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 human genetic variant carbonic anhydrase I (CA I) Michigan 1 results from a single point mutation that changes His 67 to Arg in a critical region of the active site. This variant of the zinc metalloenzyme appears to be unique in that it possesses an esterase activity that is specifically enhanced by added free zinc ions. We have determined the three-dimensional structure of human CA I Michigan 1 by X-ray crystallography to a resolution of 2.6 A. In the absence of added zinc ions, the mutated residue, Arg 67, points out of the active site, hydrogen bonding with the carboxylate of Asn 69. This contrasts with the orientation of His 67, in the native isozyme, which points into the active site. The orientations of His 94, His 96, and His 119, that coordinate the catalytic zinc ion, and of the catalytically critical Thr 199-Glu 106 hydrogen bonding system, are largely unchanged in the mutant. The structure of an enzyme adduct with a second zinc bound was determined to a resolution of 2.0 A. The second zinc ion is coordinated to His 64, His 200, and Arg 67. This arginine residue reverses its orientation on zinc binding and turns into the active site. The residues at these three positions have been implicated in determining the specific kinetic properties of native CA I. This is, to our knowledge, the first example of a zinc ion coordinating with an arginine residue in a Zn(II) enzyme.  相似文献   

2.
Human carbonic anhydrase II (HCA II) is a zinc-metalloenzyme that catalyzes the reversible interconversion of CO2 and HCO3-. The rate-limiting step of this catalysis is the transfer of a proton between the Zn-bound solvent molecule and residue His64. In order to fully characterize the active site structural features implicated in the proton transfer mechanism, the refined X-ray crystal structure of uncomplexed wild type HCA II to 1.05 A resolution with an Rcryst value of 12.0% and an Rfree value of 15.1% has been elucidated. This structure provides strong clues as to the pathway of the intramolecular proton transfer between the Zn-bound solvent and His64. The structure emphasizes the role of the solvent network, the unique positioning of solvent molecule W2, and the significance of the dual conformation of His64 in the active site. The structure is compared with molecular dynamics (MD) simulation calculations of the Zn-bound hydroxyl/His64+ (charged) and the Zn-bound water/His64 (uncharged) HCA II states. A comparison of the crystallographic anisotropic atomic thermal parameters and MD simulation root-mean-square fluctuation values show excellent agreement in the atomic motion observed between the two methods. It is also interesting that the observed active site solvent positions in the crystal structure are also the most probable positions of the solvent during the MD simulations. On the basis of the comparative study of the MD simulation results, the HCA II crystal structure observed is most likely in the Zn-bound water/His64 state. This conclusion is based on the following observations: His64 is mainly (80%) orientated in an inward conformation; electron density omit maps infer that His64 is not charged in an either inward or outward conformation; and the Zn-bound solvent is most likely a water molecule.  相似文献   

3.
Zheng J  Avvaru BS  Tu C  McKenna R  Silverman DN 《Biochemistry》2008,47(46):12028-12036
Catalysis by the zinc metalloenzyme human carbonic anhydrase II (HCA II) is limited in maximal velocity by proton transfer between His64 and the zinc-bound solvent molecule. Asn62 extends into the active site cavity of HCA II adjacent to His64 and has been shown to be one of several hydrophilic residues participating in a hydrogen-bonded solvent network within the active site. We compared several site-specific mutants of HCA II with replacements at position 62 (Ala, Val, Leu, Thr, and Asp). The efficiency of catalysis in the hydration of CO 2 for the resulting mutants has been characterized by (18)O exchange, and the structures of the mutants have been determined by X-ray crystallography to 1.5-1.7 A resolution. Each of these mutants maintained the ordered water structure observed by X-ray crystallography in the active site cavity of wild-type HCA II; hence, this water structure was not a variable in comparing with wild type the activities of mutants at residue 62. Crystal structures of wild-type and N62T HCA II showed both an inward and outward orientation of the side chain of His64; however, other mutants in this study showed predominantly inward (N62A, N62V, N62L) or predominantly outward (N62D) orientations of His64. A significant role of Asn62 in HCA II is to permit two conformations of the side chain of His64, the inward and outward, that contributes to maximal efficiency of proton transfer between the active site and solution. The site-specific mutant N62D had a mainly outward orientation of His64, yet the difference in p K a between the proton donor His64 and zinc-bound hydroxide was near zero, as in wild-type HCA II. The rate of proton transfer in catalysis by N62D HCA II was 5% that of wild type, showing that His64 mainly in the outward orientation is associated with inefficient proton transfer compared with His64 in wild type which shows both inward and outward orientations. These results emphasize the roles of the residues of the hydrophilic side of the active site cavity in maintaining efficient catalysis by carbonic anhydrase.  相似文献   

4.
The tryptophan residue Trp5, highly conserved in the α class of carbonic anhydrases including human carbonic anhydrase II (HCA II), is positioned at the entrance of the active site cavity and forms a π-stacking interaction with the imidazole ring of the proton shuttle His64 in its outward orientation. We have observed that replacement of Trp5 in HCA II caused significant structural changes, as determined by X-ray diffraction, in the conformation of 11 residues at the N-terminus and in the orientation of the proton shuttle residue His64. Most significantly, two variants W5H and W5E HCA II had His64 predominantly outward in orientation, while W5F and wild type showed the superposition of both outward and inward orientations in crystal structures. Although Trp5 influences the orientation of the proton shuttle His64, this orientation had no significant effect on the rate constant for proton transfer near 1 μs−1, determined by exchange of 18O between CO2 and water measured by mass spectrometry. The apparent values of the pKa of the zinc-bound water and the proton shuttle residue suggest that different active-site conformations influence the two stages of catalysis, the proton transfer stage and the interconversion of CO2 and bicarbonate.  相似文献   

5.
The X-ray crystallographic structure for the adduct of an activator with human carbonic anhydrase isozyme I (hCA I) is reported. L-Histidine binds deep within the enzyme active site, participating in a network of hydrogen bonds involving its carboxylate moiety and the zinc-bound water molecule, as well as the imidazole of His200, being in van der Waals contacts with Thr199, His200, His64, and His67. This binding is very different from that to the other major cytosolic isozyme hCA II.  相似文献   

6.
Acipimox, a nicotinic acid derivative in clinical use for the treatment of hyperlipidaemia, incorporates a free carboxylic acid and an N-oxide moiety, functionalities known to interact with the metalloenzyme carbonic anhydrase (CA, EC 4.2.1.1) and inhibit its activity. Herein we report that acipimox acts as a low micromolar CA inhibitor (CAI) against most human (h) isoforms possessing catalytic activity, hCA I – XIV. By using computational techniques (docking and molecular dynamics simulations), we propose that acipimox coordinates through its carboxylate group to the zinc ion from the enzyme active site cavity, whereas the N-oxide group is hydrogen-bonded to the proton shuttle His residue in some isoforms (hCA I) or to active site Thr or Gln residues in other isoforms (hCA II, III, IV, VII, etc). As some CA isoforms are involved in lipogenesis, these data may be useful for the design of more effective CAIs with antiobesity activity.  相似文献   

7.
Activation of the carbonic anhydrase (CA, EC 4.2.1.1) isoforms hCA I, II, and IV with l-histidine and some of its derivatives has been investigated by kinetic and X-ray crystallographic methods. l-His was a potent activator of isozymes I and IV (activation constants in the range of 4-33microM), and a moderate hCA II activator (activation constant of 113microM). Both carboxy- as well as amino-substituted l-His derivatives, such as the methyl ester or the dipeptide carnosine (beta-Ala-His), acted as more efficient activators as compared to l-His. The X-ray crystallographic structure of the hCA II-l-His adduct showed the activator to be anchored at the entrance of the active site cavity, participating in an extended network of hydrogen bonds with the amino acid residues His64, Asn67, and Gln92 and, with three water molecules connecting it to the zinc-bound water. Although the binding site of l-His is similar to that of histamine, the first CA activator for which the X-ray crystal structure has been reported in complex with hCA II (Briganti, F.; Mangani, S.; Orioli, P.; Scozzafava, A.; Vernaglione, G.; Supuran, C. T. Biochemistry1997, 36, 10384) there are important differences of binding between the two structurally related activators, since histamine interacts among others with Asn67 and Gln92 (similarly to l-His), but also with Asn62 and not His64, whereas the number of water molecules connecting them to the zinc-bound water is different (two for histamine, three for l-His). Furthermore, the imidazole moieties of the two activators adopt different conformations when bound to the enzyme active site. Since neither the amino- nor carboxy moieties of l-His participate in interactions with amino acid moieties of the active site, they can be derivatized for obtaining more potent activators, with pharmacological applications for the enhancement of synaptic efficacy. This may constitute a novel approach for the treatment of Alzheimer's disease, aging, and other conditions in need of achieving spatial learning and memory therapy.  相似文献   

8.
The maximal velocity of catalysis of CO(2) hydration by human carbonic anhydrase II (HCA II) requires proton transfer from zinc-bound water to solution assisted by His 64. The catalytic activity of a site-specific mutant of HCA II in which His 64 is replaced with Ala (H64A HCA II) can be rescued by exogenous proton donors/acceptors, usually derivatives of imidazole and pyridine. X-ray crystallography has identified Trp 5 as a binding site of the rescue agent 4-methylimidazole (4-MI) on H64A HCA II. This binding site overlaps with the "out" position in which His 64 in wild-type HCA II points away from the zinc. Activation by 4-MI as proton donor/acceptor in catalysis was determined in the dehydration direction using (18)O exchange between CO(2) and water and in the hydration direction by stopped-flow spectrophotometry. Replacement of Trp 5 by Ala, Leu, or Phe in H64A HCA II had no significant effect on enhancement by 4-MI of maximal rate constants for proton transfer in catalysis to levels near 10(5) s(-1). This high activity for chemical rescue indicates that the binding site of 4-MI at Trp 5 in H64A HCA II appears to be a nonproductive binding site, although it is possible that a similarly effective pathway for proton transfer exists in the mutants lacking Trp 5. Moreover, the data suggest that the out position of His 64 considered alone is not active in proton transfer in HCA II. In contrast to isozyme II, the replacement of Trp 5 by Ala in HCA III abolished chemical rescue of k(cat) by imidazole but left k(cat)/K(m) for hydration unchanged. This demonstrates that Trp 5 contributes to the predominant productive binding site for imidazole, with a maximal level for the rate constant of proton transfer near 10(4) s(-1). This difference in the susceptibility of CA II and III to chemical rescue may be related to the more sterically constrained and electrostatically positive nature of the active site cavity of CA III compared with CA II. The possibility of nonproductive binding sites for exogenous proton donors offers an explanation for the unusually low value of the intrinsic kinetic barrier obtained by application of Marcus theory to chemical rescue of H64A HCA II.  相似文献   

9.
A manual threading approach is used to model the human glutathione transferase T1–1 based on the coordinates of the related Theta class enzyme T2–2. The low level of sequence identity (about 20%), found in the C-terminal extension in conjunction with a relative deletion of about five residues makes this a challenging modeling problem. The C-terminal extension contributes to the active site of the molecule and is thus of particular interest for understanding the molecular mechanism of the enzyme. Manual docking of known substrates and non-substrates has implicated potential candidates for the T1–1 catalytic residues involved in the dehalogenation and epoxide-ring opening activities. These include the conserved Theta class residues Arg 107, Trp 115, and the conserved GSTT1 subclass residue His 176. Also, the residue at position 234 is implicated in the modulation of T1–1 activity with different substrates between species. Proteins 33:444–454, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
Histidyl-tRNA synthetase (HisRS) differs from other class II aminoacyl-tRNA synthetases (aaRS) in that it harbors an arginine at a position where the others bind a catalytic Mg2+ ion. In computer experiments, four mutants of HisRS from Escherichia coli were engineered by removing the arginine and introducing a Mg2+ ion and residues from seryl-tRNA synthetase (SerRS) that are involved in Mg2+ binding. The mutants recreate an active site carboxylate pair conserved in other class II aaRSs, in two possible orders: Glu-Asp or Asp-Glu, replacing Glu-Thr in native HisRS. The mutants were simulated by molecular dynamics in complex with histidyl-adenylate. As controls, the native HisRS was simulated in complexes with histidine, histidyl-adenylate, and histidinol. The native structures sampled were in good agreement with experimental structures and biochemical data. The two mutants with the Glu-Asp sequence showed significant differences in active site structure and Mg2+ coordination from SerRS. The others were more similar to SerRS, and one of them was analyzed further through simulations in complex with histidine, and His+ATP. The latter complex sampled two Mg2+ positions, depending on the conformation of a loop anchoring the second carboxylate. The lowest energy conformation led to an active site geometry very similar to SerRS, with the principal Mg2+ bridging the α- and β-phosphates, the first carboxylate (Asp) coordinating the ion through a water molecule, and the second (Glu) coordinating it directly. This mutant is expected to be catalytically active and suggests a basis for the previously unexplained conservation of the active site Asp-Glu pair in class II aaRSs other than HisRS. Proteins 32:362–380, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

11.
Ferrochelatase catalyzes the terminal step in the heme biosynthetic pathway, i.e., the incorporation of Fe(II) into protoporphyrin IX. Various biochemical and biophysical methods have been used to probe the enzyme for metal binding residues and the location of the active site. However, the location of the metal binding site and the path of the metal into the porphyrin are still disputed. Using site-directed mutagenesis on Bacillus subtilis ferrochelatase we demonstrate that exchange of the conserved residues His183 and Glu264 affects the metal affinity of the enzyme. We also present the first X-ray crystal structure of ferrochelatase with iron. Only a single iron was found in the active site, coordinated in a square pyramidal fashion by two amino acid residues, His183 and Glu264, and three water molecules. This iron was not present in the structure of a His183Ala modified ferrochelatase. The results strongly suggest that the insertion of a metal ion into protoporphyrin IX by ferrochelatase occurs from a metal binding site represented by His183 and Glu264.  相似文献   

12.
beta-Ketoacyl-ACP synthases catalyze the condensation steps in fatty acid and polyketide synthesis and are targets for the development of novel antibiotics and anti-obesity and anti-cancer agents. The roles of the active site residues in Streptococcus pneumoniae FabF (beta-ketoacyl-ACP synthase II; SpFabF) were investigated to clarify the mechanism for this enzyme superfamily. The nucleophilic cysteine of the active site triad was required for acyl-enzyme formation and the overall condensation activity. The two active site histidines in the elongation condensing enzyme have different electronic states and functions. His337 is essential for condensation activity, and its protonated Nepsilon stabilizes the negative charge developed on the malonyl thioester carbonyl in the transition state. The Nepsilon of His303 accelerated catalysis by deprotonating a structured active site water for nucleophilic attack on the C3 of malonate, releasing bicarbonate. Lys332 controls the electronic state of His303 and also plays a critical role in the positioning of His337. Phe396 functions as a gatekeeper that controls the order of substrate addition. These data assign specific roles for each active site residue and lead to a revised general mechanism for this important class of enzymes.  相似文献   

13.
OPs (organophosphylates) exert their acute toxicity through inhibition of acetylcholinesterase, by phosphylation of the catalytic serine residue. Engineering of human butyrylcholinesterase, by substitution of a histidine residue for the glycine residue at position 117, led to the creation of OP hydrolase activity. However, the lack of structural information and poor understanding of the hydrolytic mechanism of the G117H mutant has hampered further improvements in the catalytic activity. We have solved the crystallographic structure of the G117H mutant with a variety of ligands in its active site. A sulfate anion bound to the active site suggested the positioning for an OP prior to phosphylation. A fluoride anion was found in the active site when NaF was added to the crystallization buffer. In the fluoride complex, the imidazole ring from the His117 residue was substantially shifted, adopting a relaxed conformation probably close to that of the unliganded mutant enzyme. Additional X-ray structures were obtained from the transient covalent adducts formed upon reaction of the G117H mutant with the OPs echothiophate and VX [ethyl ({2-[bis(propan-2-yl)amino]ethyl}sulfanyl](methyl)phosphinate]. The position of the His117 residue shifted in response to the introduction of these adducts, overlaying the phosphylserine residue. These structural data suggest that the dephosphylation mechanism involves either a substantial conformational change of the His117 residue or an adjacent nucleophilic substitution by water.  相似文献   

14.
Apurinic/apyrimidinic endonuclease 1 (Ape1) is an important metal‐dependent enzyme in the base excision repair mechanism, responsible for the backbone cleavage of abasic DNA through a phosphate hydrolysis reaction. Molecular dynamics simulations of Ape1 complexed to its substrate DNA performed for models containing 1 or 2 Mg2+‐ions as cofactor located at different positions show a complex with 1 metal ion bound on the leaving group site of the scissile phosphate to be the most likely reaction‐competent conformation. Active‐site residue His309 is found to be protonated based on pKa calculations and the higher conformational stability of the Ape1‐DNA substrate complex compared to scenarios with neutral His309. Simulations of the D210N mutant further support the prevalence of protonated His309 and strongly suggest Asp210 as the general base for proton acceptance by a nucleophilic water molecule.  相似文献   

15.
We investigated a series of derivatized fullerenes possessing alcohol, amine, and amino acid pendant groups as inhibitors of the zinc enzymes carbonic anhydrases (CAs, EC 4.2.1.1). We discovered that fullerenes bind CAs with submicromolar—low micromolar affinity, despite the fact that these compounds do not possess moieties normally associated with CA inhibitors such as the sulfonamides and their isosteres, or the coumarins. The 13 different mammalian CA isoforms showed a diverse inhibition profile with these compounds. By means of computational methods we assessed the inhibition mechanism as being due to occlusion of the active site entrance by means of the fullerene cage (possessing dimension of the same order of magnitude as the opening of the enzyme cavity, of 1 nm). The pendant moieties to the fullerene cage make interactions with amino acid residues from the active site, among which His64, His94, His96, Val121, and Thr200. Fullerenes thus represent a totally new class of nanoscale CA inhibitors which may show applications for targeting physiologically relevant isoforms, such as the dominant CA II and the tumor-associated CA IX.  相似文献   

16.
We have prepared a site-specific mutant of human carbonic anhydrase (HCA) II with histidine residues at positions 7 and 64 in the active site cavity. Using a different isozyme, we have placed histidine residues in HCA III at positions 64 and 67 and in another mutant at positions 64 and 7. Each of these histidine residues can act as a proton transfer group in catalysis when it is the only nonliganding histidine in the active site cavity, except His(7) in HCA III. Using an (18)O exchange method to measure rate constants for intramolecular proton transfer, we have found that inserting two histidine residues into the active site cavity of either isozyme II or III of carbonic anhydrase results in rates of proton transfer to the zinc-bound hydroxide that are antagonistic or suppressive with respect to the corresponding single mutants. The crystal structure of Y7H HCA II, which contains both His(7) and His(64) within the active site cavity, shows the conformation of the side chain of His(64) moved from its position in the wild type and hydrogen-bonded through an intervening water molecule with the side chain of His(7). This suggests a cause of decreased proton transfer in catalysis.  相似文献   

17.
The dielectric constant in the active site cleft of subtilisin from Bacillus amyloliquefaciens has been probed by mutating charged residues on the rim and measuring the effect on the pKa value of the active site histidine (His64) by kinetics. Mutation of a negatively charged surface residue, which is 12 to 13 A from His64, to an uncharged one Asp----Ser99) lowers the pKa of the histidine by up to 0.4 unit at low ionic strength (0.005 to 0.01 M). This corresponds to an apparent dielectric constant of about 40 to 50 between Asp99 and His64. The mutation is in an external loop that is known to tolerate a serine at position 99 from homologies with subtilisins from other bacilli. The environment between His64 and Asp99 is predominantly protein. Another charged residue that is at a similar distance from His64 (14 to 15 A) and is also in an external loop that is known to tolerate a serine residue is Glu156, at the opposite side of the active site. There is only water in a direct line between His64 and Glu156. Mutation of Glu----Ser156 also lowers the pKa of His64 by up to 0.4 unit at low ionic strength. This change again corresponds to an apparent dielectric constant of about 40 to 50. The pKa values were determined from the pH dependence of kcat/KM for the hydrolysis of peptide substrates, with a precision of typically +/- 0.02 unit. The following suggests that the changes in pKa are real and not artefacts of experimental conditions: Hill plots of the data for pKa determination have gradients (h) of -1.00(+/- 0.02), showing that there are negligible systematic deviations from theoretical ionization curves involving a monobasic acid: the pH dependence for the hydrolysis of two different substrates (succinyl-L-alanyl-L-alanyl-L-prolyl-L-phenylalanyl p-nitroanilide and benzoyl-L-valyl-L-glycyl-L-arginyl p-nitroanilide) gives identical results so that the pKa is independent of substrate; the pH dependence is unaffected by changing the concentration of enzyme, so that aggregation is not affecting the results; the shift in pKa is masked by high ionic strength, as expected qualitatively for ionic shielding of electrostatic interactions.  相似文献   

18.
Tripeptidyl-peptidase II (TPP II) is a large (Mr>10(6)) tripeptide-releasing enzyme with an active site of the subtilisin-type. Compared with other subtilases, TPP II has a 200 amino-acid insertion between the catalytic Asp44 and His264 residues, and is active as an oligomeric complex. This study demonstrates that the insert is important for the formation of the active high-molecular mass complex. A recombinant human TPP II and a murine TPP II were found to display different complex-forming characteristics when over-expressed in human 293-cells; the human enzyme was mainly in a nonassociated, inactive state whereas the murine enzyme formed active oligomers. This was surprising because native human TPP II is purified from erythrocytes as an active oligomeric complex, and the amino-acid sequences of the human and murine enzymes were 96% identical. Using a combination of chimeras and a single point mutant, the amino acid responsible for this difference was identified as Arg252 in the recombinant human sequence, which corresponds to a glycine in the murine sequence. As Gly252 is conserved in all sequenced variants of TPP II, the recombinant enzyme with Arg252 is atypical. Nevertheless, as Arg252 evidently interferes with complex formation, and this residue is close to the catalytic His264, it may also explain why oligomerization influences enzyme activity. The exact mechanism for how the G252R substitution interferes with complex formation remains to be determined, but will be of importance for the understanding of the unique properties of TPP II.  相似文献   

19.
A three-dimensional structure of histo-aspartic protease (HAP), a pepsin-like enzyme from the causative agent of malaria Plasmodium falciparum, is suggested on the basis of homologous modeling followed by equilibration by the method of molecular dynamics. The presence of a His residue in the catalytic site instead of an Asp residue, which is characteristic of pepsin-like enzymes, and replacement of some other conserved residues in the active site make it possible for the enzyme to function by the covalent mechanism inherent in serine proteases. The detailed structures of HAP complexes with pepstatin, a noncovalent inhibitor of aspartic proteases, and phenylmethylsulfonyl fluoride, a covalent inhibitor of serine proteases, as well as with a pentapeptide substrate are discussed.  相似文献   

20.
The formation of the reactive enzyme–substrate complex of formate dehydrogenase has been investigated by molecular dynamics techniques accounting for different conformational states of the enzyme. Simulations revealed that the transport of substrate to the active site through the substrate channel proceeds in the open conformation of enzyme due to the crucial role of the Arg284 residue acting as a vehicle. However, formate binding in the active site of the open conformation leads to the formation of a nonproductive enzyme–substrate complex. The productive Michaelis complex is formed only in the closed enzyme conformation after the substrate and coenzyme have bound, when required rigidity of the binding site and reactive formate orientation due to interactions with Arg284, Asn146, Ile122, and His332 residues is attained. Then, the high occupancy (up to 75%) of the reactive substrate–coenzyme conformation is reached, which was demonstrated by hybrid quantum mechanics/molecular mechanics simulations using various semiempirical Hamiltonians.  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