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.
Docking and molecular dynamics were used to study the nine ligands (see Scheme 1) at the neuraminidase (NA) active sites. Their binding modes are structurally and energetically different, with details given in the text. Compared with 1A (oseltamivir carboxylate), the changes of core template or/and functional groups in the other ligands cause the reductions of interaction energies and numbers of H-bonds with the NA proteins. Nonetheless, all these ligands occupy the proximity space at the NA active sites and share some commonness in their binding modes. The fragment approach was then used to analyze and understand the binding specificities of the nine ligands. The contributions of each core template and functional group were evaluated. It was found that the core templates rather than functional groups play a larger role during the binding processes; in addition, the binding qualities are determined by the synergistic effects of the core templates and functional groups. Among the nine ligands, 1A (oseltamivir carboxylate) has the largest synergistic energy and its functional groups fit perfectly with the NA active site, consistent with the largest interaction energy, numerous H-bonds with the NA active-site residues as well as experimentally lowest IC50 value. Owing to the poorer metabolizability than oseltamivir, large contribution of the benzene core template and fine synergistic effects of the functional groups, the 4-(N-acetylamino)-5-guanidino-3-(3-pentyloxy)benzoic acid should be an ideal lead compound for optimizing NA drugs.  相似文献   

2.
Many strains of Streptococcus pyogenes bind C4b-binding protein (C4BP), an inhibitor of complement activation. The binding is mediated by surface M proteins in a fashion that has been suggested to mimic the binding of C4b. We have previously shown that a positively charged cluster at the interface between complement control protein domains 1 and 2 of C4BP alpha-chain is crucial for the C4b-C4BP interaction. To extend this observation, and to investigate the interaction with M proteins, we constructed and characterized a total of nine mutants of C4BP. We identified a key recognition surface for M proteins that overlaps with the C4b binding site because substitution of R64 and H67 by Gln dramatically reduces binding to both ligands. However, the analysis of all mutants indicates that the binding sites for C4b and M proteins are only overlapping, but not identical. Furthermore, M proteins were able to displace C4BP from immobilized C4b, whereas C4b only weakly affected binding of C4BP to immobilized M proteins. We found that the molecular mechanisms involved in these two interactions differ because the binding between M proteins and C4BP is relatively insensitive to salt in contrast to the C4BP-C4b binding. In addition, six mAbs directed against the alpha-chain interfered with C4b-C4BP interaction, whereas only two of them efficiently inhibited binding of C4BP to M proteins. Collectively, our results suggest that binding between C4b and C4BP is governed mostly by electrostatic interactions, while additional noncovalent forces cause tight binding of C4BP to streptococcal M proteins.  相似文献   

3.
The influenza neuraminidase (NA) inhibitors zanamivir, oseltamivir and peramivir were all designed based on the knowledge that the transition state analogue of the cleaved sialic acid, 2-deoxy,2,3-dehydro N-acetyl neuraminic acid (DANA) was a weak inhibitor of NA. While DANA bound rapidly to the NA, modifications leading to the improved potency of these new inhibitors also conferred a time dependent or slow binding phenotype. Many mutations in the NA leading to decreased susceptibility result in loss of slow binding, hence this is a phenotypic marker of many but not all resistant NAs. We present here a simplified approach to determine whether an inhibitor is fast or slow binding by extending the endpoint fluorescent enzyme inhibition assay to a real time assay and monitoring the changes in IC(50)s with time. We carried out two reactions, one with a 30 min preincubation with inhibitor and the second without. The enzymatic reaction was started via addition of substrate and IC(50)s were calculated after each 10 min interval up to 60 min. Results showed that without preincubation IC(50)s for the wild type viruses started high and although they decreased continuously over the 60 min reaction time the final IC(50)s remained higher than for pre-incubated samples. These results indicate a slow equilibrium of association and dissociation and are consistent with slow binding of the inhibitors. In contrast, for viruses with decreased susceptibility, preincubation had minimal effect on the IC(50)s, consistent with fast binding. Therefore this modified assay provides additional phenotypic information about the rate of inhibitor binding in addition to the IC(50), and critically demonstrates the differential effect of incubation times on the IC(50) and K(i) values of wild type and mutant viruses for each of the inhibitors.  相似文献   

4.
A model binding site was used to investigate charge-charge interactions in molecular docking. This simple site, a small (180A(3)) engineered cavity in cyctochrome c peroxidase (CCP), is negatively charged and completely buried from solvent, allowing us to explore the balance between electrostatic energy and ligand desolvation energy in a system where many of the common approximations in docking do not apply. A database with about 5300 molecules was docked into this cavity. Retrospective testing with known ligands and decoys showed that overall the balance between electrostatic interaction and desolvation energy was captured. More interesting were prospective docking scre"ens that looked for novel ligands, especially those that might reveal problems with the docking and energy methods. Based on screens of the 5300 compound database, both high-scoring and low-scoring molecules were acquired and tested for binding. Out of 16 new, high-scoring compounds tested, 15 were observed to bind. All of these were small heterocyclic cations. Binding constants were measured for a few of these, they ranged between 20microM and 60microM. Crystal structures were determined for ten of these ligands in complex with the protein. The observed ligand geometry corresponded closely to that predicted by docking. Several low-scoring alkyl amino cations were also tested and found to bind. The low docking score of these molecules owed to the relatively high charge density of the charged amino group and the corresponding high desolvation penalty. When the complex structures of those ligands were determined, a bound water molecule was observed interacting with the amino group and a backbone carbonyl group of the cavity. This water molecule mitigates the desolvation penalty and improves the interaction energy relative to that of the "naked" site used in the docking screen. Finally, six low-scoring neutral molecules were also tested, with a view to looking for false negative predictions. Whereas most of these did not bind, two did (phenol and 3-fluorocatechol). Crystal structures for these two ligands in complex with the cavity site suggest reasons for their binding. That these neutral molecules do, in fact bind, contradicts previous results in this site and, along with the alkyl amines, provides instructive false negatives that help identify weaknesses in our scoring functions. Several improvements of these are considered.  相似文献   

5.
The structure of the recombinant Trypanosoma rangeli sialidase (TrSA) has been determined at 1.6A resolution, and the structures of its complexes with the transition state analog inhibitor 2-deoxy-2,3-dehydro-N-acetyl-neuraminic acid (DANA), Neu-5-Ac-thio-alpha(2,3)-galactoside (NATG) and N-acetylneuraminic acid (NANA) have been determined at 1.64A, 2.1A and 2.85A, respectively. The 3D structure of TrSA is essentially identical to that of the natural enzyme, except for the absence of covalently attached sugar at five distinct N-glycosylation sites. The protein exhibits a topologically rigid active site architecture that is unaffected by ligand binding. The overall binding of DANA to the active site cleft is similar to that observed for other viral and bacterial sialidases, dominated by the interactions of the inhibitor carboxylate with the conserved arginine triad. However, the interactions of the other pyranoside ring substituents (hydroxyl, N-acetyl and glycerol moieties) differ between trypanosomal, bacterial and viral sialidases, providing a structural basis for specific inhibitor design. Sialic acid is found to bind the enzyme with the sugar ring in a distorted (half-chair or boat) conformation and the 2-OH hydroxyl group at hydrogen bonding distance of the carboxylate of Asp60, substantiating a direct catalytic role for this residue. A detailed comparison of TrSA with the closely related structure of T.cruzi trans-sialidase (TcTS) reveals a highly conserved catalytic center, where subtle structural differences account for strikingly different enzymatic activities and inhibition properties. The structure of TrSA in complex with NATG shows the active site cleft occupied by a smaller compound which could be identified as DANA, probably the product of a hydrolytic side reaction. Indeed, TrSA (but not TcTS) was found to cleave O and S-linked sialylated substrates, further stressing the functional differences between trypanosomal sialidases and trans-sialidases.  相似文献   

6.
Human cytosolic sialidase (Neuraminidase 2, NEU2) catalyzes the removal of terminal sialic acid residues from glycoconjugates. The effect of siastatin B, known as a sialidase inhibitor, has not been evaluated toward human NEU2 yet. We studied the regulation of NEU2 activity by siastatin B in vitro and predicted the interaction in silico. Inhibitory and stabilizing effects of siastatin B were analyzed in comparison with DANA (2-deoxy-2,3-dehydro-N-acetylneuraminic acid) toward 4-umbelliferyl N-acetylneuraminic acid (4-MU-NANA)- and α2,3-sialyllactose-degrading activities of recombinant NEU2 produced by E. coli GST-fusion gene expression. Siastatin B exhibited to have higher competitive inhibitory activity toward NEU2 than DANA at pH 4.0. We also revealed the stabilizing effect of siastatin B toward NEU2 activity at acidic pH. Docking model was constructed on the basis of the crystal structure of NEU2/DANA complex (PDB code: 1VCU). Molecular docking predicted that electrostatic neutralization of E111 and E218 residues of the active pocket should not prevent siastatin B from binding at pH 4.0. The imino group (1NH) of siastatin B can also interact with D46, neutralized at pH 4.0. Siastatin B was suggested to have higher affinity to the active pocket of NEU2 than DANA, although it has no C7–9 fragment corresponding to that of DANA. We demonstrated here the pH-dependent affinity of siastatin B toward NEU2 to exhibit potent inhibitory and stabilizing activities. Molecular interaction between siastatin B and NEU2 will be utilized to develop specific inhibitors and stabilizers (chemical chaperones) not only for NEU2 but also the other human sialidases, including NEU1, NEU3 and NEU4, based on homology modeling.  相似文献   

7.
Interactions of stromelysin with a series of inhibitors representative of three chemical templates with distinct binding modes were examined. Unfolding temperatures for inhibitor complexes were 10 degrees C to 15 degrees C greater than for apo stromelysin. Minor changes in ellipticity in the far-UV CD spectra of complexes indicated that ligand-induced conformational changes were localized to the binding site and did not involve gross changes in protein folding. Isothermal titrating calorimetry of thiadiazole-containing inhibitors, which bind in the S(1)-S(3) subsites of stromelysin, indicated that the binding interaction was exothermic and only slightly favorable entropically. Near-UV CD spectra showed large positive ellipticity increases from 250 to 300 nm, consistent with an interaction between the benzene ring of the inhibitor and stromelysin residues Tyr155 and Tyr168. Interactions between stromelysin and amide-hydroxamate ligands, which bind in the S(')(1)-S(')(3) subsites, were found to be both enthalpically and entropically driven. Binding of this class of ligands resulted in modest negative ellipticity changes at 260-285 nm and positive increases at 292 nm. Stromelysin complexed to a lactam-hydroxamate inhibitor with structure extending into both the S(1)-S(3) and S(')(1)-S(')(3) subsites showed increased ellipticity at 245 nm and negative changes at 260-285 and 295 nm.  相似文献   

8.
The paramyxovirus hemagglutinin-neuraminidase (HN) is a multifunctional protein mediating hemagglutination (HA), neuraminidase (NA), and fusion promotion activities. It has been a matter of debate whether HN contains combined or separate sites for HA and NA activities. To clear the issue, we determined the presence of the second binding site on human parainfluenza virus (hPIV) type 1, 2, and 3 and Sendai virus (SeV) HN proteins. Results of virus elution from erythrocytes at an elevated temperature and HA inhibition by NA inhibitor BCX-2798 suggest that all hPIVs bind to the receptor only through the NA catalytic site, while SeV HN has an additional receptor binding site. Comparison of SeV and hPIV1 HN sequences revealed two amino acid differences at residues 521 and 523 in the region close to the second binding site identified in Newcastle disease virus HN. We mutated hPIV1 HN at position 523 from Asn to the residue of SeV HN, Asp, and rescued a recombinant SeV that carries the mutated hPIV1 HN by a reverse genetics system. The hPIV1 HN with Asp at position 523 hemagglutinated in the presence of BCX-2798, suggesting that the amino acid difference at position 523 is critical for the formation of a second binding site. Creation of the second binding site on hPIV1 HN, however, did not significantly affect the growth or fusion activity of the recombinant virus. Our study indicates that the presence and requirement of a second binding site vary among paramyxoviruses.  相似文献   

9.
The Duffy binding-like (DBL) domain is a key adhesive module in Plasmodium falciparum, present in both erythrocyte invasion ligands (EBLs) and the large and diverse P. falciparum erythrocyte membrane protein 1 (PfEMP1) family of cytoadherence receptors. DBL domains bind a variety of different host receptors, including intercellular adhesion molecule 1 (ICAM-1), a receptor interaction that may have a role in infected erythrocyte binding to cerebral blood vessels and cerebral malaria. In this study, we expressed the nearly full complement of DBLbeta-C2 domains from the IT4/25/5 (IT4) parasite isolate and showed that ICAM-1-binding domains (DBLbeta-C2(ICAM-1)) were confined to group B and group C PfEMP1 proteins and were not present in group A, suggesting that ICAM-1 selection pressure differs between PfEMP1 groups. To further dissect the molecular determinants of binding, we modelled a DBLbeta-C2(ICAM-1) domain on a solved DBL structure and created alanine substitution mutants in two DBLbeta-C2(ICAM-1) domains. This analysis indicates that the DBLbeta-C2::ICAM-1 interaction maps to the equivalent glycan binding region of EBLs, and suggests a general model for how DBL domains evolve under dual selection for host receptor binding and immune evasion.  相似文献   

10.
Gee SH  Quenneville S  Lombardo CR  Chabot J 《Biochemistry》2000,39(47):14638-14646
PDZ domains are modular protein-protein interaction domains that bind to specific C-terminal sequences of membrane proteins and/or to other PDZ domains. Certain PDZ domains in PSD-95 and syntrophins interact with C-terminal peptide ligands and heterodimerize with the extended nNOS PDZ domain. The capacity to interact with nNOS correlates with the presence of a Lys residue in the carboxylate- binding loop of these PDZ domains. Here, we report that substitution of an Arg for Lys-165 in PSD-95 PDZ2 disrupted its interaction with nNOS, but not with the C terminus of the Shaker-type K(+) channel Kv1.4. The same mutation affected nNOS binding to alpha1- and beta1-syntrophin PDZ domains to a lesser extent, due in part to the stabilizing effect of tertiary interactions with the canonical nNOS PDZ domain. PDZ domains with an Arg in the carboxylate-binding loop do not bind nNOS; however, substitution with Lys or Ala was able to confer nNOS binding. Our results indicate that the carboxylate-binding loop Lys or Arg is a critical determinant of nNOS binding and that the identity of this residue can profoundly alter one mode of PDZ recognition without affecting another. We also analyzed the effects of mutating Asp-143, a residue in the alphaB helix of alpha1-syntrophin that forms a tertiary contact with the nNOS PDZ domain. This residue is important for both nNOS and C-terminal peptide binding and confers a preference for peptides with a positively charged residue at position -4. On this basis, we have identified the C terminus of the Kir2.1 channel as a possible binding partner for syntrophin PDZ domains. Together, our results demonstrate that single-amino acid substitutions alter the specificity and affinity of PDZ domains for their ligands.  相似文献   

11.
In previous work, we reconstituted salinixanthin, the C(40)-carotenoid acyl glycoside that serves as a light-harvesting antenna to the light-driven proton pump xanthorhodopsin, into a different protein, gloeobacter rhodopsin expressed in Escherichia coli, and demonstrated that it transfers energy to the retinal chromophore [Imasheva, E. S., et al. (2009) Biochemistry 48, 10948]. The key to binding of salinixanthin was the accommodation of its ring near the retinal β-ionone ring. Here we examine two questions. Do any of the native Gloeobacter carotenoids bind to gloeobacter rhodopsin, and does the 4-keto group of the ring play a role in binding? There is no salinixanthin in Gloeobacter violaceous, but a simpler carotenoid, echinenone, also with a 4-keto group but lacking the acyl glycoside, is present in addition to β-carotene and oscillol. We show that β-carotene does not bind to gloeobacter rhodopsin, but its 4-keto derivative, echinenone, does and functions as a light-harvesting antenna. This indicates that the 4-keto group is critical for carotenoid binding. Further evidence of this is the fact that salinixanthol, an analogue of salinixanthin in which the 4-keto group is reduced to hydroxyl, does not bind and is not engaged in energy transfer. According to the crystal structure of xanthorhodopsin, the ring of salinixanthin in the binding site is turned out of the plane of the polyene conjugated chain. A similar conformation is expected for echinenone in the gloeobacter rhodopsin. We suggest that the 4-keto group in salinixanthin and echinenone allows for the twisted conformation of the ring around the C6-C7 bond and probably is engaged in an interaction that locks the carotenoid in the binding site.  相似文献   

12.
Concanavalin A (Con A) and agglutinins from the pea (PSA), lentil (LCH), and fava bean (VFA) constitute a group of D-mannose/D-glucose binding legume lectins. In addition to their sugar binding specificity, these lectins also contain sites that bind hydrophobic ligands. The present study explores a class of nonpolar binding sites reportedly present adjacent to the carbohydrate binding site in PSA, LCH, and VFA. A series of 2-O- and 3-O-substituted nitrobenzoyl and nitrobenzyl derivatives of methyl alpha-D-glucopyranoside and methyl alpha-D-mannopyranoside were synthesized. Evaluation of their binding to Con A, PSA, LCH, and VFA was carried out by the technique of hapten inhibition of precipitation reaction. The hapten inhibition assay results reveal that the presence of a methyl or methylene group at the O-2 or O-3 position of the sugar is essential for hydrophobic interaction with PSA, LCH, and VFA. The substitution of methyl by nitrobenzyl leads to enhanced binding (1.7-16.7 times for the 2-O-substituted compounds and 7.9-40.5 times for the 3-O-substituted compounds) with the m-nitrobenzyl group contributing to maximum binding. A hydrophobic interaction is also involved between Con A and 2-O-nitrobenzyl derivatives, resulting in enhanced binding, but the corresponding 3-O-isomers bind poorly due probably to steric reasons. These results may be rationalized on the basis of the recently published X-ray data of Con A and VFA. The nitrobenzyl derivatives, after transformation to their azido analogs, have potential applications in the photoaffinity labeling of these lectins.  相似文献   

13.
Different low molecular mass ligands have been used to identify amyloid deposits. Among these markers, the dyes Thioflavin T and Congo Red interact specifically with the beta-sheet structure arranged in a cross-beta conformation, which is characteristic of amyloid. However, the molecular details of this interaction remain unknown. When labelled with technetium-99m, the proteinase inhibitor aprotinin has been shown to represent a very important radiopharmaceutical agent for in vivo imaging of extra-abdominal deposition of amyloid in amyloidosis of the immunoglobulin type. However, no information is available as to whether aprotinin binds other types of amyloid fibrils and on the nature and characteristics of the interaction. The present work shows aprotinin binding to insulin, transthyretin, beta-amyloid peptide and immunoglobulin synthetic amyloid fibrils by a specific dot-blot ligand-binding assay. Aprotinin did not bind amorphous precipitates and/or the soluble fibril precursors. A Ka of 2.9 microM-1 for the binding of aprotinin to insulin amyloid fibrils was determined by Scatchard analysis. In competition experiments, analogues such as an aprotinin variant, a spermadhesin and the soybean trypsin inhibitor were tested and results suggest that both aprotinin and the spermadhesin interact with amyloid fibrils through pairing of beta-sheets of the ligands with exposed structures of the same type at the surface of amyloid deposits. An electrostatic component may also be involved in the binding of aprotinin to amyloid fibrils because important differences in binding constants are observed when substitutions V15L17E52 are introduced in aprotinin; on the other hand beta-sheet containing acidic proteins, such as the soybean trypsin inhibitor, are unable to bind amyloid fibrils.  相似文献   

14.
Recently we designed the first small organic ligands, sulfated flavanoids and flavonoids, that act as activators of antithrombin for accelerated inhibition of factor Xa, a key proteinase of the coagulation cascade [Gunnarsson and Desai, Bioorg. Med. Chem. Lett. (2003) 13:579]. To better understand the binding properties of these activators at a molecular level, we have utilized computerized hydropathic interaction (HINT) analyses of the sulfated molecules interacting in two plausible electropositive regions, the pentasaccharide- and extended heparin-binding sites, of antithrombin in its native and activated forms. HINT analyses indicate favorable multi-point interactions of the activators in both binding sites of the two forms of antithrombin. Yet, HINT predicts better interaction of most activators, except for (-)-catechin sulfate, with the activated form of antithrombin than with the native form supporting the observation in solution that these molecules function as activators of the inhibitor. Further, whereas (+)-catechin sulfate recognized the activated form of antithrombin better in both the pentasaccharide- and extended heparin- binding sites, the native form was better recognized by (-)-catechin sulfate, thus explaining its weaker binding and activation potential in solution. A reasonable linear correlation between the overall HINT score and the solution free energy of binding of the sulfated activators was evident. This investigation indicates that HINT is a useful tool in understanding interactions of antithrombin with small sulfated organic ligands at a molecular level, has some good predictive properties, and is likely to be useful for rational design purposes.  相似文献   

15.
Neuraminidase (NA) of influenza is a key target for virus infection control and the recently discovered open 150-cavity in group-1 NA provides new opportunity for novel inhibitors design. In this study, we used a combination of theoretical methods including fragment docking, molecular linking and molecular dynamics simulations to design ligands that specifically target at the 150-cavity. Through in silico screening of a fragment compound library on the open 150-cavity of NA, a few best scored fragment compounds were selected to link with Zanamivir, one NA-targeting drug. The resultant new ligands may bind both the active site and the 150-cavity of NA simultaneously. Extensive molecular dynamics simulations in explicit solvent were applied to validate the binding between NA and the designed ligands. Moreover, two control systems, a positive control using Zanamivir and a negative control using a low-affinity ligand 3-(p-tolyl) allyl-Neu5Ac2en (ETT, abbreviation reported in the PDB) found in a recent experimental work, were employed to calibrate the simulation method. During the simulations, ETT was observed to detach from NA, on the contrary, both Zanamivir and our designed ligand bind NA firmly. Our study provides a prospective way to design novel inhibitors for controlling the spread of influenza virus.  相似文献   

16.
A low-molecular-weight biomimetic affinity ligand selective for binding elastase has been designed and synthesized. The ligand was based on mimicking part of the interaction between a natural inhibitor, turkey ovomucoid inhibitor and elastase, and modelled from the X-ray crystallographic structure of the enzyme-inhibitor complex. Limited solid-phase combinatorial chemistry was used to synthesize 12 variants of the lead ligand using the triazine moiety as the scaffold for assembly. The ligand library was screened for its ability to bind elastase and trypsin, and two ligands were studied further. Ligand C4/6 [2-alanyl-alanyl-4-tryptamino-6-(alpha-lysyl)-s-triazine] was found to bind porcine pancreatic elastase, but not trypsin, with a dissociation constant of 6 x 10(-5) M and a binding capacity of 21 mg elastase per ml gel. The adsorbent was used to purify elastase from a crude extract of porcine pancreas. Immobilized ligand C4/5 6 [2-alanyl-alanyl-4-tyramino-6-(alpha-lysyl)-s-triazine] was similarly chosen for optimal binding of elastase from cod and used to purify the enzyme from a crude extract of cod pyloric caeca. Ligand C4/6 was subsequently synthesized in solution and its structure verified by 1H-NMR.  相似文献   

17.
B‐cell lymphoma extra‐large protein (BclXL) serves as an apoptotic repressor by virtue of its ability to recognize and bind to BH3 domains found within a diverse array of proapoptotic regulators. Herein, we investigate the molecular basis of the specificity of the binding of proapoptotic BH3 ligands to BclXL. Our data reveal that while the BH3 ligands harboring the LXXX[A/S]D and [R/Q]XLXXXGD motif bind to BclXL with high affinity in the submicromolar range, those with the LXXXGD motif afford weak interactions. This suggests that the presence of a glycine at the fourth position (G+4)—relative to the N‐terminal leucine (L0) within the LXXXGD motif—mitigates binding, unless the LXXXGD motif also contains arginine/glutamine at the ?2 position. Of particular note is the observation that the residues at the +4 and ?2 positions within the LXXX[A/S]D and [R/Q]XLXXXGD motifs appear to be energetically coupled—replacement of either [A/S]+4 or [R/Q]‐2 with other residues has little bearing on the binding affinity of BH3 ligands harboring one of these motifs. Collectively, our study lends new molecular insights into understanding the binding specificity of BH3 ligands to BclXL with important consequences on the design of novel anticancer drugs. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 573–582, 2014.  相似文献   

18.
In spite of showing high sequence similarity and forming structurally similar ternary complex in vitro, the in vivo role of TGF‐β1 and TGF‐β3 ligands suggests against their functional redundancy and necessitates the importance for the study of the specificity of these ligands. A comparative computational analysis of binary and ternary complexes of these two ligands shows that anchor residues of ligand and receptor at TGF‐β:TβR2 interface are similar in both complexes. However, the potential anchor residues of TGF‐β at TGF‐β:TβR1 interface are different, Tyr50 and Lys51 in TGF‐β3 complex and Lys60 and Tyr6 in TGF‐β1 complex. Pro55 and Asp57 of TβRI may act as anchor residues in complexes of both ligands along with Ile54 for TGF‐β3 complex and Val61 for TGF‐β1 complex. Arg58 of TβR1 acts as a potential hot residue for TGF‐β3 ternary complex but not for TGF‐β1 ternary complex formation whereas Pro55 and Phe60 may act as hot residues for both complexes. The Delphi analysis of the pH dependence of the binding energy indicates that pH has a remarkable effect on the binding energy of TβR2 to the open form of TGF‐β3. Lowering of pH from 7 to 4 favors binding of the open form of TGF‐β3 to TβR2. Now, apart from the residues at pH 7, residues Arg25, Lys31 and Arg94 of TGF‐β3 and Asp118 and Glu119 of TβR2 also contribute significantly to the binding energy. Contrary to the binding energy of TβR2 to TGF‐β3/TGF‐β1, TβR1 shows appreciable pH dependence for its binding in ternary complex of TGF‐β3/TGF‐β1. In TGF‐β3 ternary complex, the TβR1 electrostatic interaction energy disfavors complex formation at pH 7 while it is favored at pH 4. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

19.
Blanchet J  Lin SX  Zhorov BS 《Biochemistry》2005,44(19):7218-7227
Crystallographic studies of ligand-protein complexes reveal most preferable ligand binding modes, but do not show less populated modes that may contribute to measurable biochemical and biophysical characteristics of the complexes. In some cases, a ligand may bind a protein in essentially different modes. An example is 17beta-hydroxysteroid dehydrogenase type 1 (17beta-HSD1), a steroidogenic enzyme that catalyzes reduction of estrone to estradiol in gonadal and peripheral tissues. The enzyme exhibits a high specificity for estrogens which bind with their C17 atom in the proximity of the NADP(H) cofactor. 17Beta-HSD1 can also bind androgens, but in a reverse binding mode, in which the steroid C3 atom is the closest carbon atom to the cofactor. Here we map the interaction energy of estradiol and dihydrotestosterone binding to 17beta-HSD1. Positions and orientations of the steroids in the ligand-binding tunnel were sampled systematically, and at each combination of these generalized coordinates, the energy was Monte Carlo minimized. The computed maps show energy minima corresponding to the X-ray structures and predict alternative binding modes, in particular, an upside-down orientation in which steroidal face alpha is exposed to protein residues that normally interact with face beta. The methodology can be used for mapping ligand-receptor interactions in various systems, for example, in ion channels and G-protein-coupled receptors that bind elongated ligands in confined space between transmembrane helices.  相似文献   

20.
Protein kinase A (cAMP dependent protein kinase catalytic subunit, EC 2.7.11.11) binds simultaneously ATP and a phosphorylatable peptide. These structurally dissimilar allosteric ligands influence the binding effectiveness of each other. The same situation is observed with substrate congeners, which reversibly inhibit the enzyme. In this review these allosteric effects are quantified using the interaction factor, which compares binding effectiveness of ligands with the free enzyme and the pre-loaded enzyme complex containing another ligand. This analysis revealed that the allosteric effect depends upon structure of the interacting ligands, and the principle “better binding: stronger allostery” observed can be formalized in terms of linear free-energy relationships, which point to similar mechanism of the allosteric interaction between the enzyme-bound substrates and/or inhibitor molecules. On the other hand, the type of effect is governed by ligand binding effectiveness and can be inverted from positive allostery to negative allostery if we move from effectively binding ligands to badly binding compounds. Thus the outcome of the allostery in this monomeric enzyme is the same as defined by classical theories for multimeric enzymes: making the enzyme response more efficient if appropriate ligands bind.  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