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 46 毫秒
1.
V H Tran  J W Brady 《Biopolymers》1990,29(6-7):977-997
Molecular dynamics simulations have been used to study the motions in vacuum of the disaccharide sucrose. Ensembles of trajectories were calculated for each of the five local minimum energy conformations identified in the adiabatic conformational energy mapping of this molecule. The model sucrose molecules were found to exhibit a variety of motions, although the global minimum energy conformation was found to be dynamically stable, and no transitions away from this structure were observed to occur spontaneously. In all but one of these vacuum trajectories, the intramolecular hydrogen bond between residues was maintained, in accord with recent nmr studies of this molecule in aqueous solution. Considerable flexibility of the furanoid ring was found in the trajectories. No "flips" to the opposite puckering for this ring were found in the simulations starting from the global minimum, although such a transition was observed for a trajectory initiated with one of the higher local minimum energy conformations. Overall, the observed structural fluctuations were consistent with the experimental picture of sucrose as a relatively rigid molecule.  相似文献   

2.
S N Ha  L J Madsen  J W Brady 《Biopolymers》1988,27(12):1927-1952
Constrained conformational energy minimizations have been used to calculate an adiabatic (Φ, ψ) potential energy surface for the disaccharide β-maltose. The inclusion of molecular flexibility in the conformational energy analysis of the disaccharide was found to significantly lower the barriers to conformational transitions, as has been observed previously for other systems. Several low energy wells were identified on the adiabatic surface which differ in energy by small amounts and with low absolute barriers separating them, indicating the possibility of a non-negligible equilibrium population distribution in each well. If such a distribution of conformations existed in the physical system, the conformation observed by NMR NOE measurements would thus be a “virtual” conformation. Molecular dynamics simulations of the motions of this molecule in vacuum were also conducted and indicate that the rate of relaxation of the molecule to the adiabatic surface may be slower than the typical timescale of conformational fluctuations. This effect is apparently due to an unphysical persistence of hydrogen bond patterns in vacuum which does not occur in aqueous solution. Trajectories undergoing transitions between wells were calculated and the effects of such conformational transitions upon the ensemble mean structure, such as might be observed in an NMR experiment, were demonstrated.  相似文献   

3.
K. Ueda  J. W. Brady 《Biopolymers》1996,38(4):461-469
Molecular mechanics calculations have been performed for the disaccharide neocarrabiose, one of the repeat units of β-carrageenan, as a general model for the (1 → 3)-linkage in the carrageenans. An adiabatic conformational energy map for this molecule has been prepared by constrained energy minimization and compared to previously reported relaxed maps. Neither the experimentally determined crystal structure of neocarrabiose nor the fiber diffraction conformation of β-carrageenan is a low energy conformation on the relaxed Ramachandran map. Molecular dynamics simulations in vacuum produced trajectories consistent with this relaxed vacuum surface. However, a simulation with explicitly included solvent water molecules produced a trajectory that remained in the region of the two experimental structures. This dramatic solvation effect is apparently the result of the breaking of an interring hydrogen bond between the O2 hydroxyl groups of neocarrabiose as both groups hydrogen bond to solvent. © 1996 John Wiley & Sons, Inc.  相似文献   

4.
Stortz CA  Cerezo AS 《Biopolymers》2003,70(2):227-239
The adiabatic potential energy surfaces (PES) of six trisaccharides-namely 3,6-An-alpha-D-Galp-(1-->3)-beta-D-Galp-(1-->4)-3,6-An-alpha-D-Galp, beta-D-Galp-(1-->4)-3,6-An-alpha-D-Galp-(1-->3)-beta-D-Galp, and their derivatives sulfated on positions 2 and 4 of the beta-galactose unit-were obtained using the MM3 force field. Each PES was described by a single contour map for which the energy is plotted against the two psi glycosidic angles, given the small variations of the phi glycosidic torsional angle in the low-energy regions of disaccharide maps. In five of the six examples, the surfaces are those expected from the maps of the disaccharidic repeating units of carrageenans, with less important factors altering the additive effect of both linkages. However, when a sulfate group is present on C2 of a beta-galactose reducing end, a new low-energy minimum in a different region is produced, originated in a hydrogen bond between the first and third monosaccharidic moieties of the trisaccharide. The flexibility of the beta-linkages is nearly identical to that in their disaccharide counterparts, while that of the alpha-linkages is slightly reduced, independent of their presence closer or further away from the reducing end. A fair agreement is observed between the x-ray fiber diffraction analysis for a kappa-carrageenan double helix and the surfaces obtained for the trisaccharide analogs of that polymer.  相似文献   

5.
K. Ueda  J. W. Brady 《Biopolymers》1997,41(3):323-330
Molecular mechanics calculations have been performed for the disaccharide carrabiose, one of the repeat units of β-carrageenan, as a general model for the (1→4)-linkage in the carrageenans. An adiabatic conformational energy map for this unsulfated molecule was prepared by constrained energy minimization and compared to a previously reported rigid-residue energy map for the sulfated molecule and to a similar adiabatic map for neocarrabiose, the related (1→3)-linked dimer repeat unit of β-carrageenan. Molecular dynamics simulations of this molecule in vacuo and in an aqueous (TIP3P) solution were calculated, and the observed motions were found to be generally consistent with the vacuum adiabatic energy map. Unlike the case observed in previous simulations of neocarrabiose, little salvation shift in the molecular conformation was observed for carrabiose. From the dynamics, the linkage was observed to be relatively flexible, as has been inferred from experiment on sulfated carrageenan polymers. © 1997 John Wiley & Sons, Inc.  相似文献   

6.
7.
Stortz CA 《Carbohydrate research》2006,341(15):2531-2542
The adiabatic potential energy surfaces (PES) of six trisaccharides, sulfated derivatives of alpha-D-Gal p-(1-->3)-beta-D-Gal p-(1-->4)-alpha-D-Gal p and beta-D-Gal p-(1-->4)-alpha-D-Gal p-(1-->3)-beta-D-Gal p representing models of lambda-, mu-, and nu-carrageenans were obtained using the MM3 force-field at epsilon = 3. Each PES was described by a single contour map for which the energy is plotted against the two psi glycosidic angles, given the small variations of the phi glycosidic torsional angle in the low-energy regions of disaccharide maps. Most surfaces appear as expected from the maps of the disaccharidic repeating units of carrageenans, with less important factors altering the additive effect of both linkages. Only small interactions between the first and third monosaccharidic moieties of the trisaccharides are observed. The flexibility of the alpha-linkages appears nearly identical to that in their disaccharide counterparts, with only one exception, where it appears reduced by the presence of the third monosaccharide. On the other hand, the flexibility of the beta-linkage appears to be equal or sometimes even higher than that observed for the corresponding disaccharide.  相似文献   

8.
The adiabatic potential energy surfaces (PES) of two trisaccharides with 2-linkages (alpha-kojitriose and beta-sophorotriose) were obtained using the MM3 force field, and are represented by a single 3D contour map for which the energy is plotted against the two psi glycosidic angles. In spite of the proximity of the positions where the two monosaccharidic units are linked to the central monosaccharide, an almost independent behavior of both linkages was found for the alpha-linked trisaccharide alpha-kojitriose, i.e., the surfaces are those expected from the maps of the disaccharide containing the same linkage. A slight shift of the position of the global minimum is found to occur, due to a hydrogen bond between the third and first monosaccharide units, which also leads to an increase in flexibility. On the other hand, for the beta-linked trisaccharide beta-sophorotriose, the surface is sharply different from that expected by observation of the disaccharide map. Some of the expected minima cannot appear unless a serious deformation of the phi and/or psi angles is produced. Furthermore, the global minimum corresponds to a combination of different conformations for each of the linkages, whereas another minimum with only slightly higher energy has both glycosidic linkages in a conformation less favored for the disaccharide, though close to that predicted in crystal diffraction studies.  相似文献   

9.
The determination of conformational preferences of oligosaccharides is best approached by describing their preferred conformations on potential energy surfaces as a function of the glycosidic linkage φ, ψ torsional angles. For proper molecular mechanics modelling the flexibility of the rotatable pendant groups must also be considered. The so called adiabatic maps partially mimic the flexibility within the 10 dimensional conformational space of the pendant groups of the given disaccharide. These molecular mechanics maps are considered to be the state-of-the art of the φ, ψ potential energy surface of disaccharides recently calculated. The RAMM (RAndom Molecular Mechanics) method was shown to be able to calculate such profiles automatically. Additionally, based on the continuum solvent approach, RAMM allows the calculation of the effects of solvent on conformational energy profiles. Molecular dynamics simulations are also useful tools to study the influence of solvent on conformational behaviour of oligosaccharides. The capability of the RAMM calculational protocol to locate low-energy conformers on the multidimensional potential energy hypersurfaces of disaccharides is illustrated and compared with molecular dynamics simulations with and without inclusion of the solvent. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

10.
The adiabatic potential energy surfaces (PES) of two trisaccharides (beta-cellotriose and alpha-maltotriose) were obtained using the MM3 force field. Each PES can be described by a single 3D contour map for which the energy is plotted against the two psi glycosidic angles. Given the usually small variations of the phi glycosidic torsional angle in the low-energy regions of disaccharide maps (at least with MM3), it is valid to leave both phi glycosidic angles to relax in the process of building the conformational map of trisaccharides. The surfaces are those expected from the map of disaccharides containing the same linkages and monosaccharide units (i.e., beta-cellobiose and alpha-maltose), with second-order factors altering the 'symmetry' of both linkages. A large low-energy region appears for beta-cellotriose, comprising four minima in close proximity, with barriers between them below 0.6 kcal/mol. On the other hand, for alpha-maltotriose a main global minimum is observed, with several surrounding local minima. The surfaces obtained agree with single-crystal X-ray data on these trisaccharides and derivatives. A reduction of the linkage flexibilities is observed when passing from the disaccharides to the trisaccharides. Furthermore, the linkage closer to the reducing end appears to be less flexible than the linkage closer to the non-reducing end.  相似文献   

11.
Energy surfaces were computed for relative orientations of the relaxed pyranosyl rings of the two anomeric forms of kojibiose, nigerose, and maltose, the (1 → 2)-, (1 → 3)-- and (1 → 4)--linked -glucosyl disaccharides, respectively. Twenty-four combinations of starting conformations of the rotatable side-groups were considered for each disaccharide. Optimized structures were calculated using MM3 on a 20° grid spacing of the torsional angles about the glycosidic bonds. The energy surfaces of the six disaccharides were similar in many respects but differed in detail within the low-energy regions. The maps also illustrate the importance of the exo-anomeric effect and linkage type in determining the conformational flexibility of disaccharides. Torsional conformations of known crystal structures of maltosyl-containing molecules lie in a lower MM3 energy range than previously reported.  相似文献   

12.
Uromodulin is the pregnancy-associated Tamm-Horsfall glycoprotein, with the enhanced ability to inhibit T-cell proliferation. Pregnancy-associated structural changes mainly occur in the O-glycosylation of this glycoprotein. These include up to 12 glycan structures, made up of an unusual core type 2 sequence terminated with one, two, or three sialyl Lewis(x) sequences; this type of O-glycans could serve as E- and P-selectin ligands. The present work focuses on the most complex one; a tetradecamer made up of a type 2 core carrying three sialyl Lewis(x) branches. Five different monosaccharides are assembled by 14 glycosidic linkages. The conformational behavior of the constituting disaccharide segments was evaluated using the flexible residue procedure of the MM3 molecular mechanics procedure. For each disaccharide, the adiabatic energy surface, along with the local energy minima were established. All these results were used for the generation, prior to complete optimization of the tetradecamer. This was followed by a complete exploration of conformational hyperspace throughout the use of the single coordinate method as implemented in the CICADA program. Despite the potential flexibility of the tetradecasaccharide, only four conformational families occur, accounting for more than 95% of the total low energy conformations. For each family, the molecular properties (electrostatic, lipophilicity, and hydrogen potential) were studied. The shape of the tetradecasaccharide is best described as a flat ribbon, flanked by three branches having terminal sialyl residues. Two of the branches interact through nonbonded interactions, bringing further energy stabilization, and limiting the conformational flexibility of the sialyl residues. Only one branch maintains the original conformational features of sialyl Lewis(x). This O-glycan can be seen as a fascinating example of 'dendrimeric' structure, where the spatial arrangement of three S-Le(x) epitopes may favor its complementary 'presentations' for the interactions with E- and P-selectins.  相似文献   

13.
Xia J  Case DA 《Biopolymers》2012,97(5):276-288
Although the crystal structure of the disaccharide sucrose was solved more than 30 years ago, its conformational distribution in aqueous solution is still a matter of debate. We report here a variety of molecular dynamics simulations (mostly of 100 ns) using the GLYCAM06 force field and various water models, paying particular attention to comparisons to NMR measurements of residual dipolar couplings and electron-mediated spin-spin couplings. We focus on the glycosidic linkage conformation, the puckering phase angle of the fructose ring, and intramolecular hydrogen bonds between the two sugars. Our results show that sucrose is indeed a dynamic molecule, but the crystal conformation is qualitatively the dominant one in dilute solution. A second conformational basin, populated in many force fields, is probably overstabilized in the calculations.  相似文献   

14.
The significant variation among solved structures of the λ Cro dimer suggests its flexibility. However, contacts in the crystal lattice could have stabilized a conformation which is unrepresentative of its dominant solution form. Here we report on the conformational space of the Cro dimer in solution using replica exchange molecular dynamics in explicit solvent. The simulated ensemble shows remarkable correlation with available x-ray structures. Network analysis and a free energy surface reveal the predominance of closed and semi-open dimers, with a modest barrier separating these two states. The fully open conformation lies higher in free energy, indicating that it requires stabilization by DNA or crystal contacts. Most NMR models are found to be unstable conformations in solution. Intersubunit salt bridging between Arg4 and Glu53 during simulation stabilizes closed conformations. Because a semi-open state is among the low-energy conformations sampled in simulation, we propose that Cro-DNA binding may not entail a large conformational change relative to the dominant dimer forms in solution.  相似文献   

15.
D R Ripoll  H A Scheraga 《Biopolymers》1990,30(1-2):165-176
The conformational space of the membrane-bound portion of melittin has been searched using the electrostatically driven Monte Carlo (EDMC) method with the ECEPP/2 (empirical conformational energy program for peptides) algorithm. The former methodology assumes that a polypeptide or protein molecule is driven toward the native structure by the combined action of electrostatic interactions and stochastic conformational changes associated with thermal movements. The algorithm produces a Monte Carlo search in the conformational hyperspace of the polypeptide using electrostatic predictions and a random sampling technique, combined with local minimization of the energy function, to locate low-energy conformations. As a result of 8 test calculations on the 20-residue membrane-bound portion of melittin, starting from six arbitrary and two completely random conformations, the method was able to locate a very low-energy region of the potential with a well-defined structure for the backbone. In all of the cases under study, the method found a cluster of similar low-energy conformations that agree well with the structure deduced from x-ray diffraction experiments and with one computed earlier by the build-up procedure.  相似文献   

16.
The structures of five basic pancreatic trypsin inhibitor (BPTI) molecules are compared to establish the extent and nature of the conformational variability resulting from crystal packing effects. BPTI is an ideal system to evaluate such factors because of the availability of high resolution X-ray models of five different BPTI structures, each in a different crystal packing environment. Differences observed among the structures are found to be distributed throughout the molecule, although the regions that display most variability are associated with the loop structures (residues 14-17 and 24-29). The regions of structure that show the largest rms deviations from the mean of the five packing motifs correlate well with the presence of intermolecular contacts in the crystal lattice. For most of the molecules there is also a correspondence between a larger number of intermolecular contacts and systematically higher B-factors, although it is not apparent whether this is induced by the crystal contact or results from the fact that the contacts are made predominantly through surface loops. The conformational differences seen among the X-ray models constitute more than local shifts at the lattice contact surfaces, and in fact involve in some cases the making and breaking of intramolecular H-bonds. The magnitudes of the differences among packing models are significantly larger than those usually associated with changes induced by mutagenesis; for instance; the structural differences at the site of mutation observed on removing an internal disulfide from the molecule are significantly less than those associated with lattice contact effects. The crystal packing conformations are compared to representative structures of BPTI generated during a 96-psec molecular dynamics (MD) simulation. This comparison shows a high level of correspondence between the protein flexibility indicated by the X-ray and MD analyses, and specifically between those regions that are most variable. This suggests that the regions that show most variability among the crystal packing models are not artifacts of crystallization, but rather represent true low-energy conformers that have been preferentially selected by crystallization factors.  相似文献   

17.
Vacuum UV CD spectra of methyl 3-O-(α-D -mannopyranosyl)-α-D -mannopyranoside in D2O and as a cast film were obtained in the 145–200 nM region. The disaccharide solution CD per residue is nearly identical to that of the monosaccharide solution CD, and to the monosaccharide film CD. Conversely, the disaccharide film spectrum exhibits a strong positive CD linkage contribution in the 160–170 nm range, which is consistent with the known crystal conformation under the aegis of previously determined sector rules. The close similarity between the monosaccharide and disaccharide solution spectra, therefore, reflects conformational averaging in which the net linkage contribution is approximately zero. The present observation of significant solution linkage flexibility confirms previous conclusions based on optical rotation, as well as conclusions of others based on nmr data. Moreover, when combined with those earlier results, the present work demonstrates the population of at least three distinct potential energy wells on the disaccharide ϕ, ψ potential energy surface. © 1996 John Wiley & Sons, Inc.  相似文献   

18.
Ab initio conformational maps for beta-lactose in both the gas phase and in aqueous solution have been constructed at the HF/6-31G(d,p) level of calculation. The results of the gas-phase ab initio calculations allow us to conclude that a rigid conformational map is able to predict the regions of the minima in the potential energy surface of beta-lactose, in full agreement with those found in the relaxed conformational map. The solvation effects do not give rise to any new local minimum in the potential energy surface of beta-lactose, but just change the relative Boltzmann populations of the conformers found in the gas-phase calculations. The values obtained for heteronuclear spin coupling constant (3J(H,C)), using the seven most stable conformers in solution are in good agreement with the available experimental values. This is a good indication that ab initio rigid conformational maps can be reliably used to sort the most stable conformers of beta-lactose.  相似文献   

19.
The crystal structure of the ternary complex of NADPH, the potent antifolate [2, 4-diamino-5-?3-[3-(2-acetyloxyethyl)-3-benzyltriazen-1-yl]-4 -chloroph enyl?-6-ethylpyrimidine] (TAB, 1) and Pneumocystis carinii dihydrofolate reductase (pcDHFR), refined to 2.1 A resolution, reveals that TAB binds similar to the antifolates trimethoprim and methotrexate. These data also reveal multiple conformations for the binding geometry of TAB with two preferred orientations of the acetyloxy and benzyl groups that results from a 180 degrees rotation about the N2-N3 triazenyl bond. The methyl of the acetyloxy and benzyl ring of TAB probes large hydrophobic regions of the p-aminobenzoyl folate binding pocket of the active site, in particular the region near Phe69, which is unique to the pcDHFR sequence. These results confirm prior molecular modeling investigations of the binding of TAB to pcDHFR that identified four low-energy binding geometries, two involving rotations about the terminal N(2)-N(3) triazenyl linkage and two involving atropisomerism about the pivotal pyrimethamine-phenyl bond. The primary differences in the molecular dynamics (MD) models and those observed in this crystal complex result from small conformational changes in active-site residues on energy minimization. However, two MD models place the acetyloxy and benzyl ring groups in a region of the active site between the cofactor-binding region and the p-aminobenzoyl folate pocket; an orientation never observed in any DHFR crystal structure to date. These conformers interact with solvent near the enzyme surface and are probably not observed due to the loss of specific hydrogen bonds with the enzyme. The high species pcDHFR selectivity of TAB could be the result of ligand flexibility that enables multiple binding orientations at the enzyme active site. Further modification of the acetyloxy region of TAB could increase its potency and selectivity for pcDHFR.  相似文献   

20.
The conformational change taking place in myoglobin concomitantly with the observed geometrical change at the heme-His(F8) linkage upon oxygenation is studied by normal mode analysis, which is based on the quadratic approximation of the conformational energy function. The heme-globin interaction energy increases for this change by 8.114 kcal/mol when both the heme group and the globin molecule are held rigid. When they are permitted flexibility, the interaction energy relaxes by 7.038 kcal/mol, and the difference (1.076 kcal/mol) is distributed as strain energy within the molecule. This increase is the work necessary for the heme group to move against the force exerted by the globin. If the force is assumed to be invariable during this move, the work is small, 0.276 kcal/mol, meaning that the force is strongly variable. Furthermore, this means that the heme group is located near the equilibrium point of the potential energy of the heme-globin interaction. The change in the localized strain energy stored in the force field at the linkage between the heme and the imidazole of HisF8 is estimated to be of the same order of magnitude as the distributed energy. The largest atomic displacements are observed at the region from the F helix to the beginning of the G helix, and secondary large displacements occur at several regions, i.e, the A helix, from the C helix to the CD corner, the E helix, and the C-terminal side of the H helix. All of these regions have strong dynamic interactions with the heme group, either directly or indirectly. Their secondary structures show complex deformations. In other parts, relatively rigid segments undergo rotational and/or bending changes in a way consistent with the large changes described above and close atomic packing within the molecule. The calculated conformational change is decomposed to vibrational normal modes of deoxymyoglobin. The magnitude of the conformational change, measured by the mass-weighted mean-square atomic displacement, is accounted for up to 92.0% by the 151 normal modes with frequencies lower than 40 cm-1. In descending order of contribution, the first six modes, each of which has a frequency lower than 12 cm-1, account for up to 57.4%. This means that the functionally important conformational change can well be expressed in terms of a relatively small number of collective low frequency normal modes.  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