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.
Explicit-solvent molecular dynamics (MD) simulations of the 11 glucose-based disaccharides in water at 300 K and 1 bar are reported. The simulations were carried out with the GROMOS 45A4 force-field and the sampling along the glycosidic dihedral angles ? and ψ was artificially enhanced using the local elevation umbrella sampling (LEUS) method. The trajectories are analyzed in terms of free-energy maps, stable and metastable conformational states (relative free energies and estimated transition timescales), intramolecular H-bonds, single molecule configurational entropies, and agreement with experimental data. All disaccharides considered are found to be characterized either by a single stable (overwhelmingly populated) state ((1→n)-linked disaccharides with n = 1, 2, 3, or 4) or by two stable (comparably populated and differing in the third glycosidic dihedral angle ; gg or gt) states with a low interconversion barrier ((1→6)-linked disaccharides). Metastable (anti-? or anti-ψ) states are also identified with relative free energies in the range of 8-22 kJ mol−1. The 11 compounds can be classified into four families: (i) the α(1→1)α-linked disaccharide trehalose (axial-axial linkage) presents no metastable state, the lowest configurational entropy, and no intramolecular H-bonds; (ii) the four α(1→n)-linked disaccharides (n = 1, 2, 3, or 4; axial-equatorial linkage) present one metastable (anti-ψ) state, an intermediate configurational entropy, and two alternative intramolecular H-bonds; (iii) the four β(1→n)-linked disaccharides (n = 1, 2, 3, or 4; equatorial-equatorial linkage) present two metastable (anti-? and anti-ψ) states, an intermediate configurational entropy, and one intramolecular H-bond; (iv) the two (1→6)-linked disaccharides (additional glycosidic dihedral angle) present no (isomaltose) or a pair of (gentiobiose) metastable (anti-?) states, the highest configurational entropy, and no intramolecular H-bonds. The observed conformational preferences appear to be dictated by four main driving forces (ring conformational preferences, exo-anomeric effect, steric constraints, and possible presence of a third glycosidic dihedral angle), leaving a secondary role to intramolecular H-bonding and specific solvation effects. In spite of the weak conformational driving force attributed to solvent-exposed H-bonds in water (highly polar protic solvent), intramolecular H-bonds may still have a significant influence on the physico-chemical properties of the disaccharide by decreasing its hydrophilicity. Along with previous work, the results also complete the suggestion of a spectrum of approximate transition timescales for carbohydrates up to the disaccharide level, namely: ∼30 ps (hydroxyl groups), ∼1 ns (free lactol group, free hydroxymethyl groups, glycosidic dihedral angle in (1→6)-linked disaccharides), ∼10 ns to 2 μs (ring conformation, glycosidic dihedral angles ? and ψ). The calculated average values of the glycosidic torsional angles agree well with the available experimental data, providing validation for the force-field and simulation methodology employed.  相似文献   

2.
We studied the temperature dependent vibrational modes of the glycosidic bond in trehalose, sucrose, and maltose at wavenumbers ranging from 1000 to 1200 cm(-1). We found that the slope of temperature dependent Raman shifts of the glycosidic bond in trehalose and sucrose changed at temperatures around 120 degrees C, indicating a bond length or a bond angle (dihedral and torsional angles) change. However, we did not observe any slope change in maltose because the melting temperature of maltose is very close to 120 degrees C. We also found, at temperatures below 120 degrees C, that Raman shifts of the vibrational modes of the glycosidic bond in trehalose showed the strongest temperature dependence among the three disaccharides.  相似文献   

3.
The potential energy surfaces of several alpha-(1-->3)- and beta-(1-->4)-linked disaccharides were obtained and plotted in terms of energy versus psi glycosidic angle. These plots were compared to those obtained previously in the way of the usual 3D contour maps, which relate the energy with the two glycosidic angles (phi and psi). Given the usually small variations of the phi angle in the low-energy regions (at least using MM3), both kinds of graphs lead to similar conclusions concerning flexibility measurements by two different methods and assessment of the effects of sulfation and/or hydroxyl group orientation. Only second-order effects were found with some sulfated disaccharides, not changing the general conclusions. The computational efforts required to produce those plots are smaller, and the plots are easier to interpret. Besides, the conversion of a 3D map into a 2D plot leaves the possibility of constructing 3D maps of carbohydrates including a second variable different to phi, e.g., the second psi angle of a trisaccharide or the omega angle of a 6-linked disaccharide.  相似文献   

4.
Three new hydroxymethyl-linked non-natural disaccharide analogues, containing an additional methylene group in between the glycosidic linkage, were synthesized by utilizing 4-C-hydroxymethyl-α-d-glucopyranoside as the glycosyl donor. A kinetic study was undertaken to assess the hydrolytic stabilities of these new disaccharide analogues toward acid-catalyzed hydrolysis, at 60 °C and 70 °C. The studies showed that the disaccharide analogues were stable, by an order of magnitude, than naturally-occurring disaccharides, such as, cellobiose, lactose, and maltose. The first order rate constants were lower than that of methyl glycosides and the trend of hydrolysis rate constants followed that of naturally-occurring disaccharides. α-Anomer showed faster hydrolysis than the β-anomer and the presence of axial hydroxyl group also led to faster hydrolysis among the disaccharide analogues. Energy minimized structures, derived through molecular modeling, showed that dihedral angles around the glycosidic bond in disaccharide analogues were nearly similar to that of naturally-occurring disaccharides.  相似文献   

5.
《Carbohydrate research》1988,173(1):89-99
Reactions of (1→4)- and (1→6)-linked disaccharides, mainly of maltose and isomaltose, with the Fenton reagent under physiological conditions were studied. Chemical characterization of oxidation products was conducted by g.l.c. and g.l.c.-m.s. of their trimethylsilyl derivatives, and the results demonstrated that (1→6)-linked disaccharides are more reactive with the hydroxyl radical (·OH) generated by the Fenton reagent than (1→4)-linked disaccharides. About 35–40% of (1→6)-and 15–20% of (1→4)-linked disaccharides were oxidatively degraded to smaller molecules after incubation for 24 h. Of the four disaccharides examined, namely, maltose, isomaltose, cellobiose, and gentiobiose, the α-(1→6)-linked disaccharide isomaltose exhibited the highest reactivity, whereas the β-(1→4)-linked disaccharide cellobiose showed the lowest. These results suggest the existence of a relationship between the configuration of the glycosidic linkage and the reactivity with ·OH in aqueous solution.  相似文献   

6.
Molecular dynamics (MD) simulations were used to study the stability and solvation of amylose and cellulose fragments. The recently developed gromos carbohydrate force field was further tested by simulating maltose, cellobiose, and maltoheptaose. The MD simulations reproduced fairly well the favorable conformations of disaccharides defined by the torsional angles related with the glycosidic bond and the radius gyration of maltoheptaose. The effects of methylation at different hydroxyl groups on the stability of amylose and cellulose fragments were investigated. The methylations of O-2 and O-3 reduce the stability of a single helix more than methylation at O-6, while the latter reduces the stability of a double helix more. Solvation free-energy differences between the unsubstituted amylose and cellulose fragments and the methylated species were studied using the single-step perturbation method. It was found that methylation at O-2 has the biggest effect, in agreement with experiment.  相似文献   

7.
Zeng Y  Ning J  Kong F 《Carbohydrate research》2003,338(4):307-311
In (1-->3)-glucosylation the glycosyl bond originally present in either donor or acceptor is shown to control the stereoselectivity of the forthcoming bond, i.e., the newly formed glycosidic linkage has the opposite anomeric configuration of that of either the donor or acceptor. Therefore, with alpha-(1-->3)-linked disaccharides with nonreducing ends that have the 3-OH free as the acceptor and an acetylated glucosyl trichloroacetimidate as the donor, or with an alpha-(1-->3)-linked acetylated disaccharide trichloroacetimidate as the donor and a glucoside with 3-OH free as the acceptor, beta-linked trisaccharides were obtained. Meanwhile, with beta-(1-->3)-linked disaccharides that have nonreducing ends with the 3-OH free as the acceptor and an acetylated glucosyl trichloroacetimidate as the donor, or with a beta-(1-->3)-linked acetylated disaccharide trichloroacetimidate as the donor and a glucoside with the 3-OH free as the acceptor, alpha-linked trisaccharides were obtained in spite of the C-2 neighboring group participation.  相似文献   

8.
A novel method for the determination of the three-dimensional (3D) structure of oligosaccharides in the solid state using experimental 13C NMR data is presented. The approach employs this information, combined with 13C chemical shift surfaces (CSSs) for the glycosidic bond carbons in the generation of NMR pseudopotential energy functions suitable for use as constraints in molecular modeling simulations. Application of the method to trehalose, cellobiose, and cellotetraose produces 3D models that agree remarkably well with the reported X-ray structures, with phi and psi dihedral angles that are within 10 degrees from the ones observed in the crystals. The usefulness of the approach is further demonstrated in the determination of the 3D structure of the cellohexaose, an hexasaccharide for which no X-ray data has been reported, as well as in the generation of accurate structural models for cellulose II and amylose V6.  相似文献   

9.
The structure of the disaccharide cellulose subunit cellobiose (4-O-β-D-glucopyranosyl-D-glucose) in solution has been determined via neutron diffraction with isotopic substitution (NDIS), computer modeling and nuclear magnetic resonance (NMR) spectroscopic studies. This study shows direct evidence for an intramolecular hydrogen bond between the reducing ring HO3 hydroxyl group and the non-reducing ring oxygen (O5′) that has been previously predicted by computation and NMR analysis. Moreover, this work shows that hydrogen bonding to the non-reducing ring O5′ oxygen is shared between water and the HO3 hydroxyl group with an average of 50% occupancy by each hydrogen-bond donor. The glycosidic torsion angles φH and ψH from the neutron diffraction-based model show a fairly tight distribution of angles around approximately 22° and −40°, respectively, in solution, consistent with the NMR measurements. Similarly, the hydroxymethyl torsional angles for both reducing and non-reducing rings are broadly consistent with the NMR measurements in this study, as well as with those from previous measurements for cellobiose in solution.  相似文献   

10.
The spatial structure of the methylamide of N-acetyl-L-lysine has been analysed taking into account non-bonded and electrostatic interactions, torsional energy, bond angles distortion and hydrogen bonding. Conformational capacities of the backbone and mutual dependence of spatial structures of the backbone and the side chain was described by conformational maps obtained by energy minimisation, the dihedral angles and the bond angles of the side chain being varied for every phi, psi point. Every possible combination for phi, psi, x1-x5-angles was used corresponding to the stable form of the backbone and to torsion potential minima of the initial approximations in the calculation of preferred conformations of the molecule. Comparisons are made between stable forms of the methylamide of N-acetyl-L-lysine and Lys residues in proteins with known structure.  相似文献   

11.
The disaccharide, alpha-maltose, forms the molecular basis for the analysis of the structure of starch, and determining the conformational energy landscape as the molecule oscillates around the glycosidic bonds is of importance. Thus, it is of interest to determine, using density functionals and a medium size basis set, a relaxed isopotential contour map plotted as a function of the phi(H) and psi(H) dihedral angles. The technical aspects include the method of choosing the starting conformations, the choice of scanning step size, the method of constraining the specific dihedral angles, and the fitting of data to obtain well defined contour maps. Maps were calculated at the B3LYP/6-31+G( *) level of theory in 5 degrees intervals around the (phi(H),psi(H))=(0 degrees ,0 degrees ) position, out to approximately +/-30 degrees or greater, for gg-gg'-c, gg-gg'-r, gt-gt'-c, gt-gt'-r, tg-tg'-c, and tg-tg'-r conformers, as well as one-split gg(c)-gg'(r) conformer. The results show that the preferred conformation of alpha-maltose in vacuo depends strongly upon the hydroxyl group orientations ('c'/'r'), but the energy landscape moving away from the minimum-energy position is generally shallow and transitions between conformational positions can occur without the addition of significant energy. Mapped deviations of selected parameters such as the dipole moment; the C1-O1-C4', H1-C1-O1, and H4'-C4'-O1 bond angles; and deviations in hydroxymethyl rotamers, O5-C5-C6-O6, O5'-C5'-C6'-O6', C5-C6-O6-H, and C5'-C6'-O6'-H', are presented. These allow visualization of the structural and energetic changes that occur upon rotation about the glycosidic bonds. Interactions across the bridge are visualized by deviations in H(O2)...O3', H(O3')...O2, and H1...H4' distances and the H(O2)-O2-C2-C1 and H'(O3')-O3'-C3'-C4' hydroxyl dihedral angles.  相似文献   

12.
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   

13.
The time-averaged solution conformation of a unique bis-sulfated glycolipid (HSO3)2-2,6Manalpha-2Glcalpha-1-sn-2,3-O-alkylglycerol , was studied in terms of the torsional angles of two glycosidic linkages, phi (H1-C1-O-Cx) and psi (C1-O-Cx-Hx), derived from heteronuclear three-bond coupling constants (3JC,H), and inter-residual proton-proton distances from J-HMBC 2D and ROESY experiments, respectively. The dihedral angles of Glcalpha1Gro in glycolipids were determined for the first time. The C1-C4 diagonal line of the alpha-glucose ring makes an angle of approximately 120 degrees with the glycerol backbone, suggesting that the alpha-glucose ring is almost parallel to the membrane surface in contrast with the perpendicular orientation of the beta-isomer. Furthermore, minimum-energy states around the conformation were estimated by Monte Carlo/stochastic dynamics (MCSD) mixed-mode simulations and the energy minimization with assisted model building and energy refinement (AMBER) force field. The Glcalpha1Gro linkage has a single minimum-energy structure. On the other hand, three conformers were observed for the Manalpha2Glc linkage. The flexibility of Manalpha2Glc was further confirmed by the absence of inter-residual hydrogen bonds which were judged from the temperature coefficients of the chemical shifts, ddelta/dT (-10-3 p.p.m. degrees C-1), of hydroxy protons. The conformational flexibility may facilitate interaction of extracellular substances with both sulfate groups.  相似文献   

14.
The crystal structure of beta-D-glucopyranosyl-(1-->4)-alpha-D-glucopyranose (alpha-cellobiose) in a complex with water and NaI was determined with Mo K(alpha) radiation at 150 K to R=0.027. The space group is P2(1) and unit cell dimensions are a=9.0188, b=12.2536, c=10.9016 A, beta=97.162 degrees. There are no direct hydrogen bonds among cellobiose molecules, and the usual intramolecular hydrogen bond between O-3 and O-5' is replaced by a bridge involving Na+, O-3, O-5', and O-6'. Both Na+ have sixfold coordination. One I(-) accepts six donor hydroxyl groups and three C-H***I(-) hydrogen bonds. The other accepts three hydroxyls, one Na+, and five C-H***I(-) hydrogen bonds. Linkage torsion angles phi(O-5) and psi(C-5) are -73.6 and -105.3 degrees, respectively (phi(H)=47.1 degrees and psi(H)=14.6 degrees ), probably induced by the Na+ bridge. This conformation is in a separate cluster in phi,psi space from most similar linkages. Both C-6-O-H and C-6'-O-H are gg, while the C-6'-O-H groups from molecules not in the cluster have gt conformations. Hybrid molecular mechanics/quantum mechanics calculations show <1.2 kcal/mol strain for any of the small-molecule structures. Extrapolation of the NaI cellobiose geometry to a cellulose molecule gives a left-handed helix with 2.9 residues per turn. The energy map and small-molecule crystal structures imply that cellulose helices having 2.5 and 3.0 residues per turn are left-handed.  相似文献   

15.
The nonbonded interaction energy of disaccharides, mannobiose and galactobiose and polysaccharides mannan and galactan have been computed as a function of dihedral angles (?,ψ). The conformation (40°, ?20°) has been preferred for the mannan chain from nonbonded interaction energy considerations. The O5…O3′ type of intramolecular hydrogen bond has been found to be possible in the above conformation. Comparison of the allowed region of mannan with those of cellulose and xylan indicates that the monomer unit, in mannan chain has slightly higher freedom of rotation than that of cellulose and less than that of xylan. As in cellulose and mannan, the freedom of rotation of the monomer units in β-1,4′ galactan is highly restricted. Unlike mannan (which prefers an extended conformation) the β-1,4′ galactan prefers a helical conformation similar to amylose. Just as in amylose the O2…O3′ type hydrogen bond between contiguous residues is also possible in β-1,4′ galactan.  相似文献   

16.
The solution conformation of model compounds for the tri'-antennary and tetraantennary (six-arm) branch point of N-linked glycans has been determined through the use of chemical shift, relaxation, and nuclear Overhauser enhancement data. The object was to establish the conformation about the glycosidic linkages in the N-linked substructure GlcNAc(beta 1,6) [GlcNAc(beta 1,2)] Man(alpha)- by estimation of values for the appropriate glycosidic torsional angles. The GlcNAc(beta 1,6) linkage in a trisaccharide model compound was found to be constrained to a narrow rotameric subpopulation about the substituted Man C5-C6 bond (omega = -60 degrees) and a narrow range of possible phi - psi values. Free rotation about the Man C5-C6 bond was obstructed by unfavorable steric interactions between the GlcNAc(beta 1,6) and GlcNAc(beta 1,2) residues. A phi, psi value of 55 degrees, 190 degrees was found to be consistent with the NMR data for the GlcNAc(beta 1,6) linkage. However, the value of psi appears to be "virtual" in that the molecule is in equilibrium between two different values (90 degrees and 252 degrees). For the GlcNAc(beta 1,2) linkage, complete agreement between all the observed NMR parameters and all the calculated ensemble average values could only be obtained with a set of potential energy functions which included hydrogen bonding. Other choices of potentials yielded calculated values that disagreed with at least two of the observed quantities. As a result, we infer that an interresidue hydrogen bond is formed, and we find it to be between the GlcNAc(beta 1,2) ring oxygen and the Man C3 hydroxyl.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The normal modes of the gramicidin-A dimer channel.   总被引:1,自引:4,他引:1       下载免费PDF全文
The dynamics of the gramicidin-A dimer channel is studied in the harmonic approximation by a vibrational analysis of the atomic motions relative to their equilibrium positions. The system is represented by an empirical potential energy function, and all degrees of freedom (bonds lengths, bond angles, and torsional angles) are allowed to vary. The thermal fluctuations in the backbone dihedral angles phi and psi, atomic root mean square displacements, and the correlations between the different amide planes are computed. It is found that only adjacent dihedral psi i and phi i+1 are strongly correlated, while different hydrogen-bonded amide planes are only weakly correlated. Modes with relatively low vibrational frequencies (75-175 cm-1) make the dominant contributions to the carbonyl librations. The general flexibility of the structure and the role of carbonyl librations in the ion transport mechanism are discussed.  相似文献   

18.
The conformational preference of alpha-D-Manp-(1-->2)-alpha-D-Manp-(1-->O)-L-Ser has been investigated by one-dimensional (1)H,(1)H T-ROESY experiments and molecular-dynamics simulations with CHARMM22 type of force fields and water as explicit solvent. Proton-proton distances were obtained from the simulations and subsequently experimentally determined distances could be derived. Measurements were performed on the title compound as well as on selectively deuterium-substituted analogues synthesized as part of this study to alleviate possible NMR spectroscopic difficulties. A very good agreement was present between the separate NMR experiments. In the subsequent analysis a key nuclear Overhauser effect between the anomeric protons in the two sugar residues was used to assess the conformational dynamics revealed by the molecular simulations. The combined results support a model in which two states are significantly populated as a result of flexibility around the bond defined by the glycosidic torsion angle psi.  相似文献   

19.
Molecular dynamics simulations have been used to calculate the potentials of mean force for separating short cellooligomers in aqueous solution as a means of estimating the contributions of hydrophobic stacking and hydrogen bonding to the insolubility of crystalline cellulose. A series of four potential of mean force (pmf) calculations for glucose, cellobiose, cellotriose, and cellotetraose in aqueous solution were performed for situations in which the molecules were initially placed with their hydrophobic faces stacked against one another, and another for the cases where the molecules were initially placed adjacent to one another in a co-planar, hydrogen-bonded arrangement, as they would be in cellulose Iβ. From these calculations, it was found that hydrophobic association does indeed favor a crystal-like structure over solution, as might be expected. Somewhat more surprisingly, hydrogen bonding also favored the crystal packing, possibly in part because of the high entropic cost for hydrating glucose hydroxyl groups, which significantly restricts the configurational freedom of the hydrogen-bonded waters. The crystal was also favored by the observation that there was no increase in chain configurational entropy upon dissolution, because the free chain adopts only one conformation, as previously observed, but against intuitive expectations, apparently due to the persistence of the intramolecular O3-O5 hydrogen bond.  相似文献   

20.
The cyclic tetrasaccharide cyclo-[-->6)-alpha-D-Glcp-(1-->3)-alpha-D-Glcp-(1-->6)-alpha-D-Glcp-(1-->3)-alpha-D-Glcp-(1-->] is the major compound obtained by the action of endo-alternases on the alternan polysaccharide. Crystals of this cyclo-tetra-glucose belong to the orthorhombic space group P2(1)2(1)2(1) with a = 7.620(5), b = 12.450(5) and c = 34.800(5) A. The asymmetric unit contains one tetrasaccharide together with five water molecules. The tetrasaccharide adopts a plate-like overall shape with a very shallow depression on one side. The shape is not fully symmetrical and this is clearly apparent on comparing the (phi, psi) torsion angles of the two alpha-(1-->6) linkages. There is almost 10 degrees differences in phi and more than 20 degrees differences in psi. The hydrogen bond network is asymmetric, with a single intramolecular hydrogen bond: O-2 of glucose ring 1 being the donor to O-2 of glucose ring 3. These two hydroxyl groups are located below the ring and their orientation, dictated by this hydrogen bond, makes the floor of the plate. Among the five water molecules, one located above the center of the plate occupies perfectly the shallow depression in the plate shape formed by the tetrasaccharide. Molecular dynamics simulation of the tetrasaccharide in explicit water allows rationalization of the discrepancies observed between the X-ray structures and data obtained previously by NMR.  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