstrate that an alternative, catalytically competent nucleophile exists in the active site of this enzyme. Next to the active-site serine is a histidine residue that is conserved in some, but not all acyl transferases. The H96A mutant is also active and can be labeled, but an H96A/S97A double mutant is inactive and cannot be labeled. The ability of H96 to form a malonyl-imidazole adduct was confirmed by proteolysis, followed by radio-HPLC and mass spectrometric analysis of the S97A mutant enzyme. Kinetic analysis revealed that the k(cat) of the S97A mutant was within 10-fold that of the wild-type enzyme, whereas the K(M)s of the two enzymes were comparable. Sequence comparison with the E. coli MAT (whose X-ray structure is known) led to the identification of H201 as the putative base in the serine-histidine catalytic dyad of the S. coelicolor enzyme. The absence of MAT activity in the H201A mutant and the detection of weak activity in the H201Q mutant was consistent with this proposal. The implications of this unexpected finding are discussed.  相似文献   

3.
Two distinct genes encode the 93% homologous type 1 (placenta, peripheral tissues) and type 2 (adrenals, gonads) 3beta-hydroxysteroid dehydrogenase/isomerase (3beta-HSD/isomerase) in humans. Mutagenesis studies using the type 1 enzyme have produced the Y154F and K158Q mutant enzymes in the Y(154)-P-H(156)-S-K(158) motif as well as the Y269S and K273Q mutants from a second motif, Y(269)-T-L-S-K(273), both of which are present in the primary structure of the human type 1 3beta-HSD/isomerase. In addition, the H156Y mutant of the type 1 enzyme has created a chimera of the type 2 enzyme motif (Y(154)-P-Y(156)-S-K(158)) in the type 1 enzyme. The mutant and wild-type enzymes have been expressed and purified. The K(m) value of dehydroepiandrosterone is 13-fold greater, and the maximal turnover rate (K(cat)) is 2-fold greater for wild-type 2 3beta-HSD compared with the wild-type 1 3beta-HSD activity. The H156Y mutant of the type 1 enzyme has substrate kinetic constants for 3beta-HSD activity that are very similar to those of the wild-type 2 enzyme. Dixon analysis shows that epostane inhibits the 3beta-HSD activity of the wild-type 1 enzyme with 14-17-fold greater affinity compared with the wild-type 2 and H156Y enzymes. The Y154F and K158Q mutants exhibit no 3beta-HSD activity, have substantial isomerase activity, and utilize substrate with K(m) values similar to those of wild-type 1 isomerase. The Y269S and K273Q mutants have low, pH-dependent 3beta-HSD activity, exhibit only 5% of the maximal isomerase activity, and utilize the isomerase substrate very poorly. From these studies, a structural basis for the profound differences in the substrate and inhibition kinetics of the wild-type 1 and 2 3beta-HSD, plus a catalytic role for the Tyr(154) and Lys(158) residues in the 3beta-HSD reaction have been identified. These advances in our understanding of the structure/function of human type 1 and 2 3beta-HSD/isomerase may lead to the design of selective inhibitors of the type 1 enzyme not only in placenta to control the onset of labor but also in hormone-sensitive breast, prostate, and choriocarcinoma tumors to slow their growth.  相似文献   

4.
Kumamolisin-As is an acid collagenase with a subtilisin-like fold. Its active site contains a unique catalytic triad, Ser278-Glu78-Asp82, and a putative transition-state stabilizing residue, Asp164. In this study, the mutants D164N and E78H/D164N were engineered in order to replace parts of the catalytic machinery of kumamolisin-As with the residues found in the equivalent positions in subtilisin. Unlike the wild-type and D164N proenzymes, which undergo instantaneous processing to produce their 37-kDa mature forms, the expressed E78H/D164N proenzyme exists as an equilibrated mixture of the nicked and intact forms of the precursor. X-ray crystallographic structures of the mature forms of the two mutants showed that, in each of them, the catalytic Ser278 makes direct hydrogen bonds with the side chain of Asn164. In addition, His78 of the double mutant is distant from Ser278 and Asp82, and the catalytic triad no longer exists. Consistent with these structural alterations around the active site, these mutants showed only low catalytic activity (relative k(cat) at pH 4.0 1.3% for D164N and 0.0001% for E78H/D164N). pH-dependent kinetic studies showed that the single D164N substitution did not significantly alter the logk(cat) vs. pH and log(k(cat)/Km) vs. pH profiles of the enzyme. In contrast, the double mutation resulted in a dramatic switch of the logk(cat) vs. pH profile to one that was consistent with catalysis by means of the Ser278-His78 dyad and Asn164, which may also account for the observed ligation/cleavage equilibrium of the precursor of E78H/D164N. These results corroborate the mechanistic importance of the glutamate-mediated catalytic triad and oxyanion-stabilizing aspartic acid residue for low-pH peptidase activity of the enzyme.  相似文献   

5.
P Carter  L Abrahmsén  J A Wells 《Biochemistry》1991,30(25):6142-6148
A mutant of the serine protease, subtilisin BPN', in which the catalytic His64 is replaced by Ala (H64A), is very specific for substrates containing a histidine, presumably by the substrate-bound histidine assisting in catalysis [Carter, P., & Wells, J.A. (1987) Science (Washington, D.C.) 237, 394-399]. Here we probe the catalytic mechanism of H64A subtilisin for cleaving His and non-His substrates. We show that the ratio of aminolysis to hydrolysis is the same for ester and amide substrates as catalyzed by the H64A subtilisin. This is consistent with formation of a common acyl-enzyme intermediate for H64A subtilisin, analogous to the mechanism of the wild-type enzyme. However, the catalytic efficiencies (kcat/KM) for amidase and esterase activities with His-containing substrates are reduced by 5000-fold and 14-fold, respectively, relative to wild-type subtilisin BPN, suggesting that acylation is more compromised than deacylation in the H64A mutant. High concentrations of imidazole are much less effective than His substrates in promoting hydrolysis by the H64A variant, suggesting that the His residue on the bound (not free) substrate is involved in catalysis. The reduction in catalytic efficiency kcat/KM for hydrolysis of the amide substrate upon replacement of the oxyanion stabilizing asparagine (N155G) is only 7-fold greater for wild-type than H64A subtilisin. In contrast, the reductions in kcat/KM upon replacement of the catalytic serine (S221A) or aspartate (D32A) are about 3000-fold greater for wild-type than H64A subtilisin, suggesting that the functional interactions between the Asp32 and Ser221 with the substrate histidine are more compromised in substrate-assisted catalysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Qian J  Khandogin J  West AH  Cook PF 《Biochemistry》2008,47(26):6851-6858
Homocitrate synthase (acetyl-coenzyme A: 2-ketoglutarate C-transferase; E.C. 2.3.3.14) (HCS) catalyzes the condensation of acetyl-CoA (AcCoA) and alpha-ketoglutarate (alpha-KG) to give homocitrate and CoA. Although the structure of an HCS has not been solved, the structure of isopropylmalate synthase (IPMS), a homologue, has been solved (Koon, N., Squire, C. J., and Baker, E. N. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 8295-8300). Three active site residues in IPMS, Glu-218, His-379, and Tyr-410, were proposed as candidates for catalytic residues involved in deprotonation of the methyl group of AcCoA prior to the Claisen condensation to give homocitrylCoA. All three of the active site residues in IPMS are conserved in the HCS from Saccharomyces cerevisiae. Site-directed mutagenesis has been carried out to probe the role of the homologous residues, Glu-155, His-309, and Tyr-320, in the S. cerevisiae HCS. No detectable activity was observed for the H309A and H309N mutant enzyme, but a slight increase in activity was observed for H309A in the presence of 300 mM imidazole, which is still 1000-fold lower than that of wild type (wt). The E155Q and E155A mutant enzymes exhibited 1000-fold lower activity than wt. The activity of E155A, but not of E155Q, could be partially rescued by formate; a K act of 60 mM with a modest 4-fold maximum activation was observed. In the presence of formate, E155A gives k cat, K AcCoA, and K alpha-KG values of 0.0031 s (-1), 13 muM, and 39 microM, respectively, while a primary kinetic deuterium isotope effect of about 1.4 was obtained on V, with deuterium in the methyl of AcCoA. The pH dependence of k cat for E155A in the presence of formate gave a p K a of 7.9 for a group that must be protonated for optimum activity, similar to that observed for the wt enzyme. However, a partial change was observed on the acid side of the profile, compared to the all or none change observed for wt giving a p K a of about 6.7. The k cat for E155Q decreased at high pH, similar to the wt enzyme, but was pH independent at low pH. The Y320F mutant enzyme only lost 25-fold activity compared to that of the wt, giving k cat, K AcCoA, and K alpha-KG values of 0.039 s (-1), 33 microM, and 140 microM, respectively, and a primary kinetic deuterium isotope effect of 1.3 and 1.8 on V/ K AcCoA and V, respectively; the pH dependence of k cat was similar to that of the wt. These data, combined with a constant pH molecular dynamics simulation study, suggest that a catalytic dyad comprising Glu-155 and His-309 acts to deprotonate the methyl group of AcCoA, while Tyr320 is likely not directly involved in catalysis, but may aid in orienting the reactant and/or the catalytic dyad.  相似文献   

7.
Marks GT  Susler M  Harrison DH 《Biochemistry》2004,43(13):3802-3813
Two detailed mechanisms [Marks et al. (2001) Biochemistry 40, 6805] have been proposed to explain the activity of methylglyoxal synthase (MGS), a homohexameric allosterically regulated enzyme that catalyzes the elimination of phosphate from DHAP to form enol pyruvaldehyde. This enol then tautomerizes to methylglyoxal in solution. In one of these mechanisms His 98 plays an obligate role in the transfer of a proton from the O(3) oxygen of DHAP to the O2 oxygen. To test this hypothesized mechanism, the variants H98N and H98Q were expressed and purified. Relative to the wild-type enzyme, the H98N variant shows a 50-fold decrease in k(cat) with all other kinetic parameters (i.e., K(m), K(PGA), etc.) being nearly the same. By contrast, the apparent catalytic rate for the H98Q variant is 250-fold lower than that of the wild-type enzyme. Inorganic phosphate acts as a competitive inhibitor (with a K(i) of 15 microM) rather than as an allosteric-type inhibitor as it does in the wild-type enzyme, and the competitive inhibitor phosphoglyolate (PGA) acts as an activator of this variant. These facts are explained by a shift in the conformational equilibrium toward an "inactive" state. When activation by PGA is accounted for, the catalytic rate for the "active" state of H98Q is estimated to be only 6-fold less than that of the wild-type enzyme, and thus His 98 is not essential for catalysis. Primary deuterium isotope effect data were measured for the wild-type enzyme and the two variants. At pH 7.0, the (D)V isotope effect (1.5) and the absence of a (D)(V/K) isotope effect for the wild-type enzyme suggest that the rate for the isotopically sensitive step is partially rate limiting but greater than the rate of substrate dissociation. Both the (D)V (2.0) and (D)(V/K) (3.4) isotope effects were more pronounced in the H98N variant, while the H98Q variant displayed an unusual inverse (D)V (0.8) isotope effect and a normal (D)(V/K) (1.5) isotope effect. The X-ray crystal structures of PGA bound to the H98Q and H98N variants were both determined to a resolution of 2.2 A. These mutations had little effect on the overall structure. However, the pattern of hydrogen bonding in the active site suggests an explanation as to how in the wild-type and H98N mutated enzymes the "inactive to active" equilibrium lies toward the active state, while with the H98Q mutated enzyme the equilibrium lies toward the inactive state. Thus, the role of His 98 appears to be more as a regulator of the enzyme's conformation rather than as a critical contributor to the catalytic mechanism.  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 →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   

9.
Replacing several serine and threonine residues on the Ser/Thr surface of the xylanase from Aspergillus niger BCC14405 with four and five arginines effectively increases the thermostability of the enzyme. The modified enzymes showed 80% of maximal activity after incubating in xylan substrate for 2h at 50 degrees C compared to only 15% activity for wild-type enzyme. The half-life of the mutated enzymes increased to 257+/-16 and 285+/-10 min for the four- and five-arginine mutants, respectively, compared to 14+/-1 min for the wild-type enzyme. Thus, the arginine substitutions effectively increase stability by 18-20-fold. Kinetic parameters of the four-arginine-substitution enzyme were maintained at the level of the wild-type enzyme with the K(m) and V(max) values of 8.3+/-0.1 mgml(-1) and 9556+/-66 (n=3) U mg(-1) protein, respectively. The five-arginine-substitution enzyme showed only slight alteration in K(m) and V(max) with K(m) of 11.7+/-1.7 mgml(-1) and V(max) of 8502+/-65 Umg(-1) protein, indicating lower substrate affinity and catalytic rate. Our study demonstrated that properly introduced arginine residues on the Ser/Thr surface of xylanase family 11 might be very effective in improvement of enzyme thermostability.  相似文献   

10.
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   

11.
The role of Ser 167 of Escherichia coli thymidylate synthase (TS) in catalysis has been characterized by kinetic and crystallographic studies. Position 167 variants including S167A, S167N, S167D, S167C, S167G, S167L, S167T, and S167V were generated by site-directed mutagenesis. Only S167A, S167G, S167T, and S167C complemented the growth of thymidine auxotrophs of E. coli in medium lacking thymidine. Steady-state kinetic analysis revealed that mutant enzymes exhibited k(cat) values 1.1-95-fold lower than that of the wild-type enzyme. Relative to wild-type TS, K(m) values of the mutant enzymes for 2'-deoxyuridylate (dUMP) were 5-90 times higher, while K(m) values for 5,10-methylenetetrahydrofolate (CH(2)H(4)folate) were 1.5-16-fold higher. The rate of dehalogenation of 5-bromo-2'-deoxyuridine 5'-monophosphate (BrdUMP), a reaction catalyzed by TS that does not require CH(2)H(4)folate as cosubstrate, by mutant TSs was analyzed and showed that only S167A and S167G catalyzed the dehalogenation reaction and values of k(cat)/K(m) for the mutant enzymes were decreased by 10- and 3000-fold, respectively. Analysis of pre-steady-state kinetics of ternary complex formation revealed that the productive binding of CH(2)H(4)folate is weaker to mutant TSs than to the wild-type enzyme. Chemical transformation constants (k(chem)) for the mutant enzymes were lower by 1.1-6.0-fold relative to the wild-type enzyme. S167A, S167T, and S167C crystallized in the I2(1)3 space group and scattered X-rays to either 1.7 A (S167A and S167T) or 2.6 A (S167C). The high-resolution data sets were refined to a R(crys) of 19.9%. In the crystals some cysteine residues were derivatized with 2-mercaptoethanol to form S,S-(2-hydroxyethyl)thiocysteine. The pattern of derivatization indicates that in the absence of bound substrate the catalytic cysteine is not more reactive than other cysteines. It is proposed that the catalytic cysteine is activated by substrate binding by a proton-transfer mechanism in which the phosphate group of the nucleotide neutralizes the charge of Arg 126', facilitating the transfer of a proton from the catalytic cysteine to a His 207-Asp 205 diad via a system of ordered water molecules.  相似文献   

12.
The Saccharomyces cerevisiae CKI1-encoded choline kinase catalyzes the committed step in phosphatidylcholine synthesis via the Kennedy pathway. The enzyme is phosphorylated on multiple serine residues, and some of this phosphorylation is mediated by protein kinase A. In this work we examined the hypothesis that choline kinase is also phosphorylated by protein kinase C. Using choline kinase as a substrate, protein kinase C activity was dose- and time-dependent and dependent on the concentrations of choline kinase (K(m) = 27 microg/ml) and ATP (K(m) = 15 microM). This phosphorylation, which occurred on a serine residue, was accompanied by a 1.6-fold stimulation of choline kinase activity. The synthetic peptide SRSSSQRRHS (V5max/K(m) = 17.5 mm(-1) micromol min(-1) mg(-1)) that contains the protein kinase C motif for Ser25 was a substrate for protein kinase C. A Ser25 to Ala (S25A) mutation in choline kinase resulted in a 60% decrease in protein kinase C phosphorylation of the enzyme. Phosphopeptide mapping analysis of the S25A mutant enzyme confirmed that Ser25 was a protein kinase C target site. In vivo the S25A mutation correlated with a decrease (55%) in phosphatidylcholine synthesis via the Kennedy pathway, whereas an S25D phosphorylation site mimic correlated with an increase (44%) in phosphatidylcholine synthesis. Although the S25A (protein kinase C site) mutation did not affect the phosphorylation of choline kinase by protein kinase A, the S30A (protein kinase A site) mutation caused a 46% reduction in enzyme phosphorylation by protein kinase C. A choline kinase synthetic peptide (SQRRHSLTRQ) containing Ser30 was a substrate (V(max)/K(m) = 3.0 mm(-1) micromol min(-1) mg(-1)) for protein kinase C. Comparison of phosphopeptide maps of the wild type and S30A mutant choline kinase enzymes phosphorylated by protein kinase C confirmed that Ser30 was also a target site for protein kinase C.  相似文献   

13.
The catalytic domains of the pterin-dependent enzymes phenylalanine hydroxylase and tyrosine hydroxylase are homologous, yet differ in their substrate specificities. To probe the structural basis for the differences in specificity, seven residues in the active site of phenylalanine hydroxylase whose side chains are dissimilar in the two enzymes were mutated to the corresponding residues in tyrosine hydroxylase. Analysis of the effects of the mutations on the isolated catalytic domain of phenylalanine hydroxylase identified three residues that contribute to the ability to hydroxylate tyrosine, His264, Tyr277, and Val379. These mutations were incorporated into full-length phenylalanine hydroxylase and the complementary mutations into tyrosine hydroxylase. The steady-state kinetic parameters of the mutated enzymes showed that the identity of the residue in tyrosine hydroxylase at the position corresponding to position 379 of phenylalanine hydroxylase is critical for dihydroxyphenylalanine formation. The relative specificity of tyrosine hydroxylase for phenylalanine versus tyrosine, as measured by the (V/K(phe))/(V/K(tyr)) value, increased by 80000-fold in the D425V enzyme. However, mutation of the corresponding valine 379 of phenylalanine hydroxylase to aspartate was not sufficient to allow phenylalanine hydroxylase to form dihydroxyphenylalanine at rates comparable to that of tyrosine hydroxylase. The double mutant V379D/H264Q PheH was the most active at tyrosine hydroxylation, showing a 3000-fold decrease in the (V/K(phe))/(V/K(tyr)) value.  相似文献   

14.
Seventeen mutations in beta-subunit of Escherichia coli F1-ATPase which had previously been characterized in strain AN1272 (Mu-induced mutant) were expressed in strain JP17 (beta-subunit gene deletion). Six showed unchanged behavior, namely: C137Y; G142D; G146S; G207D; Y297F; and Y354F. Five failed to assemble F1F0 correctly, namely: G149I; G154I; G149I,G154I; G223D; and P403S,G415D. Six assembled F1F0 correctly, but with membrane ATPase lower than in AN1272, namely: K155Q; K155E; E181Q; E192Q; D242N; and D242V. AN1272 was shown to unexpectedly produce a small amount of wild-type beta-subunit; F1-ATPase activities reported previously in AN1272 were referable to hybrid enzymes containing both mutant and wild-type beta-subunits. Purified F1 was obtained from K155Q; K155E; E181Q; E192Q; and D242N mutants in JP17. Vmax ATPase values were lower, and unisite catalysis rate and equilibrium constants were perturbed to greater extent, than in AN1272. However, general patterns of perturbation revealed by difference energy diagrams were similar to those seen previously, and the new data correlated well in linear free energy relationships for reaction steps of unisite catalysis. Correlation between multisite and unisite ATPase activity was seen in the new enzymes. Overall, the data give strong support to previously proposed mechanisms of unisite catalysis, steady-state catalysis, and energy coupling in F1-ATPases (Al-Shawi, M. K., Parsonage, D. and Senior, A. E. (1990) J. Biol. Chem. 265, 4402-4410). The K155Q, K155E, D242N, and E181Q mutations caused 5000-fold, 4000-fold, 1800-fold, and 700-fold decrease, respectively, in Vmax ATPase, implying possibly direct roles for these residues in catalysis. Experiments with the D242N mutant suggested a role for residue beta D242 in catalytic site Mg2+ binding.  相似文献   

15.
S-Adenosylhomocysteine hydrolase (AdoHcyase) catalyzes the hydrolysis of S-adenosylhomocysteine to form adenosine and homocysteine. On the bases of crystal structures of the wild type enzyme and the D244E mutated enzyme complexed with 3'-keto-adenosine (D244E.Ado*), we have identified the important amino acid residues, Asp-130, Lys-185, Asp-189, and Asn-190, for the catalytic reaction and have proposed a catalytic mechanism (Komoto, J., Huang, Y., Gomi, T., Ogawa, H., Takata, Y., Fujioka, M., and Takusagawa, F. (2000) J. Biol. Chem. 275, 32147-32156). To confirm the proposed catalytic mechanism, we have made the D130N, K185N, D189N, and N190S mutated enzymes and measured the catalytic activities. The catalytic rates (k(cat)) of D130N, K185N, D189N, and N190S mutated enzymes are reduced to 0.7%, 0.5%, 0.1%, and 0.5%, respectively, in comparison with the wild type enzyme, indicating that Asp-130, Lys-185, Asp-189, and Asn-190 are involved in the catalytic reaction. K(m) values of the mutated enzymes are increased significantly, except for the N190S mutation, suggesting that Asp-130, Lys-185, and Asp-189 participate in the substrate binding. To interpret the kinetic data, the oxidation states of the bound NAD molecules of the wild type and mutated enzymes were measured during the catalytic reaction by monitoring the absorbance at 340 nm. The crystal structures of the WT and D244E.Ado*, containing four subunits in the crystallographic asymmetric unit, were re-refined to have the same subunit structures. A detailed catalytic mechanism of AdoHcyase has been revealed based on the oxidation states of the bound NAD and the re-refined crystal structures of WT and D244E.Ado*. Lys-185 and Asp-130 abstract hydrogen atoms from 3'-OH and 4'-CH, respectively. Asp-189 removes a proton from Lys-185 and produces the neutral N zeta (-NH(2)), and Asn-190 facilitates formation of the neutral Lys-185. His-54 and His-300 hold and polarize a water molecule, which nucleophilically attacks the C5'- of 3'-keto-4',5'-dehydroadenosine to produce 3'-keto-Ado.  相似文献   

16.
A subtilisin-like serine proteinase from a psychrotrophic Vibrio species (VPR) shows distinct cold adapted traits regarding stability and catalytic properties, while sharing high sequence homology with enzymes adapted to higher temperatures. Based on comparisons of sequences and examination of 3D structural models of VPR and related enzymes of higher temperature origin, five sites were chosen to be subject to site directed mutagenesis. Three serine residues were substituted with alanine and two residues in loops were substituted with proline. The single mutations were combined to make double and triple mutants. The single Ser/Ala mutations had a moderately stabilizing effect and concomitantly decreased catalytic efficiency. Introducing a second Ser/Ala mutation did not have additive effect on stability; on the contrary a double Ser/Ala mutant had reduced stability with regard to both wild type and single mutants. The Xaa/Pro mutations stabilized the enzyme and did also tend to decrease the catalytic efficiency more than the Ser/Ala mutations.  相似文献   

17.
Zheng R  Blanchard JS 《Biochemistry》2000,39(51):16244-16251
Ketopantoate reductase (EC 1.1.1.169) catalyzes the NADPH-dependent reduction of alpha-ketopantoate to D-(-)-pantoate in the biosynthesis of pantothenate. The pH dependence of V and V/K for the E. coli enzyme suggests the involvement of a general acid/base in the catalytic mechanism. To identify residues involved in catalysis and substrate binding, we mutated the following six strictly conserved residues to Ala: Lys72, Lys176, Glu210, Glu240, Asp248, and Glu256. Of these, the K176A and E256A mutant enzymes showed 233- and 42-fold decreases in V(max), and 336- and 63-fold increases in the K(m) value of ketopantoate, respectively, while the other mutants exhibited WT kinetic properties. The V(max) for the K176A and E256A mutant enzymes was markedly increased, up to 25% and 75% of the wild-type level, by exogenously added primary amines and formate, respectively. The rescue efficiencies for the K176A and E256A mutant enzymes were dependent on the molecular volume of rescue agents, as anticipated for a finite active site volume. The protonated form of the amine is responsible for recovery of activity, suggesting that Lys176 functions as a general acid in catalysis of ketopantoate reduction. The rescue efficiencies for the K176A mutant by primary amines were independent of the pK(a) value of the rescue agents (Bronsted coefficient, alpha = -0.004 +/-0.008). Insensitivity to acid strength suggests that the chemical reaction is not rate-limiting, consistent with (a) the catalytic efficiency of the wild-type enzyme (k(cat)/K(m) = 2x10(6) M(-1) s(-1) and (b) the small primary deuterium kinetic isotope effects, (D)V = 1.3 and (D)V/K = 1.5, observed for the wild-type enzyme. Larger primary deuterium isotope effects on V and V/K were observed for the K176A mutant ((D)V = 3.0, (D)V/K = 3.7) but decreased nearly to WT values as the concentration of ethylamine was increased. The nearly WT activity of the E256A mutant in the presence of formate argues for an important role for this residue in substrate binding. The double mutant (K176A/E256A) has no detectable ketopantoate reductase activity. These results indicate that Lys176 and Glu256 of the E. coli ketopantoate reductase are active site residues, and we propose specific roles for each in binding ketopantoate and catalysis.  相似文献   

18.
Type I signal peptidase (SPase I) catalyzes the hydrolytic cleavage of the N-terminal signal peptide from translocated preproteins. SPase I belongs to a novel class of Ser proteases that utilize a Ser/Lys dyad catalytic mechanism instead of the classical Ser/His/Asp triad found in most Ser proteases. Recent X-ray crystallographic studies indicate that the backbone amide nitrogen of the catalytic Ser 90 and the hydroxyl side chain of Ser 88 might participate as H-bond donors in the transition-state oxyanion hole. In this work, contribution of the side-chain Ser 88 in Escherichia coli SPase I to the stabilization of the transition state was investigated through in vivo and in vitro characterizations of Ala-, Cys-, and Thr-substituted mutants. The S88T mutant maintains near-wild-type activity with the substrate pro-OmpA nuclease A. In contrast, substitution with Cys at position 88 results in more than a 740-fold reduction in activity (k(cat)) whereas S88A retains much less activity (>2440-fold decrease). Measurements of the kinetic constants of the individual mutant enzymes indicate that these decreases in activity are attributed mainly to decreases in k(cat) while effects on K(m) are minimal. Thermal inactivation and CD spectroscopic analyses indicate no global conformational perturbations of the Ser 88 mutants relative to the wild-type E. coli SPase I enzyme. These results provide strong evidence for the stabilization by Ser 88 of the oxyanion intermediate during catalysis by E. coli SPase I.  相似文献   

19.
L-Ribulose-5-phosphate (L-Ru5P) 4-epimerase and L-fuculose-1-phosphate (L-Fuc1P) aldolase are evolutionarily related enzymes that display 26% sequence identity and a very high degree of structural similarity. They both employ a divalent cation in the formation and stabilization of an enolate during catalysis, and both are able to deprotonate the C-4 hydroxyl group of a phosphoketose substrate. Despite these many similarities, subtle distinctions must be present which allow the enzymes to catalyze two seemingly different reactions and to accommodate substrates differing greatly in the position of the phosphate (C-5 vs C-1). Asp76 of the epimerase corresponds to the key catalytic acid/base residue Glu73 of the aldolase. The D76N mutant of the epimerase retained considerable activity, indicating it is not a key catalytic residue in this enzyme. In addition, the D76E mutant did not show enhanced levels of background aldolase activity. Mutations of residues in the putative phosphate-binding pocket of the epimerase (N28A and K42M) showed dramatically higher values of K(M) for L-Ru5P. This indicates that both enzymes utilize the same phosphate recognition pocket, and since the phosphates are positioned at opposite ends of the respective substrates, the two enzymes must bind their substrates in a reversed or "flipped" orientation. The epimerase mutant D120N displays a 3000-fold decrease in the value of k(cat), suggesting that Asp120' provides a key catalytic acid/base residue in this enzyme. Analysis of the D120N mutant by X-ray crystallography shows that its structure is indistinguishable from that of the wild-type enzyme and that the decrease in activity was not simply due to a structural perturbation of the active site. Previous work [Lee, L. V., Poyner, R. R., Vu, M. V., and Cleland, W. W. (2000) Biochemistry 39, 4821-4830] has indicated that Tyr229' likely provides the other catalytic acid/base residue. Both of these residues are supplied by an adjacent subunit. Modeling of L-Ru5P into the active site of the epimerase structure suggests that Tyr229' is responsible for deprotonating L-Ru5P and Asp120' is responsible for deprotonating its epimer, D-Xu5P.  相似文献   

20.
Li L  Zhang L  Cook PF 《Biochemistry》2006,45(42):12680-12686
Crystal structures of 6-phosphogluconate dehydrogenase (6PGDH) from sheep liver indicate that S128 and N187 are within hydrogen-bonding distance of 6PG in the E:6PG binary complex and NADPH in the E:NADPH binary complex. In addition, H186 is also within hydrogen-bonding distance of NADPH in the E:NADPH binary complex, while in the E:6PG binary complex it is within hydrogen-bonding distance of S128 and close to N187. The structures suggest that this triad of residues may play a dual role during the catalytic reaction. Site-directed mutagenesis has been performed to mutate each of the three residues to alanine. All mutant enzymes exhibit a decrease in V/E(t) (the turnover number), ranging from 7- to 67-fold. An increase in the Km for 6PG (K(6PG)) was observed for S128A and H187A mutant enzymes, while for the H186A mutation, K(6PG) is decreased by a factor of 2. K(NADP) remains the same as the wild type enzyme for the S128A and H186A mutant enzyme, while it increases by 6-fold in the N187A mutant enzyme. An increased K(iNADPH) was measured for all of the mutant enzymes. The primary kinetic 13C-isotope effect is increased, while the primary deuterium kinetic isotope effect is decreased, indicating that the decarboxylation step has become more rate limiting under conditions where substrate is limiting. A quantitative analysis of the data suggests that the S128, H186, and N187 triad is multifunctional in the 6PGDH reaction and contributes as follows. The triad (1) participates in the precatalytic conformational change; (2) provides ground state binding affinity for 6PG and NADPH; and (3) affects the relative rates of reduction or decarboxylation of the 3-keto-6PG intermediate by anchoring the cofactor after hydride transfer, which is accompanied by the rotation of the nicotinamide ring around the N-glycosidic bond and displacement of C1 of 6PG, facilitating decarboxylation.  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