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.
Zocher G  Wiesand U  Schulz GE 《The FEBS journal》2007,274(20):5382-5389
The crystal structure of the dimeric O-acetylserine sulfhydrylase isozyme B from Escherichia coli (CysM), complexed with the substrate analog citrate, has been determined at 1.33 A resolution by X-ray diffraction analysis. The C1-carboxylate of citrate was bound at the carboxylate position of O-acetylserine, whereas the C6-carboxylate adopted two conformations. The activity of the enzyme and of several active center mutants was determined using an assay based on O-acetylserine and thio-nitrobenzoate (TNB). The unnatural substrate TNB was modeled into the reported structure. The substrate model and the observed mutant activities may facilitate future protein engineering attempts designed to broaden the substrate spectrum of the enzyme. A comparison of the reported structure with previously published CysM structures revealed large conformational changes. One of the crystal forms contained two dimers, each of which comprised one subunit in a closed and one in an open conformation. Although the homodimer asymmetry was most probably caused by crystal packing, it indicates that the enzyme can adopt such a state in solution, which may be relevant for the catalytic reaction.  相似文献   

2.
The X-ray crystal structures of Escherichia coli gamma-aminobutyrate aminotransferase unbound and bound to the inhibitor aminooxyacetate are reported. The enzyme crystallizes from ammonium sulfate solutions in the P3(2)21 space group with a tetramer in the asymmetric unit. Diffraction data were collected to 2.4 A resolution for the unliganded enzyme and 1.9 A resolution for the aminooxyacetate complex. The overall structure of the enzyme is similar to those of other aminotransferase subgroup II enzymes. The ability of gamma-aminobutyrate aminotransferase to act on primary amine substrates (gamma-aminobutyrate) in the first half-reaction and alpha-amino acids in the second is proposed to be enabled by the presence of Glu211, whose side chain carboxylate alternates between interactions with Arg398 in the primary amine half-reaction and an alternative binding site in the alpha-amino acid half-reaction, in which Arg398 binds the substrate alpha-carboxylate. The specificity for a carboxylate group on the substrate side chain is due primarily to the presence of Arg141, but also requires substantial local main chain rearrangements relative to the structurally homologous enzyme dialkylglycine decarboxylase, which is specific for small alkyl side chains. No iron-sulfur cluster is found in the bacterial enzyme as was found in the pig enzyme [Storici, P., De Biase, D., Bossa, F., Bruno, S., Mozzarelli, A., Peneff, C., Silverman, R. B., and Schirmer, T. (2004) J. Biol. Chem. 279, 363-73.]. The binding of aminooxyacetate causes remarkably small changes in the active site structure, and no large domain movements are observed. Active site structure comparisons with pig gamma-aminobutyrate aminotransferase and dialkylglycine decarboxylase are discussed.  相似文献   

3.
Arylmalonate decarboxylase (AMDase) from Bordetella bronchiseptica catalyzes the enantioselective decarboxylation of arylmethylmalonates without the need for an organic cofactor or metal ion. The decarboxylation reaction is of interest for the synthesis of fine chemicals. As basis for an analysis of the catalytic mechanism of AMDase and for a rational enzyme design, we determined the X-ray structure of the enzyme up to 1.9 Å resolution. Like the distantly related aspartate or glutamate racemases, AMDase has an aspartate transcarbamoylase fold consisting of two α/β domains related by a pseudo dyad. However, the domain orientation of AMDase differs by about 30° from that of the glutamate racemases, and also significant differences in active-site structures are observed. In the crystals, four independent subunits showing different conformations of active-site loops are present. This finding is likely to reflect the active-site mobility necessary for catalytic activity.  相似文献   

4.
The crystal structure of a complex between chemically synthesized human immunodeficiency virus type 1 (HIV-1) protease and an octapeptide inhibitor has been refined to an R factor of 0.138 at 2.5-A resolution. The substrate-based inhibitor, H-Val-Ser-Gln-Asn-Leu psi [CH(OH)CH2]Val-Ile-Val-OH (U-85548e) contains a hydroxyethylene isostere replacement at the scissile bond that is believed to mimic the tetrahedral transition state of the proteolytic reaction. This potent inhibitor has Ki less than 1 nM and was developed as an active-site titrant of the HIV-1 protease. The inhibitor binds in an extended conformation and is involved in beta-sheet interactions with the active-site floor and flaps of the enzyme, which form the substrate/inhibitor cavity. The inhibitor diastereomer has the S configuration at the chiral carbon atom of the hydroxyethylene insert, and the hydroxyl group is within H-bonding distance of the two active-site carboxyl groups in the enzyme dimer. The two subunits of the enzyme are related by a pseudodyad, which superposes them at a 178 degrees rotation. The main difference between the subunits is in the beta turns of the flaps, which have different conformations in the two monomers. The inhibitor has a clear preferred orientation in the active site and the alternative conformation, if any, is a minor one (occupancy of less than 30%). A new model of the enzymatic mechanism is proposed in which the proteolytic reaction is viewed as a one-step process during which the nucleophile (water molecule) and electrophile (an acidic proton) attack the scissile bond in a concerted manner.  相似文献   

5.
L-Vinylglycine (L-VG) is both a substrate for and a mechanism-based inhibitor of 1-aminocyclopropane-1-carboxylate (ACC) synthase. The ratio of the rate constants for catalytic conversion to alpha-ketobutyrate and ammonia to inactivation is 500/1. The crystal structure of the covalent adduct of the inactivated enzyme was determined at 2.25 Angstroms resolution. The active site contains an external aldimine of the adduct of L-VG with the pyridoxal 5'-phosphate cofactor. The side chain gamma-carbon of L-VG is covalently bound to the epsilon-amino group of Lys273. This species corresponds to one of the two alternatives proposed by Feng and Kirsch [Feng, L. and Kirsch, J.F. (2000) L-Vinylglycine is an alternative substrate as well as a mechanism-based inhibitor of 1-aminocyclopropane-1-carboxylate synthase. Biochemistry 39, 2436-2444] and presumably results from Michael addition to a vinylglycine ketimine intermediate.  相似文献   

6.
Doucet N  Pelletier JN 《Proteins》2007,69(2):340-348
TEM-1 is a class A beta-lactamase that contributes to the primary defensive measure used by bacteria to hydrolyze the clinically-relevant beta-lactam antibiotics. Several crystal structures of this enzyme complexed with inhibitors display the active-site residue Tyr105 in an alternate orientation relative to that assigned in the free or in the substrate-bound forms. Thus, the alternate conformation may not be favored in the free enzyme and may be adopted only in the presence of inhibitor. As the residue at position 105 is a determinant of substrate specificity, we sought a better understanding of the relation between its conformation and its function in ligand binding. Here, we perform a molecular dynamics simulated annealing protocol to identify stable orientations adopted by Tyr105 in free TEM-1. Our results demonstrate that, in the absence of substrate, structurally validated conformers of Tyr105 predominantly adopt either of the two rotameric orientations observed in the crystal structures. This suggests that adoption of either conformation in the free enzyme is energetically favored and is not strictly promoted by ligand binding. We propose that free TEM-1 alternates between these two conformations of Tyr105 and that a dynamically heterogeneous population of both rotamers exists in solution. The conformational change significantly reshapes the active-site cavity and modifies the potential for forming specific ligand contacts. Our results add to the body of evidence suggesting that Tyr105 displays a dynamical behavior resulting in alternate ligand binding modes and are consistent with the lower affinity of TEM-1 for cephalosporins relative to penicillins.  相似文献   

7.
Triose-phosphate isomerase, a key enzyme of the glycolytic pathway, catalyzes the isomerization of dihydroxy acetone phosphate and glyceraldehyde 3-phosphate. In this communication we report the crystal structure of Plasmodium falciparum triose-phosphate isomerase complexed to the inhibitor 2-phosphoglycerate at 1.1-A resolution. The crystallographic asymmetric unit contains a dimeric molecule. The inhibitor bound to one of the subunits in which the flexible catalytic loop 6 is in the open conformation has been cleaved into two fragments presumably due to radiation damage. The cleavage products have been tentatively identified as 2-oxoglycerate and meta-phosphate. The intact 2-phosphoglycerate bound to the active site of the other subunit has been observed in two different orientations. The active site loop in this subunit is in both open and "closed" conformations, although the open form is predominant. Concomitant with the loop closure, Phe-96, Leu-167, and residues 208-211 (YGGS) are also observed in dual conformations in the B-subunit. Detailed comparison of the active-site geometry in the present case to the Saccharomyces cerevisiae triose-phosphate isomerase-dihydroxy acetone phosphate and Leishmania mexicana triose-phosphate isomerase-phosphoglycolate complexes, which have also been determined at atomic resolution, shows that certain interactions are common to the three structures, although 2-phosphoglycerate is neither a substrate nor a transition state analogue.  相似文献   

8.
Toney MD 《Biochemistry》2001,40(5):1378-1384
A computational study of nonenzymatic and enzymatic pyridoxal phosphate-catalyzed decarboxylation of 2-aminoisobutyrate (AIB) is presented. Four prototropic isomers of a model aldimine between AIB and 5'-deoxypyridoxal, with acetate interacting with the pyridine nitrogen, were employed in calculations of both gas phase and water model (PM3 and PM3-SM3) decarboxylation reaction paths. Calculations employing the transition state structures obtained for the four isomers allow the demonstration of stereoelectronic effects in transition state stabilization as well as a separation of the contributions of the Schiff base and pyridine ring moieties to this stabilization. The unprotonated Schiff base contribution (approximately 16 kcal/mol) is larger than that of the pyridine ring even when it is protonated (approximately 10 kcal/mol), providing an explanation of the catalytic power of pyruvoyl-dependent amino acid decarboxylases. An active site model of dialkylglycine decarboxylase was constructed and validated, and enzymatic decarboxylation reaction paths were calculated. The reaction coordinate is shown to be complex, with proton transfer from Lys272 to the coenzyme C4' likely simultaneous with C alpha--CO(2)(-) bond cleavage. The proposed concerted decarboxylation/proton-transfer mechanism provides a simple explanation for the observed specificity of this enzyme toward oxidative decarboxylation.  相似文献   

9.
Isocitrate dehydrogenases (IDHs) catalyze oxidative decarboxylation of isocitrate (ICT) into alpha-ketoglutarate (AKG). We report here the crystal structures of Saccharomyces cerevesiae mitochondrial NADP-IDH Idp1p in binary complexes with coenzyme NADP, or substrate ICT, or product AKG, and in a quaternary complex with NADPH, AKG, and Ca(2+), which represent different enzymatic states during the catalytic reaction. Analyses of these structures identify key residues involved in the binding of these ligands. Comparisons among these structures and with the previously reported structures of other NADP-IDHs reveal that eukaryotic NADP-IDHs undergo substantial conformational changes during the catalytic reaction. Binding or release of the ligands can cause significant conformational changes of the structural elements composing the active site, leading to rotation of the large domain relative to the small and clasp domains along two hinge regions (residues 118-124 and residues 284-287) while maintaining the integrity of its secondary structural elements, and thus, formation of at least three distinct overall conformations. Specifically, the enzyme adopts an open conformation when bound to NADP, a quasi-closed conformation when bound to ICT or AKG, and a fully closed conformation when bound to NADP, ICT, and Ca(2+) in the pseudo-Michaelis complex or with NADPH, AKG, and Ca(2+) in the product state. The conformational changes of eukaryotic NADP-IDHs are quite different from those of Escherichia coli NADP-IDH, for which significant conformational changes are observed only between two forms of the apo enzyme, suggesting that the catalytic mechanism of eukaryotic NADP-IDHs is more complex than that of EcIDH, and involves more fine-tuned conformational changes.  相似文献   

10.
Uracil phosphoribosyltransferase (UPRTase) catalyzes the conversion of 5-phosphate-alpha-1-diphosphate (PRPP) and uracil to uridine 5'-monophosphate (UMP) and diphosphate. The UPRTase from Sulfolobus solfataricus has a unique regulation by nucleoside triphosphates compared to UPRTases from other organisms. To understand the allosteric regulation, crystal structures were determined for S. solfataricus UPRTase in complex with UMP and with UMP and the allosteric inhibitor CTP. Also, a structure with UMP bound in half of the active sites was determined. All three complexes form tetramers but reveal differences in the subunits and their relative arrangement. In the UPRTase-UMP complex, the peptide bond between a conserved arginine residue (Arg80) and the preceding residue (Leu79) adopts a cis conformation in half of the subunits and a trans conformation in the other half and the tetramer comprises two cis-trans dimers. In contrast, four identical subunits compose the UPRTase-UMP-CTP tetramer. CTP binding affects the conformation of Arg80, and the Arg80 conformation in the UPRTase-UMP-CTP complex leaves no room for binding of the substrate PRPP. The different conformations of Arg80 coupled to rearrangements in the quaternary structure imply that this residue plays a major role in regulation of the enzyme and in communication between subunits. The ribose ring of UMP adopts alternative conformations in the cis and trans subunits of the UPRTase-UMP tetramer with associated differences in the interactions of the catalytically important Asp209. The active-site differences have been related to proposed kinetic models and provide an explanation for the regulatory significance of the C-terminal Gly216.  相似文献   

11.
Environmental effects on trimethoprim (TMP), an inhibitor of bacterial dihydrofolate reductase (DHFR), were investigated with energy minimizations in vacuo, in the crystal, and in aqueous solution. The conformations, harmonic dynamics, and energetics of the antibacterial drug calculated in these environments were compared with each other and with those of two enzyme-bound drugs. Valence and torsion angles and their energies and overall intra- and intermolecular energies compensated one another in the minimized TMP structures. The conformations of the isolated and aqueous molecules were similar to that of TMP bound to chicken liver DHFR, while the structures from the TMP crystal and from the Escherichia coli DHFR complex were unique. Since neither the small-molecule crystal nor a local minimum of the isolated molecule gave the conformation of TMP bound to the bacterial enzyme, a combination of several experimental and theoretical techniques may be necessary to probe accessible conformations of a molecule.  相似文献   

12.
During exocytosis of dense-core secretory vesicles (trichocysts) in Paramecium, the protein pp63/parafusin (pp63/pf) is transiently dephosphorylated. We report here the structures of two crystal forms of one isoform of this protein which has a high degree of homology with rabbit phosphoglucomutase, whose structure has been reported. As expected, both proteins possess highly similar structures, showing the same four domains forming two lobes with an active-site crevice in between. The two X-ray structures that we report here were determined after crystallization in the presence of sulfate and tartrate, and show the lobes arranged as a closed and an open conformation, respectively. While both conformations possess a bound divalent cation, only the closed (sulfate-bound) conformation shows bound sulfate ions in the "phosphate-transfer site" near the catalytic serine residue and in the "phosphate-binding site". Comparison with the open form shows that the latter dianion is placed in the centre of three arginine residues, one contributed by subunit II and two by subunit IV, suggesting that it causes a contraction of the arginine triangle, which establishes the observed conformational closure of the lobes. It is therefore likely that the closed conformation forms only when a phosphoryl group is bound to the phosphate-binding site. The previously published structure of rabbit phosphoglucomutase is intermediate between these two conformers. Several of the known reversible phosphorylation sites of pp63/pf-1 are at positions critical for transition between the conformations and for binding of the ligands and thus give hints as to possible roles of pp63/pf-1 in the course of exocytosis.  相似文献   

13.
The E. coli isozyme of gamma-aminobutyrate aminotransferase (GABA-AT) is a tetrameric pyridoxal phosphate-dependent enzyme that catalyzes transamination between primary amines and alpha-keto acids. The roles of the active site residues V241, E211, and I50 in the GABA-AT mechanism have been probed by site-directed mutagenesis. The beta-branched side chain of V241 facilitates formation of external aldimine intermediates with primary amine substrates, while E211 provides charge compensation of R398 selectively in the primary amine half-reaction and I50 forms a hydrophobic lid at the top of the substrate binding site. The structures of the I50Q, V241A, and E211S mutants were solved by X-ray crystallography to resolutions of 2.1, 2.5, and 2.52 A, respectively. The structure of GABA-AT is similar in overall fold and active site structure to that of dialkylglycine decarboxylase, which catalyzes both transamination and decarboxylation half-reactions in its normal catalytic cycle. Therefore, an attempt was made to convert GABA-AT into a decarboxylation-dependent aminotransferase similar to dialkylglycine decarboxylase by systematic mutation of E. coli GABA-AT active site residues. Two of the twelve mutants presented, E211S/I50G/C77K and E211S/I50H/V80D, have approximately 10-fold higher decarboxylation activities than the wild-type enzyme, and the E211S/I50H/V80D has formally changed the reaction specificity to that of a decarboxylase.  相似文献   

14.
After enrichment culture with indole-3-carboxylate in static culture, a novel reversible decarboxylase, indole-3-carboxylate decarboxylase, was found in Arthrobacter nicotianae FI1612 and several molds. The enzyme reaction was examined in resting-cell reactions with A. nicotianae FI1612. The enzyme activity was induced specifically by indole-3-carboxylate, but not by indole. The indole-3-carboxylate decarboxylase of A. nicotianae FI1612 catalyzed the nonoxidative decarboxylation of indole-3-carboxylate into indole, and efficiently carboxylated indole and 2-methylindole by the reverse reaction. In the presence of 1 mM dithiothreitol, 50 mM Na2 S2O3, and 20% (v/v) glycerol, indole-3-carboxylate decarboxylase was partially purified from A. nicotianae FI1612. The purified enzyme had a molecular mass of approximately 258 kDa. The enzyme did not need any cofactor for the decarboxylating and carboxylating reactions.  相似文献   

15.
DOPA decarboxylase (DDC) is responsible for the synthesis of the key neurotransmitters dopamine and serotonin via decarboxylation of L-3,4-dihydroxyphenylalanine (L-DOPA) and L-5-hydroxytryptophan, respectively. DDC has been implicated in a number of clinic disorders, including Parkinson's disease and hypertension. Peripheral inhibitors of DDC are currently used to treat these diseases. We present the crystal structures of ligand-free DDC and its complex with the anti-Parkinson drug carbiDOPA. The inhibitor is bound to the enzyme by forming a hydrazone linkage with the cofactor, and its catechol ring is deeply buried in the active site cleft. The structures provide the molecular basis for the development of new inhibitors of DDC with better pharmacological characteristics.  相似文献   

16.
We employ ensemble docking simulations to characterize the interactions of two enantiomeric forms of a Ru-complex compound (1-R and 1-S) with three protein kinases, namely PIM1, GSK-3β, and CDK2/cyclin A. We show that our ensemble docking computational protocol adequately models the structural features of these interactions and discriminates between competing conformational clusters of ligand-bound protein structures. Using the determined X-ray crystal structure of PIM1 complexed to the compound 1-R as a control, we discuss the importance of including the protein flexibility inherent in the ensemble docking protocol, for the accuracy of the structure prediction of the bound state. A comparison of our ensemble docking results suggests that PIM1 and GSK-3β bind the two enantiomers in similar fashion, through two primary binding modes: conformation I, which is very similar to the conformation presented in the existing PIM1/compound 1-R crystal structure; conformation II, which represents a 180° flip about an axis through the NH group of the pyridocarbazole moiety, relative to conformation I. In contrast, the binding of the enantiomers to CDK2 is found to have a different structural profile including a suggested bound conformation, which lacks the conserved hydrogen bond between the kinase and the ligand (i.e., ATP, staurosporine, Ru-complex compound). The top scoring conformation of the inhibitor bound to CDK2 is not present among the top-scoring conformations of the inhibitor bound to either PIM1 or GSK-3β and vice-versa. Collectively, our results help provide atomic-level insights into inhibitor selectivity among the three kinases.
Figure
Top cluster of predicted conformations based on ensemble docking simulations of a Ruthenium based compound to protein kinases  相似文献   

17.
The structures of tomato 1-aminocyclopropane-1-carboxylate synthase (ACS) in complex with either cofactor pyridoxal-5'-phosphate (PLP) or both PLP and inhibitor aminoethoxyvinylglycine have been determined by x-ray crystallography. The structures showed good conservation of the catalytic residues, suggesting a similar catalytic mechanism for ACS and other PLP-dependent enzymes. However, the proximity of Tyr152 to the C-gamma-S bond of model substrate S-adenosylmethionine implies its critical role in the catalysis. The concerted accomplishment of catalysis by cofactor PLP and a protein residue, as proposed on the basis of the ACS structures in this paper, may represent a general scheme for the diversity of PLP-dependent catalyses. PLP-dependent enzymes have been categorized into four types of folds. A structural comparison revealed that a core fragment of ACS in fold type I is superimposable over tryptophan synthase beta subunit in fold type II and mouse ornithine decarboxylase in fold type III, thus suggesting a divergent evolution of PLP-dependent enzymes.  相似文献   

18.
We report the cocrystal structures of a computationally designed and experimentally optimized retro-aldol enzyme with covalently bound substrate analogs. The structure with a covalently bound mechanism-based inhibitor is similar to, but not identical with, the design model, with an RMSD of 1.4 Å over active-site residues and equivalent substrate atoms. As in the design model, the binding pocket orients the substrate through hydrophobic interactions with the naphthyl moiety such that the oxygen atoms analogous to the carbinolamine and β-hydroxyl oxygens are positioned near a network of bound waters. However, there are differences between the design model and the structure: the orientation of the naphthyl group and the conformation of the catalytic lysine are slightly different; the bound water network appears to be more extensive; and the bound substrate analog exhibits more conformational heterogeneity than typical native enzyme–inhibitor complexes. Alanine scanning of the active-site residues shows that both the catalytic lysine and the residues around the binding pocket for the substrate naphthyl group make critical contributions to catalysis. Mutating the set of water-coordinating residues also significantly reduces catalytic activity. The crystal structure of the enzyme with a smaller substrate analog that lacks naphthyl ring shows the catalytic lysine to be more flexible than in the naphthyl–substrate complex; increased preorganization of the active site would likely improve catalysis. The covalently bound complex structures and mutagenesis data highlight the strengths and weaknesses of the de novo enzyme design strategy.  相似文献   

19.
S-Adenosylmethionine decarboxylase belongs to a small class of amino acid decarboxylases that use a covalently bound pyruvate as a prosthetic group. It is an essential enzyme for polyamine biosynthesis and provides an important target for the design of anti-parasitic and cancer chemotherapeutic agents. We have determined the structures of S-adenosylmethionine decarboxylase complexed with the competitive inhibitors methylglyoxal bis(guanylhydrazone) and 4-amidinoindan-1-one-2'-amidinohydrazone as well as the irreversible inhibitors 5'-deoxy-5'-[N-methyl-N-[(2-aminooxy)ethyl]amino]adenosine, 5'-deoxy-5'-[N-methyl-N-(3-hydrazinopropyl)amino]adenosine, and the methyl ester analogue of S-adenosylmethionine. These structures elucidate residues important for substrate binding and show how those residues interact with both covalently and noncovalently bound inhibitors. S-Adenosylmethionine decarboxylase has a four-layer alphabeta betaalpha sandwich fold with residues from both beta-sheets contributing to substrate and inhibitor binding. The side chains of conserved residues Phe7, Phe223, and Glu247 and the backbone carbonyl of Leu65 play important roles in binding and positioning the ligands. The catalytically important residues Cys82, Ser229, and His243 are positioned near the methionyl group of the substrate. One molecule of putrescine per monomer is observed between the two beta-sheets but far away from the active site. The activating effects of putrescine may be due to conformational changes in the enzyme, to electrostatic effects, or both. The adenosyl moiety of the bound ligand is observed in the unusual syn conformation. The five structures reported here provide a framework for interpretation of S-adenosylmethionine decarboxylase inhibition data and suggest strategies for the development of more potent and more specific inhibitors of S-adenosylmethionine decarboxylase.  相似文献   

20.
We have carried out NMR and molecular modeling studies of peptidomimetic HIV-1 protease inhibitors, LB71116: Qc-Asn-Phepsi[(1R,2S)-cis-epoxide]Gly-NH-CH(isopropyl)2 where Qc stands for quinaldic acid and LB71148: Qc-(SMe)Pen(O)2-Phepsi[(1R,2S)-cis-epoxide]Gly-NH-CH(isoprop yl)2 where (SMe)Pen(O)2 stands for S-methyl-S-dioxo-penicillamine. Through conformational calculations and NMR data analysis, we have obtained preferred conformations of the two inhibitors in solution. To our knowledge, this work is one of the first extensive conformational studies of peptidomimetics containing cis-epoxide amide isostere. The resulting preferred conformations contain extended structures. In these conformations, the psi of Phe(cep) is maintained about 130 degrees and the phi angle of (cep)Gly prefers +/- 150 degrees [where Phe(cep) and (cep)Gly are the residues generated by the replacement of the Phe-Gly peptide bond with cis-epoxide]. Two conformations were commonly observed in the preferred conformations of each inhibitor. Through restrained molecular dynamics simulating the hydrogen bond formation between our inhibitor and a water molecule ('flap water'), one of the conformations is assumed as the conformation which can bind to the enzyme without large conformational changes. Recently, we had the opportunity to compare the selected preferred conformation with the binding conformation of LB71116 observed from the X-ray studies of the complex between LB71116 and HIV-1 protease. These two conformations are surprisingly similar to each other. Thus, we can explain high activity and selectivity of our inhibitors to the HIV-1 protease by the similarity between the preferred conformations in solution and the binding conformation.  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