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 31 毫秒
1.
All DNA methyltransferases (MTases) have similar catalytic domains containing nine blocks of conserved amino acid residues. We have investigated by site-directed mutagenesis the function of 17 conserved residues in the EcoRV alpha-adenine-N6-DNA methyltransferase. The structure of this class of MTases has been predicted recently. The variants were characterized with respect to their catalytic activities and their abilities to bind to DNA and the S-adenosylmethionine (AdoMet) cofactor. Amino acids located in motifs X, I, and II are shown to be involved in AdoMet binding (Lys16, Glu37, Phe39, and Asp58). Some of the mutants defective in AdoMet binding are also impaired in DNA binding, suggesting allosteric interactions between the AdoMet and DNA binding site. Asp78 (motif III), which was supposed to form a hydrogen bond to the AdoMet on the basis of the structure predictions, turned out not to be important for AdoMet binding, suggesting that motif III has not been identified correctly. R128A and N130A, having mutations in the putative DNA binding domain, are unable to bind to DNA. Residues located in motifs IV, V, VI, and VIII are involved in catalysis (Asp193, Tyr196, Asp211, Ser229, Trp231, and Tyr258), some of them presumably in binding the flipped target base, because mutations at these residues fail to significantly interfere with DNA and AdoMet binding but strongly reduce catalysis. Our results are in substantial agreement with the structure prediction for EcoRV alpha-adenine-N6-methyltransferase and x-ray structures of other MTases.  相似文献   

2.
Bacterial Delta 5-3-ketosteroid isomerase (KSI) from Pseudomonas testosteroni has been intensively studied as a prototype for understanding an enzyme-catalyzed allylic rearrangement involving intramolecular proton transfer. Asp38 serves as a general base to abstract the proton from the steroid C4-H, which is a much stronger base than the carboxyl group of this residue. This unfavorable proton transfer requires 11 kcal/mol of energy which has to be provided by favorable interactions between catalytic residues and substrate in the course of the catalytic reaction. How this energy is provided at the active site of KSI has been a controversial issue, and inevitably the enzyme mechanism is not settled. To resolve these issues, we have determined the crystal structure of this enzyme at 2.3 A resolution. The crystal structure revealed that the active site environment of P. testosteroni KSI is nearly identical to that of Pseudomonas putida KSI, whose structure in complex with a reaction intermediate analogue we have determined recently. Comparison of the two structures clearly indicates that the two KSIs should share the same enzyme mechanism involving the stabilization of the dienolate intermediate by the two direct hydrogen bonds to the dienolate oxyanion, one from Tyr14 OH and the other from Asp99 COOH. Mutational analysis of the two residues and other biochemical data strongly suggest that the hydrogen bond of Tyr14 provides the more significant contribution than that of Asp99 to the requisite 11 kcal/mol of energy for the catalytic power of KSI.  相似文献   

3.
The structures of the sugar kinase/heat shock 70/actin superfamily of enzymes show that the active site is located in a deep cleft between two domains whose relative movement defines a domain closure conformational change thought to be involved in the catalytic and regulatory properties of members of the superfamily. To investigate the role of the domain closure in the regulatory behavior, site-directed mutagenesis is used to alter specific domain-domain interactions in Escherichia coli glycerol kinase (EC 2.7.1.30; ATP:glycerol 3-phosphotransferase), a member of this superfamily. Two active site aspartate residues are conserved throughout the superfamily, one (Asp245 in glycerol kinase) which is proposed to act as a general base during catalysis and one (Asp10 in glycerol kinase) which interacts with the Mg(II) ion of the bound Mg(II)-nucleotide complex. Each of these residues participates in domain-domain interactions that are mediated by the bound substrates. The enzymes containing the substitutions Asp245 to Asn (D245N) or Asp10 to Asn (D10N) were purified by affinity chromatography, and the effects of the substitutions on the catalytic properties and regulation by the allosteric effectors, fructose 1,6-bisphosphate (FBP), and the glucose-specific phosphocarrier protein, IIIGlc (also known as IIAGlc), were determined. Each of the residues participates in catalysis; kcat/Katp is decreased 300-fold by the D245N substitution and 100-fold by the D10N substitution. Affinity labeling with the glycerol analog 1,3-dichloroacetone shows that the level of activity seen for the D245N mutant enzyme is not due to deamidation of the substituted asparagine. Each of the substitutions has little effect on regulation by FBP and the apparent affinity for IIIGlc, and the D245N substitution does not affect the extent of inhibition by IIIGlc. However, the D10N substitution decreases the maximum extent of inhibition by IIIGlc from 100 to 60%, thus changing the action of IIIGlc to that of a partial inhibitor. The different sensitivities of the extents of FBP and IIIGlc inhibition to perturbation of a domain-domain interaction mediated by Asp10 suggest that the relations of the actions of these allosteric effectors to the domain closure conformational change are different.  相似文献   

4.
VanX, one of the five proteins required for the vancomycin-resistant phenotype in clinically pathogenic Enterococci, is a zinc-containing d-Ala-d-Ala dipeptidase. To identify potential zinc ligands and begin defining the active site residues, we have mutated the 2 cysteine, 5 histidine, and 4 of the 28 aspartate and glutamate residues in the 202 residue VanX protein. Of 10 mutations, 3 cause inactivation and greater than 90% loss of zinc in purified enzyme samples, implicating His116, Asp123, and His184 as zinc-coordinating residues. Homology searches using the 10 amino acid sequence SxHxxGxAxD, in which histidine and aspartate residues are putative zinc ligands, identified the metal coordinating ligands in the N-terminal domain of the murine Sonic hedgehog protein, which also exhibits an architecture for metal coordination identical to that observed in thermolysin from Bacillus thermoproteolyticus. Furthermore, this 10 amino acid consensus sequence is found in the Streptomyces albus G zinc-dependent N-acyl-d-Ala-d-Ala carboxypeptidase, an enzyme catalyzing essentially the same d-Ala-d-Ala dipeptide bond cleavage as VanX, suggesting equivalent mechanisms and zinc catalytic site architectures. VanX residue Glu181 is analogous to the Glu143 catalytic base in B. thermoproteolyticus thermolysin, and the E181A VanX mutant has no detectable dipeptidase activity, yet maintains near-stoichiometric zinc content, a result consistent with the participation of the residue as a catalytic base.  相似文献   

5.
The structure of a ternary complex of the catalytic subunit of cAMP-dependent protein kinase, MgATP, and a 20-residue inhibitor peptide was determined at a resolution of 2.7 A using the difference Fourier technique starting from the model of the binary complex (Knighton et al., 1991a). The model of the ternary complex was refined using both X-PLOR and TNT to an R factor of 0.212 and 0.224, respectively. The orientation of the nucleotide and the interactions of MgATP with numerous conserved residues at the active site of the enzyme are clearly defined. The unique protein kinase nucleotide binding site consists of a five-stranded antiparallel beta-sheet with the base buried in a hydrophobic site along beta-strands 1 and 2 and fixed by hydrogen bonds to the N6 amino and N7 nitrogens. The small lobe secures the nucleotide via a glycine-rich loop and by ion pairing with Lys72 and Glu91. While the small lobe fixes the nontransferable alpha- and beta-phosphates in this inhibitor complex, the gamma-phosphate is secured by two Mg2+ ions and interacts both directly and indirectly with several residues in the large lobe--Asp184, Asn171, Lys168. Asp166 is positioned to serve as a catalytic base. The structure is correlated with previous chemical evidence, and the features that distinguish this nucleotide binding motif from other nucleotide binding proteins are delineated.  相似文献   

6.
The role of amino acid residues located in the active site pocket of phosphatidylinositol-specific phospholipase C (PI-PLC) from Bacillus cereus[Heinz, D. W., Ryan, M., Bullock, T., & Griffith, O. H. (1995) EMBO J. 14, 3855-3863] was investigated by site-directed mutagenesis, kinetics, and crystal structure analysis. Twelve residues involved in catalysis and substrate binding (His32, Arg69, His82, Gly83, Lys115, Glu117, Arg163, Trp178, Asp180, Asp198, Tyr200, and Asp274) were individually replaced by 1-3 other amino acids, resulting in a total number of 21 mutants. Replacements in the mutants H32A, H32L, R69A, R69E, R69K, H82A, H82L, E117K, R163I, D198A, D198E, D198S, Y200S, and D274S caused essentially complete inactivation of the enzyme. The remaining mutants (G83S, K115E, R163K, W178Y, D180S, Y200F, and D274N) exhibited reduced activities up to 57% when compared with wild-type PI-PLC. Crystal structures determined at a resolution ranging from 2.0 to 2.7 A for six mutants (H32A, H32L, R163K, D198E, D274N, and D274S) showed that significant changes were confined to the site of the respective mutation without perturbation of the rest of the structure. Only in mutant D198E do the side chains of two neighboring arginine residues move across the inositol binding pocket toward the newly introduced glutamic acid. An analysis of these structure-function relationships provides new insight into the catalytic mechanism, and suggests a molecular explanation of some of the substrate stereospecificity and inhibitor binding data available for this enzyme.  相似文献   

7.
The crystal structure of the hexahistidine-tagged mouse recombinant catalytic subunit (H6-rC) of cAMP-dependent protein kinase (cAPK), complexed with a 20-residue peptide inhibitor from the heat-stable protein kinase inhibitor PKI(5-24) and adenosine, was determined at 2.2 A resolution. Novel crystallization conditions were required to grow the ternary complex crystals. The structure was refined to a final crystallographic R-factor of 18.2% with good stereochemical parameters. The "active" enzyme adopts a "closed" conformation as found in rC:PKI(5-24) [Knighton et al. (1991a,b) Science 253, 407-414, 414-420] and packs in a similar manner with the peptide providing a major contact surface. This structure clearly defines the subsites of the unique nucleotide binding site found in the protein kinase family. The adenosine occupies a mostly hydrophobic pocket at the base of the cleft between the two lobes and is completely buried. The missing triphosphate moiety of ATP is filled with a water molecule (Wtr 415) which replaces the gamma-phosphate of ATP. The glycine-rich loop between beta1 and beta2 helps to anchor the phosphates while the ribose ring is buried beneath beta-strand 2. Another ordered water molecule (Wtr 375) is pentacoordinated with polar atoms from adenosine, Leu 49 in beta-strand 1, Glu 127 in the linker strand between the two lobes, Tyr 330, and a third water molecule, Wtr 359. The conserved nucleotide fold can be defined as a lid comprised of beta-strand 1, the glycine-rich loop, and beta-strand 2. The adenine ring is buried beneath beta-strand 1 and the linker strand (120-127) that joins the small and large lobes. The C-terminal tail containing Tyr 330, a segment that lies outside the conserved core, covers this fold and anchors it in a closed conformation. The main-chain atoms of the flexible glycine-rich loop (residues 50-55) in the ATP binding domain have a mean B-factor of 41.4 A2. This loop is quite mobile, in striking contrast to the other conserved loops that converge at the active site cleft. The catalytic loop (residues 166-171) and the Mg2+ positioning loop (residues 184-186) are a stable part of the large lobe and have low B-factors in all structures solved to date. The stability of the glycine-rich loop is highly dependent on the ligands that occupy the active site cleft with maximum stability achieved in the ternary complex containing Mg x ATP and the peptide inhibitor. In this ternary complex the gamma-phosphate is secured between both lobes by hydrogen bonds to the backbone amide of Ser 53 in the glycine-rich loop and the amino group of Lys 168 in the catalytic loop. In the adenosine ternary complex the water molecule replacing the gamma-phosphate hydrogen bonds between Lys 168 and Asp 166 and makes no contact with the small lobe. This glycine-rich loop is thus the most mobile component of the active site cleft, with the tip of the loop being highly sensitive to what occupies the gamma-subsite.  相似文献   

8.
The crystal structure of phosphoenolpyruvate carboxylase (PEPC; EC 4. 1.1.31) has been determined by x-ray diffraction methods at 2.8-A resolution by using Escherichia coli PEPC complexed with L-aspartate, an allosteric inhibitor of all known PEPCs. The four subunits are arranged in a "dimer-of-dimers" form with respect to subunit contact, resulting in an overall square arrangement. The contents of alpha-helices and beta-strands are 65% and 5%, respectively. All of the eight beta-strands, which are widely dispersed in the primary structure, participate in the formation of a single beta-barrel. Replacement of a conserved Arg residue (Arg-438) in this linkage with Cys increased the tendency of the enzyme to dissociate into dimers. The location of the catalytic site is likely to be near the C-terminal side of the beta-barrel. The binding site for L-aspartate is located about 20 A away from the catalytic site, and four residues (Lys-773, Arg-832, Arg-587, and Asn-881) are involved in effector binding. The participation of Arg-587 is unexpected, because it is known to be catalytically essential. Because this residue is in a highly conserved glycine-rich loop, which is characteristic of PEPC, L-aspartate seemingly causes inhibition by removing this glycine-rich loop from the catalytic site. There is another mobile loop from Lys-702 to Gly-708 that is missing in the crystal structure. The importance of this loop in catalytic activity was also shown. Thus, the crystal-structure determination of PEPC revealed two mobile loops bearing the enzymatic functions and accompanying allosteric inhibition by L-aspartate.  相似文献   

9.
We solved the structure and traced the complete active site of the catalytic domain of the human immunodeficiency virus type 1 integrase (HIV-1 IN) with the F185H mutation. The only previously available crystal structure, the F185K mutant of this domain, lacks one of the catalytically important residues, E152, located in a stretch of 12 disordered residues [Dyda et al. (1994) Science 266, 1981-1986]. It is clear, however, that the active site of HIV-1 IN observed in either structure cannot correspond to that of the functional enzyme, since the cluster of three conserved carboxylic acids does not create a proper metal-binding site. The conformation of the loop was compared with two different conformations found in the catalytic domain of the related avian sarcoma virus integrase [Bujacz et al. (1995) J. Mol. Biol. 253, 333-346]. Flexibility of the active site region of integrases may be required in order for the enzyme to assume a functional conformation in the presence of substrate and/or cofactors.  相似文献   

10.
Endo-beta-N-acetylglucosaminidase F1 (Endo F1) is an endoglycosidase, secreted by Flavobacterium meningosepticum, that cleaves asparagine-linked oligosaccharides after the first N-acetylglucosamine residue. The enzyme is selective for high-mannose oligosaccharide chains. The crystal structure of Endo F1 has been determined at 2.0-A resolution. The molecular fold consists of a highly irregular alpha/beta-barrel, a commonly observed motif consisting of a cyclic 8-fold repeat of beta-strand/loop/alpha-helix units with an eight-stranded parallel beta-barrel at the center. Endo F1 lacks two of the alpha-helices, those of units 5 and 6. Instead, the links after beta-strands 5 and 6 consist of a short turn followed by a section in an extended conformation that replaces the helix and a long loop at the bottom of the molecule. The absence of any excursion on top of the molecule following beta-strands 5 and 6 results in a pronounced depression in the rim of the barrel. This depression forms one end of a shallow cleft that runs across the surface of the molecule, over the core of the beta-barrel to the area between the loops of units 1 and 2. The active site residues, Asp130 and Glu132, are located at the carboxyl end of beta-strand 4 and extend into this cleft. These residues are surrounded by several tyrosine residues. The cleft area formed by loops 1 and 2 is lined with polar residues, mainly asparagines. The latter area is thought to be responsible for oligosaccharide binding and recognition while the protein moiety of the substrate would be located outside the molecule but adjacent to the area of loops 5 and 6.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Enzymic catalysts of thiol:disulfide oxidoreduction contain two cysteine residues in their active sites. Another common residue is an aspartate (or glutamate), the role of which has been unclear. Escherichia coli thioredoxin (Trx) is the best characterized thiol:disulfide oxidoreductase, and in Trx these three active-site residues are Cys32, Cys35, and Asp26. Structural analyses had indicated that the carboxylate of Asp26 is positioned properly for the deprotonation of the thiol of Cys35, which would facilitate its attack on Cys32 in enzyme-substrate mixed disulfides. Here, Asp26 of Trx was replaced with isologous asparagine and leucine residues. D26N Trx and D26L Trx are reduced and oxidized more slowly than is wild-type Trx during catalysis by E.coli thioredoxin reductase. Stopped-flow spectroscopy demonstrated that the cleavage of the mixed disulfide between Trx and a substrate is slower in the D26N and D26L enzymes. Buffers increase the rate of mixed disulfide cleavage in these variants but not in wild-type Trx. These results indicate that Asp26 serves as an acid/base in the oxidation/reduction reactions catalyzed by Trx. Specifically, Asp26 protonates (during substrate oxidation) or deprotonates (during substrate reduction) the thiol of Cys35. A similar role is likely filled by the analogous aspartate (or glutamate) residue in protein disulfide isomerase, DsbA, and other thiol:disulfide oxidoreductases. Moreover, these results provide the first evidence for general acid/base catalysis in a thiol:disulfide interchange reaction.  相似文献   

12.
The known mammalian 3':5'-cyclic nucleotide phosphodiesterases (PDEs) contain a conserved region located toward the carboxyl terminus, which constitutes a catalytic domain. To identify amino acids that are important for catalysis, we introduced substitutions at 23 conserved residues within the catalytic domain of the cGMP-binding cGMP-specific phosphodiesterase (cGB-PDE; PDE5). Wild-type and mutant proteins were compared with respect to Km for cGMP, kcat, and IC50 for zaprinast. The most dramatic decrease in kcat was seen with H643A and D754A mutants with the decrease in free energy of binding (DeltaDeltaGT) being about 4.5 kcal/mol for each, which is within the range predicted for loss of a hydrogen bond involving a charged residue. His643 and Asp754 are conserved in all known PDEs and are strong candidates to be directly involved in catalysis. Substitutions of His603, His607, His647, Glu672, and Asp714 also produced marked changes in kcat, and these residues are likely to be important for efficient catalysis. The Y602A and E775A mutants exhibited the most dramatic increases in Km for cGMP, with calculated DeltaDeltaGT of 2.9 and 2.8 kcal/mol, respectively, that these two residues are important for cGMP binding in the catalytic site. Zaprinast is a potent competitive inhibitor of cGB-PDE, but the key residues for its binding differ significantly from those that bind cGMP.  相似文献   

13.
The carbohydrate-binding cleft of Bacillus licheniformis 1,3-1, 4-beta-D-glucan 4-glucanohydrolase is partially covered by the surface loop between residues 51 and 67, which is linked to beta-strand-(87-95) of the minor beta-sheet III of the protein core by a single disulfide bond at Cys61-Cys90. An alanine scanning mutagenesis approach has been applied to analyze the role of loop residues from Asp51 to Arg64 in substrate binding and stability by means of equilibrium urea denaturation, enzyme thermotolerance, and kinetics. The DeltaDeltaGU between oxidized and reduced forms is approximately constant for all mutants, with a contribution of 5.3 +/- 0.2 kcal.mol-1 for the disulfide bridge to protein stability. A good correlation is observed between DeltaGU values by reversible unfolding and enzyme thermotolerance. The N57A mutant, however, is more thermotolerant than the wild-type enzyme, whereas it is slightly less stable to reversible urea denaturation. Mutants with a <2-fold increase in Km correspond to mutations at residues not involved in substrate binding, for which the reduction in catalytic efficiency (kcat/Km) is proportional to the loss of stability relative to the wild-type enzyme. Y53A, N55A, F59A, and W63A, on the other hand, show a pronounced effect on catalytic efficiency, with Km > 2-fold and kcat < 5% of the wild-type values. These mutated residues are directly involved in substrate binding or in hydrophobic packing of the loop. Interestingly, the mutation M58A yields an enzyme that is more active than the wild-type enzyme (7-fold increase in kcat), but it is slightly less stable.  相似文献   

14.
The intrinsic fluorescence of homogeneous castor oil seed cytosolic fructose-1,6-bisphosphatase (FBPasec) was used as an indicator of conformational changes due to ligand binding. Binding of the substrate and the inhibitor fructose-2,6-bisphosphate (F-2,6-P2) was quantitatively compared to their respective kinetic effects on enzymatic activity. There are two distinct types of substrate interaction with FBPasec, corresponding to catalytic and inhibitory binding, respectively. Inhibitory substrate binding shares several characteristics with F-2,6-P2 binding which indicates that both ligands bind at the same site. However, F-2,6-P2 does not prevent fluorescence transitions attributed to catalytic substrate binding. The marked synergistic inhibition of FBPasec by AMP and F-2,6-P2 appears to arise via AMP's promotion of F-2,6-P2 binding. Based on the X-ray crystal structure of porcine kidney FBPase our modelling studies suggest the existence of a distinct F-1,6-P2/F-2,6-P2 inhibitory binding site which partially overlaps with the enzyme's catalytic site. We propose that a pronounced allosteric transition mediated by AMP binding increases access of F-1,6-P2 and F-2,6-P2 to this common inhibitory binding site.  相似文献   

15.
The catalytic triad consisting of His57, Asp102 and Ser195, which is completely conserved within the chymotrypsin-like serine protease family, plays a central role in catalysis. Highly conserved Ala55 also likely plays an important role in catalysis due to its location just behind the catalytic triad. The only exception to the conserved Ala55 in mammalian serine proteases is Val55 in bovine protein C. Interestingly, it has been demonstrated that the replacement of Ala55 with Thr results in the reduced activity of plasmin in patients with venous thrombosis and with retinochoroidal vascular disorders, which indicates the importance of Ala55 in catalysis. In the present study, we constructed a bovine protein C model which shows that Val55 causes no serious rearrangement of the catalytic site structure. We also constructed an A55T variant model of trypsin for comparison. The A55T substitution alters His57 into an inactive conformation, forming an unusual hydrogen bond between Thr55 O gamma 1 and His57 N epsilon 2. The present study shows that the Ala/Val55 residue contributes heavily to the active conformation of His57 and enables His57 to accept a proton from Ser195 O gamma effectively.  相似文献   

16.
Factor D is a serine protease essential for the activation of the alternative pathway of complement. The structures of native factor D and a complex formed with isatoic anhydride inhibitor were determined at resolution of 2.3 and 1.5 A, respectively, in an isomorphous monoclinic crystal form containing one molecule per asymmetric unit. The native structure was compared with structures determined previously in a triclinic cell containing two molecules with different active site conformations. The current structure shows greater similarity with molecule B in the triclinic cell, suggesting that this may be the dominant factor D conformation in solution. The major conformational differences with molecule A in the triclinic cell are located in four regions, three of which are close to the active site and include some of the residues shown to be critical for factor D catalytic activity. The conformational flexibility associated with these regions is proposed to provide a structural basis for the previously proposed substrate-induced reversible conformational changes in factor D. The high-resolution structure of the factor D/isatoic anhydride complex reveals the binding mode of the mechanism-based inhibitor. The higher specificity towards factor D over trypsin and thrombin is based on hydrophobic interactions between the inhibitor benzyl ring and the aliphatic side-chain of Arg218 that is salt bridged with Asp189 at the bottom of the primary specificity (S1) pocket. Comparison of factor D structural variants with other serine protease structures revealed the presence of a unique "self-inhibitory loop". This loop (214-218) dictates the resting-state conformation of factor D by (1) preventing His57 from adopting active tautomer conformation, (2) preventing the P1 to P3 residues of the substrate from forming anti-parallel beta-sheets with the non-specific substrate binding loop, and (3) blocking the accessibility of Asp189 to the positive1y charged P1 residue of the substrate. The conformational switch from resting-state to active-state can only be induced by the single macromolecular substrate, C3b-bound factor B. This self-inhibitory mechanism is highly correlated with the unique functional properties of factor D, which include high specificity toward factor B, low esterolytic activity toward synthetic substrates, and absence of regulation by zymogen and serpin-like or other natural inhibitors in blood.  相似文献   

17.
Bacterial Delta5-3-ketosteroid isomerase (KSI) catalyzes a stereospecific isomerization of steroid substrates at an extremely fast rate, overcoming a large disparity of pKa values between a catalytic residue and its target. The crystal structures of KSI from Pseudomonas putida and of the enzyme in complex with equilenin, an analogue of the reaction intermediate, have been determined at 1.9 and 2.5 A resolution, respectively. The structures reveal that the side chains of Tyr14 and Asp99 (a newly identified catalytic residue) form hydrogen bonds directly with the oxyanion of the bound inhibitor in a completely apolar milieu of the active site. No water molecule is found at the active site, and the access of bulk solvent is blocked by a layer of apolar residues. Asp99 is surrounded by six apolar residues, and consequently, its pKa appears to be elevated as high as 9.5 to be consistent with early studies. No interaction was found between the bound inhibitor and the residue 101 (phenylalanine in Pseudomonas testosteroni and methionine in P. putida KSI) which was suggested to contribute significantly to the rate enhancement based on mutational analysis. This observation excludes the residue 101 as a potential catalytic residue and requires that the rate enhancement should be explained solely by Tyr14 and Asp99. Kinetic analyses of Y14F and D99L mutant enzymes demonstrate that Tyr14 contributes much more significantly to the rate enhancement than Asp99. Previous studies and the structural analysis strongly suggest that the low-barrier hydrogen bond of Tyr14 (>7.1 kcal/mol), along with a moderate strength hydrogen bond of Asp99 ( approximately 4 kcal/mol), accounts for the required energy of 11 kcal/mol for the transition-state stabilization.  相似文献   

18.
We report on the role of HLA-DQA1 and DQB1 alleles in determining susceptibility to insulin-dependent diabetes mellitus (IDDM) in Hong Kong Chinese and investigate whether these alleles affect the age of onset of the disease. We studied 76 unrelated Chinese patients and 250 controls. There was no apparent predisposing effect of non-aspartic acid residues at position 57 of the DQ beta chain (Asp57-) but there was an excess of homozygous genotypes containing arginine at position 52 of the DQ alpha chain (Arg52+). This excess was mainly attributable to the genotype DQA1*0301/DQA1*05011 in early-onset disease. There was a significant excess of heterodimers of DQ alpha and DQ beta carrying Arg52+ and Asp57- in both early-onset and late-onset disease, but the excess in early-onset disease was mainly attributable to a single heterodimer formed by DQA1*05011 and DQB1*0201. Of three DQA1/DQB1 genotypes containing a double dose of Arg52+ and Asp57-, only one had a strong association with both early-onset and late-onset disease. We show that early-onset IDDM and late-onset IDDM in Chinese may be separated on the basis of their associated DQA1 and DQB1 genotypes and we conclude that previously reported associations of IDDM with Arg52+ and Asp57- residues in Chinese are secondary to specific combinations of DQA1 and DQB1 alleles. We also show that DRB1 molecules play a distinct role in determining susceptibility to early-onset IDDM but the greatest effect is exerted by specific DR/DQ genotypic combinations.  相似文献   

19.
The conserved glycines in the glycine-rich loop (Leu-Gly50-Thr-Gly52-Ser-Phe-Gly55-Arg-Val) of the catalytic (C) subunit of cAMP-dependent protein kinase were each mutated to Ser (G50S, G52S, and G55S). The effects of these mutations were assessed here using both steady-state and pre-steady-state kinetic methods. While G50S and G52S reduced the apparent affinity for ATP by approximately 10-fold, substitution at Gly55 had no effect on nucleotide binding. In contrast to ATP, only mutation at position 50 interfered with ADP binding. These three mutations lowered the rate of phosphoryl transfer by 7-300-fold. The combined data indicate that G50 and G52 are the most critical residues in the loop for catalysis, with replacement at position 52 being the most extreme owing to a larger decrease in the rate of phosphoryl transfer (29 vs 1.6 s-1 in contrast to 500 s-1 for wild-type C). Surprisingly, all three mutations lowered the affinity for Kemptide by approximately 10-fold, although none of the loop glycines makes direct contact with the substrate. The inability to correlate the rate constant for net product release with the dissociation constant for ADP implies that other steps may limit the decomposition of the ternary product complex. The observations that G52S (a) selectively affects ATP binding and (b) significantly lowers the rate of phosphoryl transfer without making direct contact with either the nucleotide or the peptide imply that this residue serves a structural role in the loop, most likely by positioning the backbone amide of Ser53 for contacting the gamma-phosphate of ATP. Energy-minimized models of the mutant proteins are consistent with the observed kinetic consequences of each mutation. The models predict that only mutation of Gly52 will interfere with the observed hydrogen bonding between the backbone and ATP.  相似文献   

20.
Glu-198 of human matrilysin is a conserved residue in the matrix metalloproteinases and is considered to play an important role in catalysis by acting as a general base catalyst toward the zinc-bound water molecule, on the basis of mechanistic proposals for other zinc proteases. In the present study, Glu-198 is mutated into Asp, Cys, Gln, and Ala, and the zinc binding properties, kinetic parameters, and pH dependence of each mutant are determined in order to examine the role of Glu-198 in catalysis. The mutations chosen either modify (C and D) or eliminate (A and Q) the general base properties of residue-198. All the mutants bind 2 mol of zinc per mol of enzyme, indicating that Glu-198 is not crucial to the binding of the catalytic zinc to the enzyme. The value of kcat/Km for the E198D mutant is only 4-fold lower than that of wild-type enzyme at the pH optimum of 7.5, while that for the E198C mutant is decreased by 160-fold. The E198Q and E198A enzymes containing the mutations that have eliminated the nucleophilic and acid/base properties of the residue are still active, having lower kcat/Km values of 590- and 1900-fold, respectively. The decrease in activity of all the mutants is essentially due to a decrease in kcat. The kcat/Km values of the mutants as a function of pH display broad bell-shaped curves that are similar to the wild-type enzyme. The acidic pKa value is not greatly affected by the change in the chemical properties of residue-198. The similarity in the pH profiles for the mutant and wild-type enzymes indicates that the ionization of Glu-198 is not responsible for the acidic pKa. Ionization of the zinc-bound water may be responsible for this pKa since the three His ligands and the scaffolding of the matrilysin catalytic zinc site are different from that observed in carboxypeptidase A and would predict a lower pKa for the metal-bound water. If the zinc-bound water is the nucleophile in the reaction, the role of Glu-198 in catalysis may be to stabilize the transition state or act as a general acid catalyst after the rate-determining step.  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