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.
The role of conserved Asp-199 in chloramphenicol acetyltransferase (CAT) has been investigated by site-directed mutagenesis. Substitution of Asp-199 by alanine results in a thermolabile mutant enzyme (Ala-199 CAT) with reduced kcat(13-fold) but similar Km values to wild type CAT. Replacement by asparagine gives rise to a thermostable mutant enzyme (Asn-199 CAT) with much reduced kcat(1500-fold). Furthermore, Asn-199 CAT shows anomalous inactivation kinetics with the affinity reagent 3-(bromo-acetyl)chloramphenicol. These results favor a structural role for Asp-199 rather than a catalytic one, in keeping with crystallographic evidence for involvement of Asp-199 in a tight salt bridge with Arg-18. Replacement of Arg-18 by valine results in a mutant enzyme (Val-18 CAT) with similar properties to Ala-199 CAT. The catalytic imidazole of His-19 appears to be conformationally constrained by hydrogen bonding between N1-H and the carbonyl oxygen of the same residue and by ring stacking with Tyr-25.  相似文献   

2.
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   

3.
The role of Asp-177 in the His-Asp catalytic dyad of glucose 6-phosphate dehydrogenase from Leuconostoc mesenteroides has been investigated by a structural and functional characterization of the D177N mutant enzyme. Its three-dimensional structure has been determined by X-ray cryocrystallography in the presence of NAD(+) and in the presence of glucose 6-phosphate plus NADPH. The structure of a glucose 6-phosphate complex of a mutant (Q365C) with normal enzyme activity has also been determined and substrate binding compared. To understand the effect of Asp-177 on the ionization properties of the catalytic base His-240, the pH dependence of kinetic parameters has been determined for the D177N mutant and compared to that of the wild-type enzyme. The structures give details of glucose 6-phosphate binding and show that replacement of the Asp-177 of the catalytic dyad with asparagine does not affect the overall structure of glucose 6-phosphate dehydrogenase. Additionally, the evidence suggests that the productive tautomer of His-240 in the D177N mutant enzyme is stabilized by a hydrogen bond with Asn-177; hence, the mutation does not affect tautomer stabilization. We conclude, therefore, that the absence of a negatively charged aspartate at 177 accounts for the decrease in catalytic activity at pH 7.8. Structural analysis suggests that the pH dependence of the kinetic parameters of D177N glucose 6-phosphate dehydrogenase results from an ionized water molecule replacing the missing negative charge of the mutated Asp-177 at high pH. Glucose 6-phosphate binding orders and orients His-178 in the D177N-glucose 6-phosphate-NADPH ternary complex and appears to be necessary to form this water-binding site.  相似文献   

4.
Two active site residues, Asp-98 and His-255, of copper-containing nitrite reductase (NIR) from Alcaligenes faecalis have been mutated to probe the catalytic mechanism. Three mutations at these two sites (D98N, H255D, and H255N) result in large reductions in activity relative to native NIR, suggesting that both residues are involved intimately in the reaction mechanism. Crystal structures of these mutants have been determined using data collected to better than 1. 9-A resolution. In the native structure, His-255 Nepsilon2 forms a hydrogen bond through a bridging water molecule to the side chain of Asp-98, which also forms a hydrogen bond to a water or nitrite oxygen ligated to the active site copper. In the D98N mutant, reorientation of the Asn-98 side chain results in the loss of the hydrogen bond to the copper ligand water, consistent with a negatively charged Asp-98 directing the binding and protonation of nitrite in the native enzyme. An additional solvent molecule is situated between residues 255 and the bridging water in the H255N and H255D mutants and likely inhibits nitrite binding. The interaction of His-255 with the bridging water appears to be necessary for catalysis and may donate a proton to reaction intermediates in addition to Asp-98.  相似文献   

5.
In a survey for unknown bioactive peptides in frog (Rana catesbeiana) brain and intestine, we isolated four novel peptides that exhibit potent stimulant effects on smooth muscle preparation of guinea pig ileum. By microsequencing and synthesis, these peptides were identified as Lys- Pro- Ser- Pro- Asp- Arg- Phe- Tyr- Gly- Leu- Met- NH2 (ranatachykinin A), Tyr- Lys- Ser- Asp- Ser- Phe- Tyr- Gly- Leu- Met- NH2 (ranatachykinin B), His- Asn- Pro- Ala- Ser- Phe- Ile- Gly- Leu- Met- NH2 (ranatachykinin C) and Lys- Pro- Ans- Pro- Glu- Arg- Phe- Tyr- Ala- Pro- Met- NH2 (ranatachykinin D). Ranatachykinin (RTK) A, B and C conserve the C- terminal sequence, Phe- X- Gly- Leu- Met- NH2, which is common to known members of the tachykinin family. On the other hand, RTK-D has a striking feature in its C-terminal sequence, Phe- Tyr- Ala- Pro- Met- NH2, which has never been found in other known tachykinins, and may constitute a new subclass in the tachykinin family.  相似文献   

6.
The imidazole of His-195 plays an essential role in the proposed general base mechanism of chloramphenicol acetyltransferase (CAT). The structure of the binary complex of CATIII and chloramphenicol suggests that two unusual interactions might determine the conformation of the side chain of His-195: (i) an intraresidue hydrogen bond between its main chain carbonyl and the protonated N delta 1 of the imidazole ring and (ii) face-to-face van der Waals contact between the His-195 imidazole group and the aromatic side chain of Tyr-25. Tyr-25 also makes a hydrogen bond, via its phenolic hydroxyl, to the carbonyl oxygen of the substrate chloramphenicol. Replacement of Tyr-25 of CATIII by phenylalanine results in a modest increase in the Km for chloramphenicol (from 11.6 to 14.6 microM) and a 2-fold fall in kcat (599 to 258 s-1), indicative of a free energy contribution to transition state binding of 0.6 kcal mol-1 for the hydrogen bond between Tyr-25 and chloramphenicol. In contrast, substitution of Tyr-25 by alanine yields an enzyme that is dramatically impaired in its ability to bind chloramphenicol (Km = 173 microM). As kcat for Ala-25 CAT is also reduced (130 s-1), the loss of the aryl group results in a 69-fold decrease in kcat/Km, corresponding to a free energy contribution to binding and catalysis of 2.5 kcal mol-1. In addition to the loss of the hydrogen bond between Tyr-25 and chloramphenicol, the loss of substrate affinity in Ala-25 CAT may be a direct consequence of reduced hydrophobicity of the chloramphenicol-binding site and/or the loss of critical constraints on the precise conformation of the catalytic imidazole. However, as with wild type CAT, inactivation of Ala-25 CAT by the affinity reagent 3-(bromoacetyl) chloramphenicol is accompanied by modification solely at N epsilon 2 of His-195. Hence, the results demonstrate that tautomeric stabilization of the imidazole ring persists in the absence of van der Waals interactions with the side chain of Tyr-25, probably as a consequence of hydrogen bonding between the protonated N delta 1 and the carbonyl oxygen of His-195.  相似文献   

7.
The enzyme phospholipase A2 (PLA2) catalyzes the hydrolysis of the sn-2 ester bond of membrane phospholipids. The highly conserved Tyr residues 52 and 73 in the enzyme form hydrogen bonds to the carboxylate group of the catalytic Asp-99. These hydrogen bonds were initially regarded as essential for the interfacial recognition and the stability of the overall catalytic network. The elimination of the hydrogen bonds involving the phenolic hydroxyl groups of the Tyr-52 and -73 by changing them to Phe lowered the stability but did not significantly affect the catalytic activity of the enzyme. The X-ray crystal structure of the double mutant Y52F/Y73F has been determined at 1.93 A resolution to study the effect of the mutation on the structure. The crystals are trigonal, space group P3(1)21, with cell parameters a = b = 46.3 A and c = 102.95 A. Intensity data were collected on a Siemens area detector, 8,024 reflections were unique with an R(sym) of 4.5% out of a total of 27,203. The structure was refined using all the unique reflections by XPLOR to a final R-factor of 18.6% for 955 protein atoms, 91 water molecules, and 1 calcium ion. The root mean square deviation for the alpha-carbon atoms between the double mutant and wild type was 0.56 A. The crystal structure revealed that four hydrogen bonds were lost in the catalytic network; three involving the tyrosines and one involving Pro-68. However, the hydrogen bonds of the catalytic triad, His-48, Asp-99, and the catalytic water, are retained. There is no additional solvent molecule at the active site to replace the missing hydroxyl groups; instead, the replacement of the phenolic OH groups by H atoms draws the Phe residues closer to the neighboring residues compared to wild type; Phe-52 moves toward His-48 and Asp-99 of the catalytic diad, and Phe-73 moves toward Met-8, both by about 0.5 A. The closing of the voids left by the OH groups increases the hydrophobic interactions compensating for the lost hydrogen bonds. The conservation of the triad hydrogen bonds and the stabilization of the active site by the increased hydrophobic interactions could explain why the double mutant has activity similar to wild type. The results indicate that the aspartyl carboxylate group of the catalytic triad can function alone without additional support from the hydrogen bonds of the two Tyr residues.  相似文献   

8.
Type II signal peptidases (SPase II) remove signal peptides from lipid-modified preproteins of eubacteria. As the catalytic mechanism employed by type II SPases was not known, the present studies were aimed at the identification of their potential active site residues. Comparison of the deduced amino acid sequences of 19 known type II SPases revealed the presence of five conserved domains. The importance of the 15 best conserved residues in these domains was investigated using the type II SPase of Bacillus subtilis, which, unlike SPase II of Escherichia coli, is not essential for viability. The results showed that only six residues are important for SPase II activity. These are Asp-14, Asn-99, Asp-102, Asn-126, Ala-128, and Asp-129. Only Asp-14 was required for stability of SPase II, indicating that the other five residues are required for catalysis, the active site geometry, or the specific recognition of lipid-modified preproteins. As Asp-102 and Asp-129 are the only residues invoked in the known catalytic mechanisms of proteases, we hypothesize that these two residues are directly involved in SPase II-mediated catalysis. This implies that type II SPases belong to a novel family of aspartic proteases.  相似文献   

9.
NagA catalyzes the hydrolysis of N-acetyl-d-glucosamine-6-phosphate to d-glucosamine-6-phosphate and acetate. X-ray crystal structures of NagA from Escherichia coli were determined to establish the number and ligation scheme for the binding of zinc to the active site and to elucidate the molecular interactions between the protein and substrate. The three-dimensional structures of the apo-NagA, Zn-NagA, and the D273N mutant enzyme in the presence of a tight-binding N-methylhydroxyphosphinyl-d-glucosamine-6-phosphate inhibitor were determined. The structure of the Zn-NagA confirms that this enzyme binds a single divalent cation at the beta-position in the active site via ligation to Glu-131, His-195, and His-216. A water molecule completes the ligation shell, which is also in position to be hydrogen bonded to Asp-273. In the structure of NagA bound to the tight binding inhibitor that mimics the tetrahedral intermediate, the methyl phosphonate moiety has displaced the hydrolytic water molecule and is directly coordinated to the zinc within the active site. The side chain of Asp-273 is positioned to activate the hydrolytic water molecule via general base catalysis and to deliver this proton to the amino group upon cleavage of the amide bond of the substrate. His-143 is positioned to help polarize the carbonyl group of the substrate in conjunction with Lewis acid catalysis by the bound zinc. The inhibitor is bound in the alpha-configuration at the anomeric carbon through a hydrogen bonding interaction of the hydroxyl group at C-1 with the side chain of His-251. The phosphate group of the inhibitor attached to the hydroxyl at C-6 is ion paired with Arg-227 from the adjacent subunit. NagA from Thermotoga maritima was shown to require a single divalent cation for full catalytic activity.  相似文献   

10.
The molecular structure of cytoplasmic malate dehydrogenase from pig heart has been refined by alternating rounds of restrained least-squares methods and model readjustment on an interactive graphics system. The resulting structure contains 333 amino acids in each of the two subunits, 2 NAD molecules, 471 solvent molecules, and 2 large noncovalently bound molecules that are assumed to be sulfate ions. The crystallographic study was done on one entire dimer without symmetry restraints. Analysis of the relative position of the two subunits shows that the dimer does not obey exact 2-fold rotational symmetry; instead, the subunits are related by a 173 degrees rotation. The structure results in a R factor of 16.7% for diffraction data between 6.0 and 2.5 A, and the rms deviations from ideal bond lengths and angles are 0.017 A and 2.57 degrees, respectively. The bound coenzyme in addition to hydrophobic interactions makes numerous hydrogen bonds that either are directly between NAD and the enzyme or are with solvent molecules, some of which in turn are hydrogen bonded to the enzyme. The carboxamide group of NAD is hydrogen bonded to the side chain of Asn-130 and via a water molecule to the backbone nitrogens of Leu-157 and Asp-158 and to the carbonyl oxygen of Leu-154. Asn-130 is one of the corner residues in a beta-turn that contains the lone cis peptide bond in cytoplasmic malate dehydrogenase, situated between Asn-130 and Pro-131. The active site histidine, His-186, is hydrogen bonded from nitrogen ND1 to the carboxylate of Asp-158 and from its nitrogen NE2 to the sulfate ion bound in the putative substrate binding site. In addition to interacting with the active site histidine, this sulfate ion is also hydrogen bonded to the guanidinium group of Arg-161, to the carboxamide group of Asn-140, and to the hydroxyl group of Ser-241. It is speculated that the substrate, malate or oxaloacetate, is bound in the sulfate binding site with the substrate 1-carboxyl hydrogen bonded to the guanidinium group of Arg-161.  相似文献   

11.
We have determined the structures and thermodynamic stabilities of the wild type Asn-52 and unusually thermostable mutant Ile-52 yeast iso-1-cytochromes c (Das, G., Hickey, D. R. McLendon, D., McLendon, G., and Sherman, F. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 496-499). Although both structures were similar, Water-166, buried within the wild type protein, is excluded from the Ile-52 mutant, which substantially reorganizes the local hydrogen bonding. Wild type Cys-102 was replaced with alanine or serine to eliminate dimerization in vitro. The Cys-102 (wild type), Ala-102, and Ser-102 proteins were equally stable, whereas the chemically modified Cys-102-SCH3 was less stable. The order of stability observed with replacements at positions 52 and 102 was as follows: Ile-52 Ala-102 greater than Ala-52 Ala-102 greater than Asn-52 Ala-102 ("normal") greater than Gly-52 Ala-102. No significant stabilization was attributed to potential energy interactions expressed as helix-forming propensities of replacements at position 52. A high correlation between differences in free energy changes and transfer free energies suggests hydrophobic interactions are the main factor for enhancing stability in the Ile-52 mutant. Additional possible contributions to the thermostability of the Ile-52 variant are energetic effects due to packing and hydrogen bonding changes surrounding position 52.  相似文献   

12.
Tryptophan indole-lyase (Trpase) from Proteus vulgaris is a pyridoxal 5'-phosphate dependent enzyme that catalyzes the reversible hydrolytic cleavage of L-Trp to yield indole and ammonium pyruvate. Asp-133 and His-458 are strictly conserved in all sequences of Trpase, and they are located in the proposed substrate-binding region of Trpase. These residues were mutated to alanine to probe their role in substrate binding and catalysis. D133A mutant Trpase has no measurable activity with L-Trp as substrate, but still retains activity with S-(o-nitrophenyl)-L-cysteine, S-alkyl-L-cysteines, and beta-chloro-L-alanine. H458A mutant Trpase has 1.6% of wild-type Trpase activity with L-Trp, and high activity with S-(o-nitrophenyl)-L-cysteine, S-alkyl-L-cysteines, and beta-chloro-L-alanine. H458A mutant Trpase does not exhibit the pK(a) of 5.3 seen in the pH dependence of k(cat)/K(m) of L-Trp for wild-type Trpase. Both mutant enzymes are inhibited by L-Ala, L-Met, and L-Phe, with K(i) values similar to those of wild-type Trpase, but oxindolyl-L-alanine and beta-phenyl-DL-serine show much weaker binding to the mutant enzymes, suggesting that Asp-133 and His-458 are involved in the binding of these ligands. D133A and H458A mutant Trpase exhibit absorption and CD spectra in the presence of substrates and inhibitors that are similar to wild-type Trpase, with peaks at about 420 and 500 nm. The rate constants for formation of the 500 nm bands for the mutant enzymes are equal to or greater than those of wild-type Trpase, indicating that Asp-133 and His-458 do not play a role in the formation of quinonoid intermediates. In constrast to wild-type and H458A mutant Trpase, D133A mutant Trpase forms an intermediate from S-ethyl-L-Cys that absorbs at 345 nm, and is likely to be an alpha-aminoacrylate. Crystals of D133A and H458A mutant Trpase bind amino acids with similar affinity as the proteins in solution, except for L-Ala, which binds to D133A mutant Trpase crystals about 20-fold stronger than in solution. These results suggest that Asp-133 and His-458 play an important role in the elimination reaction of L-Trp. Asp-133 likely forms a hydrogen bond directly to the indole NH of the substrate, while His-458 probably is hydrogen bonded to Asp-133.  相似文献   

13.
Hénot F  Pollack RM 《Biochemistry》2000,39(12):3351-3359
3-oxo-Delta(5)-steroid isomerase (KSI) from Comamonas (Pseudomonas) testosteroni catalyzes the isomerization of beta,gamma-unsaturated 3-oxosteroids to their conjugated isomers through an intermediate dienolate. Residue Asp-38 (pK(a) 4.57) acts as a base to abstract a proton from C-4 of the substrate to form an intermediate dienolate, which is then reprotonated on C-6. Both Tyr-14 (pK(a) 11.6) and Asp-99 (pK(a) >/= 9.5) function as hydrogen-bond donors to O-3 of the steroid, helping to stabilize the transition states. Mutation of the active-site base Asp-38 to the weakly basic Asn (D38N) has previously been shown to result in a >10(8)-fold decrease of catalytic activity. In this work, we describe the preparation and kinetic analysis of the Ala-38 (D38A) mutant. Unexpectedly, D38A has a catalytic turnover number (k(cat)) that is ca. 10(6)-fold greater than the value for D38N and only about 140-fold less than that for wild type. Kinetic studies as a function of pH show that D38A-catalyzed isomerization involves two groups, with pK(a) values of 4.2 and 10.4, respectively, in the free enzyme, which are assigned to Asp-99 and either Tyr-14 or Tyr-55. A mechanism for D38A is proposed in which Asp-99 is recruited as the catalytic base, with stabilization of the intermediate dienolate ion and the flanking transition states provided by hydrogen bonding from both Tyr-14 and Tyr-55. This mechanism is supported by the lack of detectable activity of the D38A/D99N, D38A/Y14F, and D38A/Y55F double mutants.  相似文献   

14.
The structure of the complex of bovine trypsin and bovine pancreatic trypsin inhibitor has been determined by crystal structure analysis at 2.8 Å resolution. The structure is closely similar to the model predicted from the structures of the components. The complex is a tetrahedral adduct with a covalent bond between the carbonyl carbon of Lys-15I of the inhibitor and the γ-oxygen of Ser-195 of the enzyme. The imidazole of His-57 is hydrogen-bonded to Asp-102 and the bound seryl γ-oxygen in accord with the histidine being charged. The negatively charged carbonyl oxygen of Lys-15I forms two hydrogen bonds with the amide nitrogens of Gly-193 and Ser-195. Protonation of the leaving group N-H of Ala-16I to form an acyl-complex requires a conformational change of the imidazole of His-57. The tetrahedral adduct is further stabilized by hydrogen bonds between groups at the leaving group side and inhibitor and enzyme, which would be weakened in the acyl-enzyme. The kinetic data of inhibitor-enzyme interaction are reconciled with the structural model, and relations between enzyme-inhibitor interaction and productive enzyme-substrate interaction are proposed.  相似文献   

15.
Choi G  Ha NC  Kim SW  Kim DH  Park S  Oh BH  Choi KY 《Biochemistry》2000,39(5):903-909
Delta 5-3-ketosteroid isomerase (KSI) catalyzes the allylic isomerization of Delta 5-3-ketosteroids at a rate approaching the diffusion limit by an intramolecular transfer of a proton. Despite the extensive studies on the catalytic mechanism, it still remains controversial whether the catalytic residue Asp-99 donates a hydrogen bond to the steroid or to Tyr-14. To clarify the role of Asp-99 in the catalysis, two single mutants of D99E and D99L and three double mutants of Y14F/D99E, Y14F/D99N, and Y14F/D99L have been prepared by site-directed mutagenesis. The D99E mutant whose side chain at position 99 is longer by an additional methylene group exhibits nearly the same kcat as the wild-type while the D99L mutant exhibits ca. 125-fold lower kcat than that of the wild-type. The mutations made at positions 14 and 99 exert synergistic or partially additive effect on kcat in the double mutants, which is inconsistent with the mechanism based on the hydrogen-bonded catalytic dyad, Asp-99 COOH...Tyr-14 OH...C3-O of the steroid. The crystal structure of D99E/D38N complexed with equilenin, an intermediate analogue, at 1.9 A resolution reveals that the distance between Tyr-14 O eta and Glu-99 O epsilon is ca. 4.2 A, which is beyond the range for a hydrogen bond, and that the distance between Glu-99 O epsilon and C3-O of the steroid is maintained to be ca. 2.4 A, short enough for a hydrogen bond to be formed. Taken together, these results strongly support the idea that Asp-99 contributes to the catalysis by donating a hydrogen bond directly to the intermediate.  相似文献   

16.
The major soluble avian eye lens protein, delta crystallin, is highly homologous to the housekeeping enzyme argininosuccinate lyase (ASL). ASL is part of the urea and arginine-citrulline cycles and catalyzes the reversible breakdown of argininosuccinate to arginine and fumarate. In duck lenses, there are two delta crystallin isoforms that are 94% identical in amino acid sequence. Only the delta2 isoform has maintained ASL activity and has been used to investigate the enzymatic mechanism of ASL. The role of the active site residues Ser-29, Asp-33, Asp-89, Asn-116, Thr-161, His-162, Arg-238, Thr-281, Ser-283, Asn-291, Asp-293, Glu-296, Lys-325, Asp-330, and Lys-331 have been investigated by site-directed mutagenesis, and the structure of the inactive duck delta2 crystallin (ddeltac2) mutant S283A with bound argininosuccinate was determined at 1.96 A resolution. The S283A mutation does not interfere with substrate binding, because the 280's loop (residues 270-290) is in the open conformation and Ala-283 is more than 7 A from the substrate. The substrate is bound in a different conformation to that observed previously indicating a large degree of conformational flexibility in the fumarate moiety when the 280's loop is in the open conformation. The structure of the S283A ddeltac2 mutant and mutagenesis results reveal that a complex network of interactions of both protein residues and water molecules are involved in substrate binding and specificity. Small changes even to residues not involved directly in anchoring the argininosuccinate have a significant effect on catalysis. The results suggest that either His-162 or Thr-161 are responsible for proton abstraction and reinforce the putative role of Ser-283 as the catalytic acid, although we cannot eliminate the possibility that arginine is released in an uncharged form, with the solvent providing the required proton. A detailed enzymatic mechanism of ASL/ddeltac2 is presented.  相似文献   

17.
Makde RD  Mahajan SK  Kumar V 《Biochemistry》2007,46(8):2079-2090
The Salmonella typhimurium PhoN protein is a nonspecific acid phosphatase and belongs to the phosphatidic acid phosphatase type 2 (PAP2) superfamily. We report here the crystal structures of phosphate-bound PhoN, the PhoN-tungstate complex, and the T159D mutant of PhoN along with functional characterization of three mutants: L39T, T159D, and D201N. Invariant active site residues, Lys-123, Arg-130, Ser-156, Gly-157, His-158, and Arg-191, interact with phosphate and tungstate oxyanions. Ser-156 also accepts a hydrogen bond from Thr-159. The T159D mutation, surprisingly, severely diminishes phosphatase activity, apparently by disturbing the active site scaffold: Arg-191 is swung out of the active site resulting in conformational changes in His-158 and His-197 residues. Our results reveal a hitherto unknown functional role of Arg-191, namely, restricting the active conformation of catalytic His-158 and His-197 residues. Consistent with the conserved nature of Asp-201 in the PAP2 superfamily, the D201N mutation completely abolished phosphatase activity. On the basis of this observation and in silico analysis we suggest that the crucial mechanistic role of Asp-201 is to stabilize the positive charge on the phosphohistidine intermediate generated by the transfer of phosphoryl to the nucleophile, His-197, located within hydrogen bond distance to the invariant Asp-201. This is in contrast to earlier suggestions that Asp-201 stabilizes His-197 and the His197-Asp201 dyad facilitates formation of the phosphoenzyme intermediate through a charge-relay system. Finally, the L39T mutation in the conserved polyproline motif (39LPPPP43) of dimeric PhoN leads to a marginal reduction in activity, in contrast to the nearly 50-fold reduction observed for monomeric Prevotella intermedia acid phosphatase, suggesting that the varying quaternary structure of PhoN orthologues may have functional significance.  相似文献   

18.
The 18.1-kDa protein IIIGlc from Escherichia coli acts as both a phosphocarrier protein in the phosphoenolpyruvate:glycose phosphotransferase system (PTS) and as a signal-transducing protein with respect to the uptake of non-PTS sugars. Phosphorylation of IIIGlc at the N epsilon (N3) position of His-90 was effected through a regeneration system that included MgCl2, DTT, excess PEP, and catalytic amounts of Enzyme I and HPr. NH, 15N, and 13C alpha signal assignments for P-IIIGlc were made through comparison of 15N-1H correlation spectra (HSQC) of uniformly 15N-labeled preparations of phosphorylated and unphosphorylated protein and through analysis of three-dimensional triple-resonance HNCA spectra of P-IIIGlc uniformly labeled with both 15N and 13C. Backbone and side-chain 1H and 13C beta signals were assigned using 3D heteronuclear HCCH-COSY and HCCH-TOCSY spectra of P-IIIGlc. Using this approach, the assignments were made without reference to nuclear Overhauser effect data or assumptions regarding protein structure. The majority of NH, 15N, H alpha, and 13C alpha chemical shifts measured for P-IIIGlc were identical to those obtained for the unphosphorylated protein [Pelton, J. G., Torchia, D. A., Meadow, N. D., Wong, C.-Y., & Roseman, S. (1991) Biochemistry 30, 10043]. Those signals that exhibited shifts corresponded to residues within four segments (1) Leu-87-Gly-100, (2) Val-36-Val-46, (3) His-75-Ser-78, and (4) Ala-131-Val-138. These four segments are in close proximity to the active site residues His-75 and His-90 in the unphosphorylated protein [Worthylake, D., Meadow, N. D., Roseman, S., Liao, D., Hertzberg, O., & Remington, S.J. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 10382], and the chemical shift data provide strong evidence that if any structural changes accompany phosphorylation, they are confined to residues in these four segments. This conclusion is confirmed by comparing NOEs observed in 3D 15N/13C NOESY-HMQC spectra of the two forms of the protein. No NOE differences are seen for residues having the same chemical shifts in IIIGlc and P-IIIGlc. Furthermore, with the exception of residues Ala-76, Asp-94, and Val-96, the NOEs of residues (in the four segments) which exhibited chemical shift differences also had the same NOEs in IIIGlc and P-IIIGlc. In the case of residues Ala-76, Asp-94, and Val-96, minor differences in NOEs, corresponding to interproton distances changes of less than 1.5 A, were observed.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Okar DA  Live DH  Devany MH  Lange AJ 《Biochemistry》2000,39(32):9754-9762
The histidines in the bisphosphatase domain of rat liver 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase were labeled with (15)N, both specifically at N1' and globally, for use in heteronuclear single quantum correlation (HSQC) NMR spectroscopic analyses. The histidine-associated (15)N resonances were assigned by correlation to the C2' protons which had been assigned previously [Okar et al., Biochemistry 38, 1999, 4471-79]. Acquisition of the (1)H-(15)N HSQC from a phosphate-free sample demonstrated that the existence of His-258 in the rare N1' tautomeric state is dependent upon occupation of the phosphate binding site filled by the O2 phosphate of the substrate, fructose-2,6-bisphosphate, and subsequently, the phosphohistidine intermediate. The phosphohistidine intermediate is characterized by two hydrogen bonds involving the catalytic histidines, His-258 and His-392, which are directly observed at the N1' positions of the imidazole rings. The N1' of phospho-His-258 is protonated ((1)H chemical shift, 14.0 ppm) and hydrogen bonded to the backbone carbonyl of Gly-259. The N1' of cationic His-392 is hydrogen bonded ((1)H chemical shift, 13.5 ppm) to the phosphoryl moiety of the phosphohistidine. The existence of a protonated phospho-His-258 intermediate and the observation of a fairly strong hydrogen bond to the same phosphohistidine implies that hydrolysis of the covalent intermediate proceeds without any requirement for an "activated" water. Using the labeled histidines as probes of the catalytic site mutation of Glu-327 to alanine revealed that, in addition to its function as the proton donor to fructose-6-phosphate during formation of the transient phosphohistidine intermediate at the N3' of His-258, this residue has a significant role in maintaining the structural integrity of the catalytic site. The (1)H-(15)N HSQC data also provide clear evidence that despite being a surface residue, His-446 has a very acidic pK(a), much less than 6.0. On the basis of these observations a revised mechanism for fructose-2,6-bisphosphatase that is consistent with all of the previously published kinetic data and X-ray crystal structures is proposed. The revised mechanism accounts for the structural and kinetic consequences produced by mutation of the catalytic histidines and Glu-327. It also provides the basis for a hypothetical mechanism of bisphosphatase activation by cAMP-dependent phosphorylation of Ser-32, which is located in the N-terminal kinase domain.  相似文献   

20.
The crotonase homolog, 6-oxo camphor hydrolase (OCH), catalyzes the desymmetrization of bicyclic beta-diketones to optically active keto acids via an enzymatic retro-Claisen reaction, resulting in the cleavage of a carbon-carbon bond. We have previously reported the structure of OCH (Whittingham, J. L., Turkenburg, J. P., Verma, C. S., Walsh, M. A., and Grogan, G. (2003) J. Biol. Chem. 278, 1744-1750), which suggested the involvement of five residues, His-45, His-122, His-145, Asp-154, and Glu-244, in catalysis. Here we report mutation studies on OCH that reveal that H145A and D154N mutants of OCH have greatly reduced values of k(cat)/K(m) derived from a very large increase in K(m) for the native substrate, 6-oxo camphor. In addition, H122A has a greatly reduced value of k(cat), and its K(m) is five times that of the wild-type. The location of the active site is confirmed by the 1.9-A structure of the H122A mutant of OCH complexed with the minor diastereoisomer of (2S,4S)-alpha-campholinic acid, the natural product of the enzyme. This shows the pendant acetate of the product hydrogen bonded to a His-145/Asp-154 dyad and the endocyclic carbonyl of the cyclopentane ring hydrogen bonded to Trp-40. The results are suggestive of a base-catalyzed mechanism of C-C bond cleavage and provide clues to the origin of prochiral selectivity by the enzyme and to the recruitment of the crotonase fold for alternate modes of transition state stabilization to those described for other crotonase superfamily members.  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