共查询到20条相似文献,搜索用时 0 毫秒
1.
Patrick M. Collins Kazuya I. P. J. Hidari Helen Blanchard 《Acta Crystallographica. Section D, Structural Biology》2007,63(3):415-419
Galectin‐3 is a multifunctional carbohydrate‐binding protein that has roles in cancer progression. In addition to carbohydrate‐dependent extracellular functions, galectin‐3 participates in carbohydrate‐independent intracellular signalling pathways, including apoptosis, via protein–protein interactions, some of which engage the carbohydrate‐binding groove. When ligands bind within this site, conformational rearrangements are induced and information on unliganded galectin‐3 is therefore valuable for structure‐based drug design. Removal of cocrystallized lactose from the human galectin‐3 carbohydrate‐recognition domain was achieved via crystal soaking, but took weeks despite low affinity. Pre‐soaking to remove lactose enabled the subsequent binding of cryoprotectant glycerol, whereas when the lactose was not removed a priori the glycerol could not displace it in the short cryosoaking time frame. This slow diffusion of lactose out of the crystals contrasts with the entrance of glycerol, which takes place within minutes. The importance of the removal of incumbent ligands prior to attempts to introduce alternative ligands is indicated, even for proteins exhibiting low affinity for ligands, and has significance for ligand exchange in structure‐based drug design. 相似文献
2.
A method is described to dock a ligand into a binding site in a protein on the basis of the complementarity of the inter-molecular atomic contacts. Docking is performed by maximization of a complementarity function that is dependent on atomic contact surface area and the chemical properties of the contacting atoms. The generality and simplicity of the complementarity function ensure that a wide range of chemical structures can be handled. The ligand and the protein are treated as rigid bodies, but displacement of a small number of residues lining the ligand binding site can be taken into account. The method can assist in the design of improved ligands by indicating what changes in complementarity may occur as a result of the substitution of an atom in the ligand. The capabilities of the method are demonstrated by application to 14 protein–ligand complexes of known crystal structure. © 1996 Wiley Liss, Inc. 相似文献
3.
4.
H. Jhannesson V. P. Denisov B. Halle 《Protein science : a publication of the Protein Society》1997,6(8):1756-1763
The 2H magnetic relaxation dispersion (NMRD) technique was used to characterize interactions of dimethyl sulfoxide (DMSO) with globular proteins. A difference NMRD experiment involving the N-acetylglucosamine trisaccharide inhibitor, demonstrated that the DMSO 2H NMRD profile in lysozyme solution is due to a single DMSO molecule bound in the active cleft, with a molecular order parameter of 0.47 +/- 0.05 and a residence time in the range 10 ns to 5 ms. With the aid of transverse 2H relaxation data, the upper bound of the residence time was further reduced to 100 microns. A 1H shift titration experiment was also performed, yielding a binding constant of 2.3 +/- 0.3 M-1 at 27 degrees C. In contrast to lysozyme, no DMSO dispersion was observed for bovine pancreatic trypsin inhibitor (BPTI), indicating that a stable DMSO-protein complex requires a cleft of appropriate geometry in addition to hydrogen-bond and hydrophobic interactions. 相似文献
5.
The plant lectin ricin B chain binds oligosaccharide with more affinity than the mono- or disaccharide ligands. The experiments indicated that a biantennary oligosaccharide could bind itself to any of the crystallographically established 1st or 2nd binding sites. After manual docking of either terminal galactose residues of the oligosaccharide in the 1st and 2nd binding sites of Ricin B and simulating the systems over nanosecond trajectories in implicit solvent, it was observed that the protein bound the oligosaccharide strongly through both its 1st and 2nd binding sites. Not only were the terminal galactose residues, several other residues of the oligosaccharide were involved in the binding scheme. Average gas phase energies were calculated molecular mechanically, solvation energies were calculated by Generalized Born model and the normal mode analysis was used to calculate the entropic contribution of binding. The entropy/enthalpy compensation has been observed for the protein-oligosaccharide interactions. The binding was found to be enthalpically favorable and compensating for the unfavorable entropic contribution. Comparison of the calculated free energy with the experimental data clearly suggests that binding is mono-dentate rather than bi-dentate through a single Gal-containing antenna. 相似文献
6.
P. C. Sanschagrin L. A. Kuhn 《Protein science : a publication of the Protein Society》1998,7(10):2054-2064
Cluster analysis is presented as a technique for analyzing the conservation and chemistry of water sites from independent protein structures, and applied to thrombin, trypsin, and bovine pancreatic trypsin inhibitor (BPTI) to locate shared water sites, as well as those contributing to specificity. When several protein structures are superimposed, complete linkage cluster analysis provides an objective technique for resolving the continuum of overlaps between water sites into a set of maximally dense microclusters of overlapping water molecules, and also avoids reliance on any one structure as a reference. Water sites were clustered for ten superimposed thrombin structures, three trypsin structures, and four BPTI structures. For thrombin, 19% of the 708 microclusters, representing unique water sites, contained water molecules from at least half of the structures, and 4% contained waters from all 10. For trypsin, 77% of the 106 microclusters contained water sites from at least half of the structures, and 57% contained waters from all three. Water site conservation correlated with several environmental features: highly conserved microclusters generally had more protein atom neighbors, were in a more hydrophilic environment, made more hydrogen bonds to the protein, and were less mobile. There were significant overlaps between thrombin and trypsin conserved water sites, which did not localize to their similar active sites, but were concentrated in buried regions including the solvent channel surrounding the Na+ site in thrombin, which is associated with ligand selectivity. Cluster analysis also identified water sites conserved in thrombin but not trypsin, and vice versa, providing a list of water sites that may contribute to ligand discrimination. Thus, in addition to facilitating the analysis of water sites from multiple structures, cluster analysis provides a useful tool for distinguishing between conserved features within a protein family and those conferring specificity. 相似文献
7.
Pellecchia M Meininger D Dong Q Chang E Jack R Sem DS 《Journal of biomolecular NMR》2002,22(2):165-173
8.
9.
V. Chu S. Freitag I. Le Trong R. E. Stenkamp P. S. Stayton 《Protein science : a publication of the Protein Society》1998,7(4):848-859
A circularly permuted streptavidin (CP51/46) has been designed to remove the flexible polypeptide loop that undergoes an open to closed conformational change when biotin is bound. The original termini have been joined by a tetrapeptide linker, and four loop residues have been removed, resulting in the creation of new N- and C-termini. Isothermal titration calorimetric studies show that the association constant has been reduced approximately six orders of magnitude below that of wild-type streptavidin to 10(7) M(-1). The deltaH degrees of biotin association for CP51/46 is reduced by 11.1 kcal/mol. Crystal structures of CP51/46 and its biotin complex show no significant alterations in the binding site upon removal of the loop. A hydrogen bond between Ser45 and Ser52 found in the absence of biotin is broken in the closed conformation as the side-chain hydroxyl of Ser45 moves to hydrogen bond to a ureido nitrogen of biotin. This is true in both the wild-type and CP51/46 forms of the protein, and the hydrogen bonding interaction might thus help nucleate closure of the loop. The reduced entropic cost of binding biotin to CP51/46 is consistent with the removal of this loop and a reduction in entropic costs associated with loop closure and immobilization. The reduced enthalpic contribution to the free energy of binding is not readily explainable in terms of the molecular structure, as the binding contacts are nearly entirely conserved, and only small differences in solvent accessible surfaces are observed relative to wild-type streptavidin. 相似文献
10.
11.
Ronald M. Levy Wei Dai Nan‐Jie Deng Dmitrii E. Makarov 《Protein science : a publication of the Protein Society》2013,22(11):1459-1465
How long does it take to equilibrate the unfolded state of a protein? The answer to this question has important implications for our understanding of why many small proteins fold with two state kinetics. When the equilibration within the unfolded state U is much faster than the folding, the folding kinetics will be two state even if there are many folding pathways with different barriers. Yet the mean first passage times (MFPTs) between different regions of the unfolded state can be much longer than the folding time. This seems to imply that the equilibration within U is much slower than the folding. In this communication we resolve this paradox. We present a formula for estimating the time to equilibrate the unfolded state of a protein. We also present a formula for the MFPT to any state within U, which is proportional to the average lifetime of that state divided by the state population. This relation is valid when the equilibration within U is very fast as compared with folding as it often is for small proteins. To illustrate the concepts, we apply the formulas to estimate the time to equilibrate the unfolded state of Trp-cage and MFPTs within the unfolded state based on a Markov State Model using an ultra-long 208 microsecond trajectory of the miniprotein to parameterize the model. The time to equilibrate the unfolded state of Trp-cage is ∼100 ns while the typical MFPTs within U are tens of microseconds or longer. 相似文献
12.
We have performed molecular dynamics simulations on a set of nine unfolded conformations of the fastest-folding protein yet discovered, a variant of the villin headpiece subdomain (HP-35 NleNle). The simulations were generated using a new distributed computing method, yielding hundreds of trajectories each on a time scale comparable to the experimental folding time, despite the large (10,000 atom) size of the simulation system. This strategy eliminates the need to assume a two-state kinetic model or to build a Markov state model. The relaxation to the folded state at 300 K from the unfolded configurations (generated by simulation at 373 K) was monitored by a method intended to reflect the experimental observable (quenching of tryptophan by histidine). We also monitored the relaxation to the native state by directly comparing structural snapshots with the native state. The rate of relaxation to the native state and the number of resolvable kinetic time scales both depend upon starting structure. Moreover, starting structures with folding rates most similar to experiment show some native-like structure in the N-terminal helix (helix 1) and the phenylalanine residues constituting the hydrophobic core, suggesting that these elements may exist in the experimentally relevant unfolded state. Our large-scale simulation data reveal kinetic complexity not resolved in the experimental data. Based on these findings, we propose additional experiments to further probe the kinetics of villin folding. 相似文献
13.
Jenni Leppiniemi Amit Meir Niklas Kähkönen Sampo Kukkurainen Juha A. Määttä Markus Ojanen Janne Jänis Markku S. Kulomaa Oded Livnah Vesa P. Hytönen 《Protein science : a publication of the Protein Society》2013,22(7):980-994
Bradavidin II is a biotin‐binding protein from Bradyrhizobium japonicum that resembles chicken avidin and bacterial streptavidin. A biophysical characterization was carried out using dynamic light scattering, native mass spectrometry, differential scanning calorimetry, and isothermal titration calorimetry combined with structural characterization using X‐ray crystallography. These observations revealed that bradavidin II differs from canonical homotetrameric avidin protein family members in its quaternary structure. In contrast with the other avidins, bradavidin II appears to have a dynamic (transient) oligomeric state in solution. It is monomeric at low protein concentrations but forms higher oligomeric assemblies at higher concentrations. The crystal structure of bradavidin II revealed an important role for Phe42 in shielding the bound ligand from surrounding water molecules, thus functionally replacing the L7,8 loop essential for tight ligand binding in avidin and streptavidin. This bradavidin II characterization opens new avenues for oligomerization‐independent biotin‐binding protein development. 相似文献
14.
Noriaki Okimoto Atsushi Suenaga Makoto Taiji 《Journal of biomolecular structure & dynamics》2017,35(15):3221-3231
In computational drug design, ranking a series of compound analogs in a manner that is consistent with experimental affinities remains a challenge. In this study, we evaluated the prediction of protein–ligand binding affinities using steered molecular dynamics simulations. First, we investigated the appropriate conditions for accurate predictions in these simulations. A conic harmonic restraint was applied to the system for efficient sampling of work values on the ligand unbinding pathway. We found that pulling velocity significantly influenced affinity predictions, but that the number of collectable trajectories was less influential. We identified the appropriate pulling velocity and collectable trajectories for binding affinity predictions as 1.25 Å/ns and 100, respectively, and these parameters were used to evaluate three target proteins (FK506 binding protein, trypsin, and cyclin-dependent kinase 2). For these proteins using our parameters, the accuracy of affinity prediction was higher and more stable when Jarzynski’s equality was employed compared with the second-order cumulant expansion equation of Jarzynski’s equality. Our results showed that steered molecular dynamics simulations are effective for predicting the rank order of ligands; thus, they are a potential tool for compound selection in hit-to-lead and lead optimization processes. 相似文献
15.
T. G. Davies R. E. Hubbard J. R. Tame 《Protein science : a publication of the Protein Society》1999,8(7):1432-1444
The oligopeptide-binding protein OppA provides a useful model system for studying the physical chemistry underlying noncovalent interactions since it binds a variety of readily synthesized ligands. We have studied the binding of eight closely related tripeptides of the type Lysine-X-Lysine, where X is an abnormal amino acid, by isothermal titration calorimetry (ITC) and X-ray crystallography. The tripeptides fall into three series of ligands, which have been designed to examine the effects of small changes to the central side chain. Three ligands have a primary amine as the second side chain, two have a straight alkane chain, and three have ring systems. The results have revealed a definite preference for the binding of hydrophobic residues over the positively charged side chains, the latter binding only weakly due to unfavorable enthalpic effects. Within the series of positively charged groups, a point of lowest affinity has been identified and this is proposed to arise from unfavorable electrostatic interactions in the pocket, including the disruption of a key salt bridge. Marked entropy-enthalpy compensation is found across the series, and some of the difficulties in designing tightly binding ligands have been highlighted. 相似文献
16.
Aisha Laguerre Jerome Wielens Michael W. Parker Christopher J. H. Porter Martin J. Scanlon 《Acta Crystallographica. Section F, Structural Biology Communications》2011,67(2):291-295
Fatty‐acid binding proteins (FABPs) are abundantly expressed proteins that bind a range of lipophilic molecules. They have been implicated in the import and intracellular distribution of their ligands and have been linked with metabolic and inflammatory responses in the cells in which they are expressed. Despite their high sequence identity, human intestinal FABP (hIFABP) and rat intestinal FABP (rIFABP) bind some ligands with different affinities. In order to address the structural basis of this differential binding, diffraction‐quality crystals have been obtained of hIFABP and rIFABP in complex with the fluorescent fatty‐acid analogue 11‐(dansylamino)undecanoic acid. 相似文献
17.
The characteristics of heme prosthetic groups and their binding sites have been analyzed in detail in a data set of nonhomologous heme proteins. Variations in the shape, volume, and chemical composition of the binding site, in the mode of heme binding and in the number and nature of heme–protein interactions are found to result in significantly different heme environments in proteins with different functions in biology. Differences are also seen in the properties of the apo states of the proteins. The apo states of proteins that bind heme permanently in their functional form show some disorder, ranging from local unfolding in the heme binding pocket to complete unfolding to give a random coil. In contrast, proteins that bind heme transiently are fully folded in their apo and holo states, presumably allowing both apo and holo forms to remain biologically active resisting aggregation or proteolysis. The principles identified here provide a framework for the design of de novo proteins that will exhibit tight heme ligand binding and for the identification of the function of structural genomic target proteins with heme ligands. Proteins 2010. © 2010 Wiley‐Liss, Inc. 相似文献
18.
George Kontopidis Paul Taylor Malcolm D. Walkinshaw 《Acta Crystallographica. Section D, Structural Biology》2004,60(3):479-485
Piperidine ligands are described that provide the first examples of non‐peptidic ligand structures for the cyclophilin family of proteins. Crystal structures of two ligand complexes are compared with the unliganded protein and show ligand‐induced changes in side‐chain conformation and water binding. A peptidylprolyl cis–trans‐isomerase assay showed the dissociation constants of the two ligands to be 320 and 25 mM. This study also provides the first published data for both enzymatic activity and three‐dimensional structure for any protein–ligand complex that binds with a high‐millimolar dissociation constant. The structures may be of relevance in the field of drug design, as they suggest starting points for the design of larger tighter‐binding analogues. 相似文献
19.
Tschapek B Pittelkow M Sohn-Bösser L Holtmann G Smits SH Gohlke H Bremer E Schmitt L 《Journal of molecular biology》2011,411(1):36-67
The substrate binding protein AfProX from the Archaeoglobus fulgidus ProU ATP binding cassette transporter is highly selective for the compatible solutes glycine betaine (GB) and proline betaine, which confer thermoprotection to this hyperthermophilic archaeon. A detailed mutational analysis of the substrate binding site revealed the contribution of individual amino acids for ligand binding. Replacement of Arg149 by an Ala residue displayed the largest impact on substrate binding. The structure of a mutant AfProX protein (substitution of Tyr111 with Ala) in complex with GB was solved in the open liganded conformation to gain further insight into ligand binding. In this crystal structure, GB is bound differently compared to the GB closed liganded structure of the wild-type AfProX protein. We found that a network of amino acid side chains communicates the presence of GB toward Arg149, which increases ligand affinity and induces domain closure of AfProX. These results were corroborated by molecular dynamics studies and support the view that Arg149 finalizes the high-affinity state of the AfProX substrate binding protein. 相似文献