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.
Differential interactions among nonpolar moieties at protein/ligand interfaces, and of these nonpolar groups with water, collectively termed hydrophobic interactions, are widely believed to make important energetic contributions to the stability of protein/ligand complexes. Quantitative estimates of hydrophobic interactions, and an evaluation of their structural basis, are essential for obtaining structure-based predictions of the free energies of binding for the purpose of drug design. Two largely nonpolar, immunosuppressive agents, FK506 and rapamycin, each bind with high affinity to a common hydrophobic pocket on a small peptidylproline cis-trans isomerase known as FK506 binding protein (FKBP-12) and inhibit its activity. In an effort to elucidate the structural features of these ligands responsible for the observed energetics, we have undertaken an investigation of the thermodynamics of binding of FK506 and rapamycin to FKBP-12. Enthalpies of binding have been determined by high-precision titration calorimetry over a range of temperature, allowing estimates of heat capacity changes. By analyzing the distribution of changes in solvent-accessible surface area upon binding of FK506 to FKBP-12 from crystallographic data, it is found that 99% of the net surface buried upon binding involves nonpolar groups. This leads to a heat capacity change of FK506 binding, normalized to the amount of nonpolar surface, of -0.40 +/- 0.02 cal.K-1.mol-1.A-2 (1 cal = 4.18 J), a value similar to that obtained for the aqueous dissolution of hydrophobic substances. Our observations are discussed in view of the general nature of hydrophobic interaction processes.  相似文献   

2.
The binding of charged ligands benzamidine and diazamidine to trypsin was investigated by using a polarizable potential energy function and explicit-water molecular dynamics simulations. The binding free energies were computed from the difference between the free energies of decoupling the ligand from water and protein environments. Both the absolute and the relative free energies from the perturbation simulations agree with experimental measurements to within 0.5 kcal.mol(-1). Comparison of free-energy components sampled from different thermodynamic paths indicates that electrostatics is the main driving force behind benzamidine recognition of trypsin. The contribution of electronic polarization to binding appears to be crucial. By computing the free-energy contribution caused by the polarization between the ligand and its surroundings, we found that polarization has the opposite effect in dissimilar environments. Although polarization favors ligand solvation in water, it weakens the protein-ligand attraction by screening the electrostatic interaction between trypsin and benzamidine. We also examined the relative binding free energies of a benzamidine analog diazamidine to trypsin. The changes in free energy on benzamidine-diazamidine substitution were tens of kilocalories in both water and trypsin environments; however, the change in the total binding free energy is <2 kcal.mol(-1) because of cancellation, consistent with the experimental results. Overall, our results suggest that the use of a polarizable force field, given adequate sampling, is capable of achieving chemical accuracy in molecular simulations of protein-ligand recognition.  相似文献   

3.
The hydrophobic effect--a rationalization of the insolubility of nonpolar molecules in water--is centrally important to biomolecular recognition. Despite extensive research devoted to the hydrophobic effect, its molecular mechanisms remain controversial, and there are still no reliably predictive models for its role in protein-ligand binding. Here we describe a particularly well-defined system of protein and ligands--carbonic anhydrase and a series of structurally homologous heterocyclic aromatic sulfonamides--that we use to characterize hydrophobic interactions thermodynamically and structurally. In binding to this structurally rigid protein, a set of ligands (also defined to be structurally rigid) shows the expected gain in binding free energy as hydrophobic surface area is added. Isothermal titration calorimetry demonstrates that enthalpy determines these increases in binding affinity, and that changes in the heat capacity of binding are negative. X-ray crystallography and molecular dynamics simulations are compatible with the proposal that the differences in binding between the homologous ligands stem from changes in the number and organization of water molecules localized in the active site in the bound complexes, rather than (or perhaps in addition to) release of structured water from the apposed hydrophobic surfaces. These results support the hypothesis that structured water molecules--including both the molecules of water displaced by the ligands and those reorganized upon ligand binding--determine the thermodynamics of binding of these ligands at the active site of the protein. Hydrophobic effects in various contexts have different structural and thermodynamic origins, although all may be manifestations of the differences in characteristics of bulk water and water close to hydrophobic surfaces.  相似文献   

4.
The thermodynamic properties and phase behavior of water in confined regions can vary significantly from that observed in the bulk. This is particularly true for systems in which the confinement is on the molecular-length scale. In this study, we use molecular dynamics simulations and a powerful solvent analysis technique based on inhomogenous solvation theory to investigate the properties of water molecules that solvate the confined regions of protein active sites. Our simulations and analysis indicate that the solvation of protein active sites that are characterized by hydrophobic enclosure and correlated hydrogen bonds induce atypical entropic and enthalpic penalties of hydration. These penalties apparently stabilize the protein-ligand complex with respect to the independently solvated ligand and protein, which leads to enhanced binding affinities. Our analysis elucidates several challenging cases, including the super affinity of the streptavidin-biotin system.  相似文献   

5.
The synthetic, biologically active, [Phe4Val6] analog of natural antamanide has been crystallized from a solution containing calcium nitrate, acetone, and acetonitrile. The crystal does not contain any Ca2+ ions but does contain 12 water molecules per peptide molecule. The conformation of this dodecahydrate is identical to the trihydrate crystallized from n-hexane/methyl acetate. The packing in both crystals is very similar, governed by parallel bands of phenyl and pyrrolidine ring stacking and by continuous channels for the solvement molecules, water in this case and n-hexane/methyl acetate in the previous study. The two structures are not ideally isomorphous, since the c cell parameter differs by more than 1.0 A in the two cells. There are three functions for H2O molecules: the three intrinsic H2O molecules in the interior of the peptide ring, the H2O bound to the exposed C=O groups, and the H2O molecules that fill space in the solvent channels. There are no direct hydrogen bonds between neighboring peptide molecules and there are only two intramolecular NH-O=C bonds (of the 5 leads to 1 type).  相似文献   

6.
We report the three-dimensional structures, at 1.8-A resolution, of the Fv fragment of the anti-hen egg white lysozyme antibody D1.3 in its free and antigen-bound forms. These structures reveal a role for solvent molecules in stabilizing the complex and provide a molecular basis for understanding the thermodynamic forces which drive the association reaction. Four water molecules are buried and others form a hydrogen-bonded network around the interface, bridging antigen and antibody. Comparison of the structures of free and bound Fv fragment of D1.3 reveals that several of the ordered water molecules in the free antibody combining site are retained and that additional water molecules link antigen and antibody upon complex formation. This solvation of the complex should weaken the hydrophobic effect, and the resulting large number of solvent-mediated hydrogen bonds, in conjunction with direct protein-protein interactions, should generate a significant enthalpic component. Furthermore, a stabilization of the relative mobilities of the antibody heavy- and light-chain variable domains and of that of the third complementarity-determining loop of the heavy chain seen in the complex should generate a negative entropic contribution opposing the enthalpic and the hydrophobic (solvent entropy) effects. This structural analysis is consistent with measurements of enthalpy and entropy changes by titration calorimetry, which show that enthalpy drives the antigen-antibody reaction. Thus, the main forces stabilizing the complex arise from antigen-antibody hydrogen bonding, van der Waals interactions, enthalpy of hydration, and conformational stabilization rather than solvent entropy (hydrophobic) effects.  相似文献   

7.
Efficient protein folding implies a microscopic funnel-like multidimensional free-energy landscape. Macroscopically, conformational entropy reduction can manifest itself as part of an empirical barrier in the traditional view of folding, but experiments show that such barriers can also entail significant unfavorable enthalpy changes. This observation raises the puzzling possibility, irrespective of conformational entropy, that individual microscopic folding trajectories may encounter large uphill moves and thus the multidimensional free-energy landscape may not be funnel-like. Here, we investigate how nanoscale hydrophobic interactions might underpin this salient enthalpic effect in biomolecular assembly by computer simulations of the association of two preformed polyalanine or polyleucine helices in water. We observe a high, positive enthalpic signature at room temperature when the helix separation is less than a single layer of water molecules. Remarkably, this unfavorable enthalpy change, with a parallel increase in void volume, is largely compensated for by a concomitant increase in solvent entropy, netting only a small or nonexistent microscopic free-energy barrier. Thus, our findings suggest that high enthalpic folding barriers can be consistent with a funnel picture of folding and are mainly a desolvation phenomenon indicative of a cooperative mechanism of simultaneous formation of multiple side-chain contacts at the rate-limiting step.  相似文献   

8.
Stimulated by the recent surprising results from Wentworth et al. [Wentworth, A. D., Jones, L. H., Wentworth, P., Janda, K. D. & Lerner, R. A. (2000) Proc. Natl. Acad. Sci. USA 97, 10930-10935] that Abs efficiently catalyze the conversion of molecular singlet oxygen ((1)O(2)) plus water to hydrogen peroxide (HOOH), we used quantum chemical methods (B3LYP density functional theory) to delineate the most plausible mechanisms for the observed efficient conversion of water to HOOH. We find two reasonable pathways. In Pathway I, (i) H(2)O catalyzes the reaction of (1)O(2) with a second water to form HOOOH; (ii) two HOOOH form a dimer, which rearranges to form the HOO-HOOO + H(2)O complex; (iii) HOO-HOOO rearranges to HOOH-OOO, which subsequently reacts with H(2)O to form H(2)O(4) + HOOH; and (iv) H(2)O(4) rearranges to the cyclic dimer (HO(2))(2), which in turn forms HOOH plus (1)O(2) or (3)O(2). Pathway II differs in that step ii is replaced with the reaction between HOOOH and (1)O(2), leading to the formation of HOO-HOOO. This then proceeds to similar products. For a system with (18)O H(2)O, Pathway I leads to a 2.2:1 ratio of (16)O:(18)O in the product HOOH, whereas Pathway II leads to 3:1. These ratios are in good agreement with the 2.2:1 ratio observed in isotope experiments by Wentworth et al. These mechanisms lead to two HOOH per initial (1)O(2) or one, depending on whether the product of step iv is (1)O(2) or (3)O(2), in good agreement with the experimental result of 2.0. In addition to the Ab-induced reactions, the hydrogen polyoxides (H(2)O(3) and H(2)O(4)) formed in these mechanisms and their decomposition product polyoxide radicals (HO(2), HO(3)) may play a role in combustion, explosions, atmospheric chemistry, and the radiation chemistry in aqueous systems.  相似文献   

9.
An equation of state describing hydrophobic interactions   总被引:10,自引:5,他引:5       下载免费PDF全文
Several thermodynamic properties for the process of dissolution of pure hydrocarbons into water are found to be linearly related to the number of hydrogens on the hydrocarbon molecule. From the correlations found for the Gibbs energy change, enthalpy change, and heat capacity change, along with the use of an average minimum solubility temperature, an equation of state for the hydrophobic effect is derived. The entropy change upon dissolution per hydrocarbon hydrogen atom is close to -R ln 2. A model based upon a “tetrahedrally” localized water molecule with one corner defined by a carbon-hydrogen group and the other three corners defined by water molecules is used to estimate the observed entropy and heat capacity changes.  相似文献   

10.
Wentworth et al. [Wentworth, P., Jones, L. H., Wentworth, A. D., Zhu, X. Y., Larsen, N. A., Wilson, I. A., Xu, X., Goddard, W. A., Janda, K. D., Eschenmoser, A. & Lerner, R. A. (2001) Science 293, 1806-1811] recently reported the surprising result that antibodies and T cell receptors efficiently catalyze the conversion of molecular singlet oxygen (1O2) plus water to hydrogen peroxide (HOOH). Recently, quantum mechanical calculations were used to delineate a plausible mechanism, involving reaction of 1O2 with two waters to form HOOOH (plus H2O), followed by formation of HOOOH dimer, which rearranges to form HOO-HOOO + H2O, which rearranges to form two HOOH plus 1O2 or 3O2. For a system with 18O H2O, this mechanism leads to a 2.2:1 ratio of 16O:18O in the product HOOH, in good agreement with the ratio 2.2:1 observed in isotope experiments by Wentworth et al. In this paper we use docking and molecular dynamics techniques (HierDock) to search various protein structures for sites that stabilize these products and intermediates predicted from quantum mechanical calculations. We find that the reaction intermediates for production of HOOH from 1O2 are stabilized at the interface of light and heavy chains of antibodies and T cell receptors. This inter Greek key domain interface structure is unique to antibodies and T cell receptors, but is not present in beta2-microglobulin, which does not show any stabilization in our docking studies. This result is consistent with the experimentally observed lack of HOOH production in this system. Our results provide a plausible mechanism for the reactions and provide an explanation of the specific structural character of antibodies responsible for this unexpected chemistry.  相似文献   

11.
We theorize that intoxicants and modern anesthetics bind at the membrane-water interface and displace (dehydrate) bound water molecules by breaking the hydrogen bonds. We tested this hypothesis by examining the effect of butanol on the binding of water to the polar regions of lipids in reversed micelles. Understanding the mechanisms of intoxication requires studies in physiologically relevant systems such as systems containing sialoglycoconjugates, especially gangliosides, which concentrate in the synapses of neural tissue. Therefore, we compared butanol effects on phospholipid with effects on ganglioside. Hydrogen-bond breaking activity of 1-butanol was studied in reversed micelles made of dipalmitoylphosphotidylcholine (DPPC), ganglioside (GM1 and GT1b) or the lipid mixture in a D2O-CCl4 medium. Fourier transform infrared spectroscopy (FTIR) data indicated that 1-butanol binds to DPPC and to gangliosides. Adding GM1 to the DPPC micelles introduces a new binding site for the alcohol. GT1b binds more butanol than GM1, because of more binding sites provided by extra sialic acid moieties. Spectral red shifts indicate that both water and butanol bind to the C = O group of sialic acid. Butanol partially releases the surface-bound water by disrupting hydrogen bonds, as indicated by an appearance of a sharp new free OD stretching band of the released D2O molecules. However, control studies with lipid-free systems in CCl4 revealed that a free OD peak could occur from a deuterium exchange reaction between D2O and 1-butanol(ol-h).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The 12-kDa FK506-binding protein (FKBP-12) is a cytosolic receptor for the immunosuppressants FK506 and rapamycin. Here we report the molecular cloning and subcellular localization of a 13-kDa FKBP (FKBP-13), which has a 21-amino acid signal peptide and appears to be membrane-associated. Although no internal hydrophobic region, and thus no transmembrane domain, is apparent within the 120 amino acids of mature FKBP-13, a potential endoplasmic reticulum retention sequence (Arg-Thr-Glu-Leu) is found at its C terminus. FKBP-13 has 51% nucleotide sequence identity and 43% amino acid sequence identity to FKBP-12; the N-terminal sequences are divergent, but the 92-amino acid C-terminal sequence of FKBP-13 has 46 identical and 20 related residues when compared with FKBP-12. The conserved residues that comprise the drug binding site and rotamase active site of FKBP-12 are completely conserved in FKBP-13. Therefore, the three-dimensional structures of FKBP-12 and the FKBP-12/FK506 complex are likely to be excellent models of the corresponding FKBP-13 structure.  相似文献   

13.
Summary Arguments are presented for a unified theory of insulin binding which differs from presently accepted viewpoints: (1) Insulin receptors are bivalent with some motional freedom between binding units; (2) following ligand formation a conformational change occurs which leads to a restriction of receptor flexibility; (3) when both binding units are occupied, they can approach sufficiently closely with an antiparallel symmetry to permit dimer formation-related interaction of receptor-bound insulin molecules; (4) this event leads to enhanced dissociation (negative cooperativity) with kinetic features identical to the model of de Meyts; (5) the binding surface of insulin cannot include residues involved in dimer formation. If this is so, then the accepted receptor binding surface of insulin, which includes the residues involved in dimer formation, is incorrect. Arguments are brought suggesting that strongly conserved hydrophilic residues of the A chain may be involved.  相似文献   

14.
Binding of ligands to the active site of carboxypeptidase A.   总被引:7,自引:5,他引:2       下载免费PDF全文
We compare the detailed binding modes of the 39-amino acid inhibitor from potatoes, glycyl-L-tyrosine, the ester analogue CH3OC6H4(CO)CH2CH(CO2(-))C6H5, and indole acetate to the exopeptidase carboxypeptidase A (EC 3.4.17.1). In the potato inhibitor, cleavage of the COOH-terminal glycine-39 leaves a new carboxylate anion of valine-38 having one oxygen on zinc and the other as a receptor of a hydrogen bond from tyrosine-248 of carboxypeptidase. Tyrosine-248 also receives a hydrogen bond from the amide proton of the originally penultimate peptide bond between tyrosine-37 and valine-38. This hydrogen bond suggests product stabilization which is available to peptides and depsipeptides but not to esters lacking an equivalent peptide bond (nonspecific esters). Also, this structure may represent the intermediate binding step for the uncleaved substrate as it moves along the binding subsites. In particular, this may be the binding mode for the substrate after association of the COOH-terminal region of the substrate with the residues at binding subsite S2 (tyrosine-198, phenylalanine-279, and arginine-71) and preceding entry into the catalytic site S1'. These stabilized complexes allow some understanding of the effect of indole acetate, shown here to bind in the pocket at S1', as a competitive inhibitor for esters (for which entry into S1' precedes the rate-determining catalytic step for hydrolysis) and as a noncompetitive inhibitor for peptides (for which entry into S1' is rate limiting). These results, including the binding mode of the ester analogue, are consistent with the original proposal from x-ray studies that both esters and peptides are cleaved with the carboxy terminus at S1', although not necessarily by the same chemical steps.  相似文献   

15.
Photoreduction of [P(2)W(18)O(62)](6-), [S(2)Mo(18)O(62)](4-), and [S(2)W(18)O(62)](4-) polyoxometalate anions (POMs) and oxidation of water occurs when water-ionic liquid and water-diethylether interfaces are irradiated with white light (275-750 nm) or sunlight. The ionic liquids (ILs) employed were aprotic ([Bmim]X; Bmim = (1-butyl-3-methylimidazolium, X = BF(4), PF(6)) and protic (DEAS = diethanolamine hydrogen sulphate; DEAP = diethanolamine hydrogen phosphate). Photochemical formation of reduced POMs at both thermodynamically stable and unstable water-IL interfaces led to their initial diffusion into the aqueous phase and subsequent extraction into the IL phase. The mass transport was monitored visually by color change and by steady-state voltammetry at microelectrodes placed near the interface and in the bulk solution phases. However, no diffusion into the organic phase was observed when [P(2)W(18)O(62)](6-) was photo-reduced at the water-diethylether interface. In all cases, water acted as the electron donor to give the overall process: 4POM + 2H(2)O + hν → 4POM(-) + 4H(+) + O(2). However, more highly reduced POM species are likely to be generated as intermediates. The rate of diffusion of photo-generated POM(-) was dependent on the initial concentration of oxidized POM and the viscosity of the IL (or mixed phase system produced in cases in which the interface is thermodynamically unstable). In the water-DEAS system, the evolution of dioxygen was monitored in situ in the aqueous phase by using a Clark-type oxygen sensor. Differences in the structures of bulk and interfacial water are implicated in the activation of water. An analogous series of reactions occurred upon irradiation of solid POM salts in the presence of water vapor.  相似文献   

16.
We report the fabrication of enthalpy arrays and their use to detect molecular interactions, including protein-ligand binding, enzymatic turnover, and mitochondrial respiration. Enthalpy arrays provide a universal assay methodology with no need for specific assay development such as fluorescent labeling or immobilization of reagents, which can adversely affect the interaction. Microscale technology enables the fabrication of 96-detector enthalpy arrays on large substrates. The reduction in scale results in large decreases in both the sample quantity and the measurement time compared with conventional microcalorimetry. We demonstrate the utility of the enthalpy arrays by showing measurements for two protein-ligand binding interactions (RNase A + cytidine 2'-monophosphate and streptavidin + biotin), phosphorylation of glucose by hexokinase, and respiration of mitochondria in the presence of 2,4-dinitrophenol uncoupler.  相似文献   

17.
The stability of the recently characterized type II hydrogen clathrate [Mao, W. L., Mao, H.-K., Goncharov, A. F., Struzhkin, V. V., Guo, Q., et al. (2002) Science 297, 2247-2249] with respect to hydrogen occupancy is examined with a statistical mechanical model in conjunction with first-principles quantum chemistry calculations. It is found that the stability of the clathrate is mainly caused by dispersive interactions between H2 molecules and the water forming the cage walls. Theoretical analysis shows that both individual hydrogen molecules and nH2 guest clusters undergo essentially free rotations inside the clathrate cages. Calculations at the experimental conditions--2,000 bar (1 bar = 100 kPa) and 250 K confirm multiple occupancy of the clathrate cages with average occupations of 2.00 and 3.96 H2 molecules per D-5(12) (small) and H-5(12)6(4) (large) cage, respectively. The H2-H2O interactions also are responsible for the experimentally observed softening of the H[bond]H stretching modes. The clathrate is found to be thermodynamically stable at 25 bar and 150 K.  相似文献   

18.
Biological processes often depend on protein-ligand binding events, yet accurate calculation of the associated energetics remains as a significant challenge of central importance to structure-based drug design. Recently, we have proposed that the displacement of unfavorable waters by the ligand, replacing them with groups complementary to the protein surface, is the principal driving force for protein-ligand binding, and we have introduced the WaterMap method to account this effect. However, in spite of the adage "nature abhors vacuum," one can occasionally observe situations in which a portion of the receptor active site is so unfavorable for water molecules that a void is formed there. In this paper, we demonstrate that the presence of dry regions in the receptor has a nontrivial effect on ligand binding affinity, and suggest that such regions may represent a general motif for molecular recognition between the dry region in the receptor and the hydrophobic groups in the ligands. With the introduction of a term attributable to the occupation of the dry regions by ligand atoms, combined with the WaterMap calculation, we obtain excellent agreement with experiment for the prediction of relative binding affinities for a number of congeneric ligand series binding to the major urinary protein receptor. In addition, WaterMap when combined with the cavity contribution is more predictive than at least one specific implementation [Abel R, Young T, Farid R, Berne BJ, Friesner RA (2008) J Am Chem Soc 130:2817-2831] of the popular MM-GBSA approach to binding affinity calculation.  相似文献   

19.
Proton transfer is crucial for many enzyme reactions. Here, we show that in addition to protonatable amino acid side chains, water networks could constitute proton-binding sites in proteins. A broad IR continuum absorbance change during the proton pumping photocycle of bacteriorhodopsin (bR) indicates most likely deprotonation of a protonated water cluster at the proton release site close to the surface. We investigate the influence of several mutations on the proton release network and the continuum change, to gain information about the location and extent of the protonated water network and to reveal the participating residues necessary for its stabilization. We identify a protonated water cluster consisting in total of one proton and about five water molecules surrounded by six side chains and three backbone groups (Tyr-57, Arg-82, Tyr-83, Glu-204, Glu-194, Ser-193, Pro-77, Tyr-79, and Thr-205). The observed perturbation of proton release by many single-residue mutations is now explained by the influence of numerous side chains on the protonated H bonded network. In situ hydrogen/deuterium exchange Fourier transform IR measurements of the bR ground state, show that the proton of the release group becomes localized on Glu-204 and Asp-204 in the ground state of the mutants E194D and E204D, respectively, even though it is delocalized in the ground state of wild-type bR. Thus, the release mechanism switches between the wild-type and mutated proteins from a delocalized to a localized proton-binding site.  相似文献   

20.
The ligand binding domain of the human vitamin D receptor (VDR) was modeled based on the crystal structure of the retinoic acid receptor. The ligand binding pocket of our VDR model is spacious at the helix 11 site and confined at the beta-turn site. The ligand 1alpha, 25-dihydroxyvitamin D(3) was assumed to be anchored in the ligand binding pocket with its side chain heading to helix 11 (site 2) and the A-ring toward the beta-turn (site 1). Three residues forming hydrogen bonds with the functionally important 1alpha- and 25-hydroxyl groups of 1alpha,25-dihydroxyvitamin D(3) were identified and confirmed by mutational analysis: the 1alpha-hydroxyl group is forming pincer-type hydrogen bonds with S237 and R274 and the 25-hydroxyl group is interacting with H397. Docking potential for various ligands to the VDR model was examined, and the results are in good agreement with our previous three-dimensional structure-function theory.  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