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 15 毫秒
1.
We investigated the temperature and pressure effects on conformational equilibria of N-acetyl-L-alanine-N'-methylamide (AAlaMA) in aqueous solution by Raman spectroscopy. Scattering intensities in the skeletal stretching mode of AAlaMA in aqueous solution were decomposed into some component bands by the spectra analysis. Our results indicate that each component band for AAlaMA adopts not only the P(II) and alpha(R) conformations but also the C(7eq) conformation. From temperature and pressure dependencies of the band intensities, we determined the enthalpy differences and the volume differences between the conformers. The C(7eq) conformer is enthalpically most stable due to the intramolecular hydrogen bond. The partial molar volume of the C(7eq) conformer is the smallest through the solvent-exclusion effect rather than the solute-solvent electrostatic interaction effect.  相似文献   

2.
In this paper we discuss the problem of including solvation free energies in evaluating the relative stabilities of loops in proteins. A conformational search based on a gas-phase potential function is used to generate a large number of trial conformations. As has been found previously, the energy minimization step in this process tends to pack charged and polar side chains against the protein surface, resulting in conformations which are unstable in the aqueous phase. Various solvation models can easily identify such structures. In order to provide a more severe test of solvation models, gas phase conformations were generated in which side chains were kept extended so as to maximize their interaction with the solvent. The free energies of these conformations were compared to that calculated for the crystal structure in three loops of the protein E. coli RNase H, with lengths of 7, 8, and 9 residues. Free energies were evaluated with a finite difference Poisson-Boltzmann (FDPB) calculation for electrostatics and a surface area-based term for nonpolar contributions. These were added to a gas-phase potential function. A free energy function based on atomic solvation parameters was also tested. Both functions were quite successful in selecting, based on a free energy criterion, conformations quite close to the crystal structure for two of the three loops. For one loop, which is involved in crystal contacts, conformations that are quite different from the crystal structure were also selected. A method to avoid precision problems associated with using the FDPB method to evaluate conformational free energies in proteins is described. © 1994 John Wiley & Sons, Inc.  相似文献   

3.
The conformational preferences of blocked alanine dipeptide (ADP), Ac‐Ala‐NHMe, in aqueous solution were studied using vibrational circular dichroism (VCD) together with density functional theory (DFT) calculations. DFT calculations of three most representative conformations of ADP surrounded by six explicit water molecules immersed in a dielectric continuum have proven high sensitivity of amide III VCD band shape that is characteristic for each conformation of the peptide backbone. The polyproline II (PII) and αR conformation of ADP are associated with a positive VCD band while β conformation has a negative VCD band in amide III region. Knowing this spectral characteristic of each conformation allows us to assign the experimental amide III VCD spectrum of ADP. Moreover, the amide III region of the VCD spectrum was used to determine the relative populations of conformations of ADP in water. Based on the interpretation of the amide III region of VCD spectrum we have shown that dominant conformation of ADP in water is PII which is stabilized by hydrogen bonded water molecules between CO and NH groups on the peptide backbone. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 814–818, 2014.  相似文献   

4.
A G Anderson  J Hermans 《Proteins》1988,3(4):262-265
A direct attack on the protein-folding problem has been initiated with the free energy perturbation methods of molecular dynamics. The complete conformational probability map for the alanine dipeptide is presented. This work uses the SPC model for the explicit hydration of the dipeptide. Free energy differences for the four observed minima (beta, alpha R, alpha L, C7ax) are given, and the free energy barriers between minima are outlined.  相似文献   

5.
Predicting absolute ligand binding free energies to a simple model site   总被引:2,自引:0,他引:2  
A central challenge in structure-based ligand design is the accurate prediction of binding free energies. Here we apply alchemical free energy calculations in explicit solvent to predict ligand binding in a model cavity in T4 lysozyme. Even in this simple site, there are challenges. We made systematic improvements, beginning with single poses from docking, then including multiple poses, additional protein conformational changes, and using an improved charge model. Computed absolute binding free energies had an RMS error of 1.9 kcal/mol relative to previously determined experimental values. In blind prospective tests, the methods correctly discriminated between several true ligands and decoys in a set of putative binders identified by docking. In these prospective tests, the RMS error in predicted binding free energies relative to those subsequently determined experimentally was only 0.6 kcal/mol. X-ray crystal structures of the new ligands bound in the cavity corresponded closely to predictions from the free energy calculations, but sometimes differed from those predicted by docking. Finally, we examined the impact of holding the protein rigid, as in docking, with a view to learning how approximations made in docking affect accuracy and how they may be improved.  相似文献   

6.
R W Dixon  P Kollman 《Proteins》1999,36(4):471-473
We have carried out calculations on the relative free energy of binding of biotin and its S27A and W79A mutants to streptavidin. Consistent with earlier suggestions by Miyamoto and Kollman from free energy component analysis and recent experiments by Stayton and coworkers, the reduction in binding strength by the W79A mutant is significantly larger than that of the S27A mutant. Proteins 1999;36:471-473.  相似文献   

7.
Protein-protein interactions are governed by the change in free energy upon binding, ΔG = ΔH - TΔS. These interactions are often marginally stable, so one must examine the balance between the change in enthalpy, ΔH, and the change in entropy, ΔS, when investigating known complexes, characterizing the effects of mutations, or designing optimized variants. To perform a large-scale study into the contribution of conformational entropy to binding free energy, we developed a technique called GOBLIN (Graphical mOdel for BiomoLecular INteractions) that performs physics-based free energy calculations for protein-protein complexes under both side-chain and backbone flexibility. Goblin uses a probabilistic graphical model that exploits conditional independencies in the Boltzmann distribution and employs variational inference techniques that approximate the free energy of binding in only a few minutes. We examined the role of conformational entropy on a benchmark set of more than 700 mutants in eight large, well-studied complexes. Our findings suggest that conformational entropy is important in protein-protein interactions--the root mean square error (RMSE) between calculated and experimentally measured ΔΔGs decreases by 12% when explicit entropic contributions were incorporated. GOBLIN models all atoms of the protein complex and detects changes to the binding entropy along the interface as well as positions distal to the binding interface. Our results also suggest that a variational approach to entropy calculations may be quantitatively more accurate than the knowledge-based approaches used by the well-known programs FOLDX and Rosetta--GOBLIN's RMSEs are 10 and 36% lower than these programs, respectively.  相似文献   

8.
An adaptive binding mechanism, requiring large conformational rearrangements, occurs commonly with many RNA-protein associations. To explore this process of reorganization, we have investigated the conformational change upon spliceosomal U1A-RNA binding with molecular dynamics (MD) simulations and free energy analyses. We computed the energetic cost of conformational change in U1A-hairpin and U1A-internal loop binding using a hybrid of molecular mechanics and continuum solvent methods. Encouragingly, in all four free energy comparisons (two slightly different proteins, two different RNAs), the free macromolecule was more stable than the bound form by the physically reasonable value of approximately 10 kcal/mol. We calculated the absolute binding free energies for both complexes to be in the same range as that found experimentally.  相似文献   

9.
Abstract

The confinement method is a robust and conceptually simple free energy simulation method that allows the calculation of conformational free energy differences between highly dissimilar states. Application of the method to explicitly solvated systems requires a multi-stage simulation protocol for the calculation of desolvation free energies. Here we show that these desolvation free energies can be readily obtained from an implicit treatment, which is simpler and less costly. The accuracy and robustness of this protocol was shown by the calculation of conformational free energy differences of a series of explicitly solvated test systems. Given the accuracy and ease by which these free energy differences were obtained, the confinement method is promising for the treatment of conformational changes in large and complex systems.  相似文献   

10.
4-Amino-(D3)-trishomocubane-4-carboxylic acid is a constrained alpha-amino acid residue that exhibits promising conformational characteristics, i.e., helical and beta-turns. As part of the development of conformational guidelines for the design of peptides and protein surrogates, the conformational energy calculations on trishomocubane using molecular mechanics and ab initio methods are presented. The C(alpha) carbon of trishomocubane forms part of the cyclic structure, and consequently a peptidic environment was simulated with an acetyl group on its N-terminus and a methylamide group on its C-terminus. Ramachandran maps computed at the molecular mechanics level using the standard AMBER (parm94) force field libraries compared reasonably well with the corresponding maps computed at the Hartree Fock level, using the 6-31G* basis set. Trishomocubane peptide (Ac-Tris-NHMe) is characterized by four low energy conformers corresponding to the C7ax, C7eq, 3(10), and alpha(L) helical structures.  相似文献   

11.
The location of the active site of membrane-bound factor Xa relative to the phospholipid surface was determined both in the presence and absence of factor Va using fluorescence energy transfer. Factor Xa was reacted with 5-(dimethylamino)-1-naphthalenesulfonyl- glutamylglycylarginyl(DEGR) chloromethyl ketone to yield DEGR-Xa, an analogue of factor Xa with a fluorescent dye attached covalently to the active site. When DEGR-Xa was titrated with phosphatidylcholine/phosphatidylserine vesicles containing octadecylrhodamine, fluorescence energy transfer was observed between the donor dyes in the active sites of the membrane-bound enzymes and the acceptor dyes at the outer surface of the phospholipid bilayer. Based on the dependence of the efficiency of singlet-singlet energy transfer upon the acceptor density and assuming kappa 2 = 2/3, the distance of closest approach between the active site probe and the surface of the phospholipid bilayer averaged 61 A in the absence of factor Va and 69 A in the presence of factor Va. These direct measurements show that the active site of factor Xa is located far above the membrane surface. Also, association of factor Xa with factor Va on the membrane surface to form the prothrombinase complex results in a substantial movement of the active site of the enzyme relative to the membrane surface. The 5-(dimethylamino)-1-naphthalenesulfonyl emission in the complete prothrombinase complex was distinct from that in any other combination of components. It therefore appears that the optimum conformation of the prothrombinase active site is achieved only when factor Va, Ca2+, and a membrane surface interact simultaneously with factor Xa. Thus, in addition to its previously demonstrated ability to stimulate factor Xa binding to membranes, factor Va, upon association with factor Xa on a phospholipid surface, allosterically induces a particular active site conformation in factor Xa and also positions the active site at the correct distance above the membrane for prothrombin activation.  相似文献   

12.
Thomas A  Milon A  Brasseur R 《Proteins》2004,56(1):102-109
Using a semiempirical quantum mechanical procedure (FCPAC) we have calculated the partial atomic charges of amino acids from 494 high-resolution protein structures. To analyze the influence of the protein's environment, we considered each residue under two conditions: either as the center of a tripeptide with PDB structure geometry (free) or as the center of 13-16 amino acid clusters extracted from the PDB structure (buried). The partial atomic charges from residues in helices and in sheets were separated. The FCPAC partial atomic charges of the Cbeta and Calpha of most residues correlate with their helix propensity, positively for Cbeta and negatively for Calpha (r2 = 0.76 and 0.6, respectively). The main consequence of burying residues in proteins is the polarization of the backbone C=O bond, which is more pronounced in helices than in sheets. The average shift of the oxygen partial charges that results from burying is -0.120 in helix and -0.084 in sheet with the charge of the proton as unit. Linear correlations are found between the average NMR chemical shifts and the average FCPAC partial charges of Calpha (r2 = 0.8-0.85), N (r3 = 0.67-0.72), and Cbeta (r2 = 0.62) atoms. Correlations for helix and beta-sheet FCPAC partial charges show parallel regressions, suggesting that the charge variations due to burying in proteins differentiate between the dihedral angle effects and the polarization of backbone atoms.  相似文献   

13.
It is well-known that properties of poly(vinyl alcohol) (PVA) in the pure and solution states depend largely on the hydrogen bonding networks formed. In the context of molecular simulation, such networks are handled through the Coulombic interactions. Therefore, a good set of partial atom charges (PACs) for simulations involving PVA is highly desirable. In this work, we calculated the PACs for PVA using a few commonly used population analysis schemes with a hope to identify an accurate set of PACs for PVA monomers. To evaluate the quality of the calculated parameters, we have benchmarked their predictions for free energy of solvation (FES) in selected solvents by molecular dynamics simulations against the ab initio calculated values. Selected solvents were water, ethanol and benzene as they covered a range of size and polarity. Also, PVA with different tacticities were used to capture their effect on the calculated FESs. Based on our results, neither PACs nor FESs are affected by the chain tacticity. While PACs predicted by the Merz-Singh-Kollman scheme were close to original values in the OPLS-AA force field in way that no significant difference in properties of pure PVA was observed, free energy of solvation calculated using such PACs showed greater agreement with ab initio calculated values than those calculated by OPLS-AA (and all other schemes used in this work) in all three solvents considered.  相似文献   

14.
In drug optimization calculations, the molecular mechanics Poisson‐Boltzmann surface area (MM‐PBSA) method can be used to compute free energies of binding of ligands to proteins. The method involves the evaluation of the energy of configurations in an implicit solvent model. One source of errors is the force field used, which can potentially lead to large errors due to the restrictions in accuracy imposed by its empirical nature. To assess the effect of the force field on the calculation of binding energies, in this article we use large‐scale density functional theory (DFT) calculations as an alternative method to evaluate the energies of the configurations in a “QM‐PBSA” approach. Our DFT calculations are performed with a near‐complete basis set and a minimal parameter implicit solvent model, within the self‐consistent calculation, using the ONETEP program on protein–ligand complexes containing more than 2600 atoms. We apply this approach to the T4‐lysozyme double mutant L99A/M102Q protein, which is a well‐studied model of a polar binding site, using a set of eight small aromatic ligands. We observe that there is very good correlation between the MM and QM binding energies in vacuum but less so in the solvent. The relative binding free energies from DFT are more accurate than the ones from the MM calculations, and give markedly better agreement with experiment for six of the eight ligands. Furthermore, in contrast to MM‐PBSA, QM‐PBSA is able to correctly predict a nonbinder. Proteins 2014; 82:3335–3346. © 2014 Wiley Periodicals, Inc.  相似文献   

15.
  1. Download : Download high-res image (147KB)
  2. Download : Download full-size image
  相似文献   

16.
Binding ATP to tryptophanyl-tRNA synthetase (TrpRS) in a catalytically competent configuration for amino acid activation destabilizes the enzyme structure prior to forming the transition state. This conclusion follows from monitoring the titration of TrpRS with ATP by small angle solution X-ray scattering, enzyme activity, and crystal structures. ATP induces a significantly smaller radius of gyration at pH=7 with a transition midpoint at approximately 8mM. A non-reciprocal dependence of Trp and ATP dissociation constants on concentrations of the second substrate show that Trp binding enhances affinity for ATP, while the affinity for Trp falls with the square of the [ATP] over the same concentration range ( approximately 5mM) that induces the more compact conformation. Two distinct TrpRS:ATP structures have been solved, a high-affinity complex grown with 1mM ATP and a low-affinity complex grown at 10mM ATP. The former is isomorphous with unliganded TrpRS and the Trp complex from monoclinic crystals. Reacting groups of the two individually-bound substrates are separated by 6.7A. Although it lacks tryptophan, the low-affinity complex has a closed conformation similar to that observed in the presence of both ATP and Trp analogs such as indolmycin, and resembles a complex previously postulated to form in the closely-related TyrRS upon induced-fit active-site assembly, just prior to catalysis. Titration of TrpRS with ATP therefore successively produces structurally distinct high- and low-affinity ATP-bound states. The higher quality X-ray data for the closed ATP complex (2.2A) provide new structural details likely related to catalysis, including an extension of the KMSKS loop that engages the second lysine and serine residues, K195 and S196, with the alpha and gamma-phosphates; interactions of the K111 side-chain with the gamma-phosphate; and a water molecule bridging the consensus sequence residue T15 to the beta-phosphate. Induced-fit therefore strengthens active-site interactions with ATP, substantially intensifying the interaction of the KMSKS loop with the leaving PP(i) group. Formation of this conformation in the absence of a Trp analog implies that ATP is a key allosteric effector for TrpRS. The paradoxical requirement for high [ATP] implies that Gibbs binding free energy is stored in an unfavorable protein conformation and can then be recovered for useful purposes, including catalysis in the case of TrpRS.  相似文献   

17.
Abstract The surface free energies ( γ b) of a variety of oral streptococci were determined from contact angle measurements on bacterial deposits, using the concept of dispersion and polar components. At least four strains of each species were tested. Strains of Streptococcus mutans, S. sanguis and S. salivarius possessed relatively high surface free energies (103 ± 12 mJ · m−2) and at the species level no significant difference was found. In contrast, the strains of S. mitis had remarkably low surface free energies (45 ± 14 mJ · m−2). S. milleri appeared to be a heterogeneous species, showing surface free energies over a range of 32–119 mJ · m−2. No significant differences were observed between laboratory strains and strains freshly isolated from the oral cavity.  相似文献   

18.
Julien M  Gros P 《Biochemistry》2000,39(15):4559-4568
Limited trypsin digestion was used to monitor nucleotide-induced conformational changes in wild-type P-glycoprotein (Pgp) as well as in nucleotide binding domain (NBD) Pgp mutants. Purified and reconstituted wild-type or mutant mouse Mdr3 Pgps were preincubated with different hydrolyzable or nonhydrolyzable nucleotides, followed by limited proteolytic cleavage at different trypsin:protein ratios. The Pgp tryptic digestion products were separated by SDS-PAGE followed by immunodetection with the mouse monoclonal anti-Pgp antibody C219, which recognizes a conserved epitope (VVQE/AALD) in each half of the protein. Different trypsin digestion patterns were observed for wild-type Pgp incubated with MgCl(2) alone, MgADP, MgAMP.PNP, MgATP, and MgATP + vanadate. A unique trypsin digestion profile suggestive of enhanced resistance to trypsin was observed under conditions of vanadate-induced trapping of nucleotides (MgATP + vanadate). The trypsin sensitivity profiles of Pgp mutants bearing either single or double mutations in Walker A (K429R, K1072R) and Walker B (D551N, D1196N) sequence signatures of NBD1 and NBD2 were analyzed under conditions of vanadate-induced trapping of nucleotides. The proteolytic cleavage pattern observed for the double mutants K429R/K1072R and D551N/D1196N, and for the single mutants K429R, K1072R, and D1196N were similar and clearly distinct from wild-type Pgp under the same conditions. This is consistent with the absence of ATP hydrolysis and of vanadate-induced trapping of 8-azido-ADP previously reported for these mutants [Urbatsch et al. (1998) Biochemistry 37, 4592-4602]. Interestingly, the trypsin digestion profiles observed under vanadate-induced trapping for the D551N and D1196N mutants were quite different, with the D551N mutant showing a profile resembling that seen for wild-type Pgp. The different sensitivity profiles of Pgp mutants bearing mutations at the homologous residue in NBD1 (D551N) and NBD2 (D1196N) suggest possible structural and functional differences between the two sites.  相似文献   

19.
We present an energy function for predicting binding free energies of protein-protein complexes, using the three-dimensional structures of the complex and unbound proteins as input. Our function is a linear combination of nine terms and achieves a correlation coefficient of 0.63 with experimental measurements when tested on a benchmark of 144 complexes using leave-one-out cross validation. Although we systematically tested both atomic and residue-based scoring functions, the selected function is dominated by residue-based terms. Our function is stable for subsets of the benchmark stratified by experimental pH and extent of conformational change upon complex formation, with correlation coefficients ranging from 0.61 to 0.66.  相似文献   

20.
The distance between the phospholipid surface and the active site of membrane-bound meizothrombin, a derivative of prothrombin, was determined directly using fluorescence energy transfer. The active site of prothrombin was exposed after a single cleavage by Echis carinatus protease in the presence of [5-(dimethylamino)-1-naphthalenesulfonyl]glutamylglycylarginyl+ ++ (DEGR) chloromethyl ketone to yield DEGR-meizothrombin and thereby minimize secondary proteolysis. When DEGR-meizothrombin was titrated with 80% phosphatidylcholine, 20% phosphatidylserine vesicles containing octadecylrhodamine, singlet-singlet energy transfer was observed between the donor dyes in the active sites of the membrane-bound proteins and the acceptor dyes at the outer surface of the phospholipid bilayer. This energy transfer required both Ca2+ and phosphatidylserine. Assuming k2 = 2/3, the dependence of the efficiency of energy transfer upon the acceptor density showed that the distance of closest approach between the active site probe and the bilayer surface was 71 +/- 2 A. In the presence of factor Va, the distance was 67 +/- 3 A. These direct measurements show that the active site of meizothrombin is located far above the membrane surface. Also, association of factor Va with meizothrombin on the phospholipid surface appears to cause a slight movement of the meizothrombin protease domain toward the membrane surface. The environment of the dansyl dye covalently attached to the active site of meizothrombin was particularly sensitive to the presence of calcium: addition of Ca2+ ions to metal-free DEGR-meizothrombin reduced the dansyl fluorescence lifetime from 11.7 to 9.0 ns and the dansyl emission intensity by 24%. Hence, the conformation of the active site changed when Ca2+ ions bound to meizothrombin. Since the intensity change was half-maximal at 0.2 mM and was also elicited by the binding of Mg2+ ions, this spectral change correlates with the calcium-dependent conformational change previously observed in fragment 1. We conclude, therefore, that the binding of Ca2+ ions to meizothrombin and, by extension, perhaps to prothrombin, elicits a conformational change that extends beyond the fragment 1 domains into the distant (cf. above) active site or protease domain. The association of factor Va with membrane-bound DEGR-meizothrombin increased both the dansyl emission intensity (by 7%) and polarization. This intensity change and the factor-Va dependent change in energy transfer indicate that the cofactor of the prothrombinase complex functions to modulate the conformation and orientation of both the substrate and the enzyme of the complex.  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