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.
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   

2.
A beta-(1-->6)-branched beta-(1-->3)-glucohexaose, present in many biologically active polysaccharides from traditionally herbal medicines such as Ganoderma lucidum, Schizophyllum commune and Lentinus edodes, was synthesized as its lauryl glycoside 32, and its analogues 18, 20 and 33 containing an alpha-(1-->3) linked bond were synthesized. It is interesting to find that coupling of a 3,6-branched acylated trisaccharide trichloroacetimidate donor 9 with 3,6-branched acceptors 13 and 16 with 3'-OH gave the alpha-(1--> 3)-linked hexasaccharides 17 and 19, respectively, in spite of the presence of C-2 ester capable of neighboring group participation. However, coupling of 9 with 4-methoxyphenyl 4,6-O-benzylidene-beta-D-glucopyranoside (27) selectively gave beta-(1-->3)-linked tetrasaccharide 28. Simple chemical transformation of the tetrasaccharide 28 gave acylated tetrasaccharide trichloroacetimidate 29. Coupling of 29 with lauryl (1-->6)-linked disaccharide 26 with 3-OH gave beta-(1-->3)-linked hexasaccharide 30 as the major product. Bioassay showed that in combination with the chemotherapeutic agent cyclophospamide (CPA), the hexaose 18 at a dose of 0.5-1mg/kg substantially increased the inhibition of S(180) for CPA, but decreased the toxicity caused by CPA. Some of these oligosaccharides also inhibited U(14) noumenal tumor in mice effectively.  相似文献   

3.
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   

4.
Stortz CA 《Carbohydrate research》2004,339(14):2381-2390
The adiabatic potential energy surfaces (PES) of alpha-L-Fuc-(1-->3)-alpha-L-Fuc and their counterparts disulfated at 2,2' and 4,4', and tetrasulfated at 2,2',4,4', which are representative of fucoidan structures, were obtained using the mm3 force field, and plotted as contour maps and as 2D graphs representing the energy versus the psi angle. The surfaces of the corresponding trisaccharides were also obtained and represented by a single 3D contour map for which the energy is plotted against the two psi glycosidic angles. For the nonsulfated disaccharide, similar populations of two minima occur. A substantial sulfate effect is observed. Whereas sulfation on both of the 2-positions shift the global minimum to positive psiH angles, sulfation on both of the 4-positions deepen the well at negative psiH values. A similar effect occurred in their galactose counterparts. Sulfation on the 2- and 4-positions carry the additive effect of both groups. The same trend was observed for both linkages present in the trisaccharides, with minor differences. For instance, the 4,4',4" trisulfated compound exhibits a trend by which the glycosidic linkage closer to the nonreducing end appears to be highly flexible, with similar energies in both conformers. Raising the dielectric constant on nonsulfated oligosaccharides was found to give a better agreement with experimental determinations.  相似文献   

5.
The 'adiabatic' (phi, psi) potential-energy surface of the disaccharide alpha-D-galactopyranosyl-(1-->3)-beta-D-galactopyranose was obtained by several established methods, using the MM3 molecular mechanics force field. The constrained minimizations throughout the whole grid were carried out using sharply different dielectric constants. The attainment of the 'true' adiabatic map is very difficult due to the 'multiple minimum problem', originating in the large number of exocyclic pendant groups present in a disaccharide. However, these results suggest that at low dielectric constants, the usual approach starting with conformers carrying cooperative hydrogen bonds results in a good approximation to the true adiabatic map, while at high dielectric constants this approach fails due to the damping of electrostatic and hydrogen-bonding interactions.  相似文献   

6.
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   

7.
phi-psi maps of N-acetyl alanine N'-methyl amide have been computed using the CHARMM potential, the all-atom AMBER potential, and the ECEPP/2 potential, before and after adiabatic relaxation. Maps using the CHARMM and AMBER potentials were determined with values of 1.0 and 4.0 for the dielectric constant epsilon, and with a distance dependent dielectric constant. Adiabatic relaxation was carried out using flexible geometry for the CHARMM and AMBER potentials, and using rigid geometry for the AMBER and ECEPP potentials. In all cases, the lowest energy was found in the C7eq region (phi approximately -70 degrees, psi approximately 70 degrees). The maps with CHARMM and AMBER with epsilon = 4.0 and with ECEPP, without adiabatic relaxation, were broadly similar but differed in the relative energies allotted to high-energy regions of the map. After adiabatic relaxation with rigid geometry, the map with ECEPP, and the map with AMBER using a distance-dependent dielectric constant, agreed fairly well apart from differences in the relative energies of the alpha R, alpha L, and C7ax regions. After adiabatic relaxation with flexible geometry, the maps with CHARMM and AMBER became very similar; the lowest energies were observed in the C7eq region, the C5 region (phi approximately -150 degrees, psi approximately 150 degrees) and the C7ax region (phi approximately 70 degrees, psi approximately -70 degrees). Breakdown of the energies, after adiabatic relaxation, into electrostatic, nonbonded, and geometric (including torsional) contributions, showed that (1) with fixed geometry, the nonbonded and torsional contribution to the ECEPP and AMBER potentials were very similar, but the electrostatic contributions were markedly different; (2) with flexible geometry, the nonbonded contribution to the CHARMM and AMBER potentials did not vary greatly over the whole map. The phi-psi maps were subjected to three simple comparisons with experiment. (1) The maps were used to predict the characteristic ratio for poly-L-alanine, and the results were compared with experimental findings (D.A. Brant and P.J. Flory, J. Amer. Chem. Soc. 87, 2788-2791, 1965). The agreement with experiment was acceptable for ECEPP, and for CHARMM after adiabatic relaxation, marginal for AMBER after adiabatic relaxation, and unsatisfactory for CHARMM or AMBER without adiabatic relaxation. (2) Deviations of bond angles from their equilibrium values, in energy-minimized conformations, were compared with values deduced from crystals of terminally-blocked amino acids. With both the CHARMM and AMBER potentials using flexible geometry, one or more excessive deviations was observed in the C7ax local minimum.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
Chen L  Kong F 《Carbohydrate research》2002,337(15):1373-1380
A practical synthesis of beta-D-GlcA-(1-->3)-beta-D-Gal-(1-->3)-beta-D-Gal-(1-->4)-beta-D-Xyl-(1-->OMe) was achieved by coupling of methyl 2,3,4-tri-O-acetyl-alpha-D-glucopyranosyluronate trichloroacetimidate with a trisaccharide acceptor. The trisaccharide acceptor was obtained by condensation of 3-O-allyl-2,4,6-tri-O-benzoyl-beta-D-galactopyranosyl-(1-->3)-2,4,6-tri-O-benzoyl-alpha-D-galactopyranosyl trichloroacetimidate with methyl 2,3-di-O-benzoyl-beta-D-xylopyranoside, followed by deallylation. The beta-(1-->3)-linked disaccharide was prepared readily with p-methoxyphenyl 3-O-allyl-2,4,6-tri-O-benzoyl-beta-D-galactopyranoside as the key synthon. The alpha-(1-->3)-linkage was formed in considerable amount with galactose mono- and disaccharide trichloroacetimidate donors with C-2 neighboring group participation.  相似文献   

9.
Different conformations of methyl 3,6-anhydro-4-O-methyl-alpha-d-galactoside (1) and 3,6-anhydro-4-O-methylgalactitol (2) were studied by molecular mechanics (using the program mm3) and by quantum mechanical (QM) methods at the B3LYP/6-31+G( * *) and MP2/6-311++G( * *) levels, with and without solvent emulation. In 2, where the five-membered ring is free to move, two main stable conformations of this ring were found, identified as North (N) and South (S). The latter appears to be more stable, by either calculation, though the energy difference is reduced when emulating solution behavior. In order to find out the possible influence of a glycosidic bond over its shape, and to explain the marked NMR chemical shift displacements observed by opening of the ring, the adiabatic maps of two disaccharides carrying an analog of beta-galactoside linked to O-4 of 1 and 2 were generated. It was shown that the characteristics of the 3,6-AnGal terminal influence the characteristics of the map, especially at lower dielectric constants. On the other hand, different glycosidic angles also promote distinct stable conformations of the five-membered ring, changing from N to S, or even variants. Comparison with experimental results leads to the idea of highly flexible disaccharides, with variable values for both the five-membered ring and the glycosidic angles.  相似文献   

10.
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   

11.
Kim H  Jeong K  Cho KW  Paik SR  Jung S 《Carbohydrate research》2006,341(8):1011-1019
The conformational preferences of a cyclic osmoregulated periplasmic glucan of Ralstonia solanacearum (OPGR), which is composed of 13 glucose units and linked entirely via beta-(1-->2) linkages excluding one alpha-(1-->6) linkage, were characterized by molecular dynamics simulations. Of the three force fields modified for carbohydrates that were applied to select a suitable one for the cyclic glucan, the carbohydrate solution force field (CSFF) was found to most accurately simulate the cyclic molecule. To determine the conformational characteristics of OPGR, we investigated the glycosidic dihedral angle distribution, fluctuation, and the potential energy of the glucan and constructed hypothetical cyclic (CYS13) and linear (LINEAR) glucans. All beta-(1-->2)-glycosidic linkages of OPGR adopted stable conformations, and the dihedral angles fluctuated in this energy region with some flexibility. However, despite the inherent flexibility of the alpha-(1-->6) linkage, the dihedral angles have no transition and are more rigid than that in a linear glucan. CYS13, which consists of only beta-(1-->2) linkages, is somewhat less flexible than other glycans, and one of its linkages adopts a higher energy conformation. In addition, the root-mean-square fluctuation of this linkage is lower than that of other linkages. Furthermore, the potential energy of glucans increases in the order of LINEAR, OPGR, and CYS13. These results provide evidence of the existence of conformational constraints in the cyclic glucan. The alpha-(1-->6)-glycosidic linkage can relieve this constraint more efficiently than the beta-(1-->2) linkage. The conformation of OPGR can reconcile the tendency for individual glycosidic bonds to adopt energetically favorable conformations with the requirement for closure of the macrocyclic ring by losing the inherent flexibility of the alpha-(1-->6)-glycosidic linkage.  相似文献   

12.
Li A  Kong F 《Carbohydrate research》2005,340(12):1949-1962
Effective syntheses of galactose hepta-, octa-, nona-, and decasaccharides that exist in the rhizomes of Atractylodes lancea DC were achieved with 2,3,4,6-tetra-O-benzoyl-alpha-d-galactopyranosyl trichloroacetimidate (1), 4-methoxyphenyl 2,3,4-tri-O-benzoyl-beta-d-galactopyranoside (2), 6-O-acetyl-2,3,4-tri-O-benzoyl-alpha-d-galactopyranosyl trichloroacetimidate (5), 4-methoxyphenyl 6-O-acetyl-2,4-di-O-benzoyl-beta-d-galactopyranoside (22), and 4-methoxyphenyl 2,4,6-tri-O-benzoyl-beta-d-galactopyranoside (26) as the key synthons. Coupling of 2 with 1, followed by oxidative cleavage of 1-OMP and subsequent trichloroacetimidate formation gave the beta-(1-->6)-linked disaccharide donor 4. Condensation of 2 with 5 and subsequent selective deacetylation by methanolysis produced the beta-(1-->6)-linked disaccharide acceptor 7. Reaction of 7 with 4, oxidative cleavage of 1-OMP, and trichloroacetimidate formation produced the tetrasaccharide donor 9. The penta- (15), the hexa- (17), and the heptasaccharide donor 19 were synthesized similarly. Meanwhile, treatment of 1 with 22 yielded beta-(1-->3)-linked disaccharide 23 and alpha-(1-->3)-linked disaccharide 25. Oxidative cleavage of 1-OMp of 23 followed by trichloroacetimidate formation produced the disaccharide donor 24. Coupling of 26 with 24, again, gave beta-linked 27 and alpha-linked 29. Selective 6-O-deacetylation of 27 afforded the trisaccharide acceptor 28. TMSOTf-promoted condensation 28 of with the tetra- (9), penta- (15), hexa-(17), and heptasaccharide donor 19, followed by deprotection, gave the target compounds.  相似文献   

13.
An empirical force field, denoted AMB99C, has been used to study molecular properties of alpha-(1-->4)-linked carbohydrates in solution. AMB99C was parameterized using structural and energetic parameters from density functional ab initio methodology. In this work we examine the solution behavior of the beta anomer of maltose and cyclohexa-, cyclohepta-, and cyclooctaamyloses (alpha-, beta-, and gamma-cyclodextrins or alpha-, beta-, and gamma-CDs, respectively), as well as of two larger (DP 10, epsilon-CD; DP 21) cyclomaltooligosaccharides, CA10 and CA21. Experimental data used for comparison purposes include X-ray structures, small-angle scattering radius of gyration values, NMR nuclear Overhauser enhancements (NOEs), and proton coupling constants. Molecular dynamics simulations were carried out using explicit water molecules (TIP3P) to establish equilibrium populations of conformations in solution, and these results are compared with other calculated values and a variety of experimental parameters, such as average H-1-H-4' distances between the rings in beta-maltose, and the primary hydroxyl groups' conformational populations. Medium-to-large cyclomaltooligosaccharide molecules were studied to test for glucose ring puckering and stability of kinked and 'flipped' conformations. The results of the solvation studies are in excellent agreement with experimental structural parameters.  相似文献   

14.
alpha-Galactosyl epitopes are carbohydrate structures bearing an alpha-Gal-(1-->3)-Gal terminus (alpha-Gal epitopes). The interaction of these epitopes on the surface of animal cells with anti alpha-Gal antibodies in human serum is believed to be the main cause in antibody-mediated hyperacute rejection in xenotransplantation. In this paper, conformational analysis of an N-linked alpha-D-Galp-(1-->3)-beta-D-Galp-(1-->4)-beta-D-Glcp trisaccharide epitope was conducted in terms of each monosaccharide residue conformation, primary hydroxymethyl group configuration, and interglycosidic conformations. Selective 2D J-delta INEPT experiments have been carried out at three different temperatures to evaluate three-bond, long-range 13C-1H coupling constants for the crucial alpha-(1-->3) linkage. The NMR experimental data were complemented by theoretical calculations. The flexibility and dynamics of the trisaccharide have been studied by Metropolis Monte Carlo simulations. Ensemble-averaged three-bond, long-range 13C-1H coupling constants and nuclear Overhauser effects were in good agreement with the experimental data. The alpha-(1-->3) glycosidic linkage has shown a restricted flexibility as indicated by NMR spectroscopy and molecular modeling.  相似文献   

15.
Kuttel MM 《Carbohydrate research》2008,343(6):1091-1098
Four Ramachandran maps of the conformational potential of mean force (PMF) for the galactose disaccharide globobiose (alpha-D-Galp-(1-->4)-beta-D-Galp) were calculated in vacuum, explicit water, with a simple high dielectric constant and a distance-dependent dielectric coefficient, respectively. This simple model of the galactan alpha-(1-->4)-linkage is shown to be conformationally restricted, with only a small range of syn-phi/syn-psi conformations predominating at standard temperature and pressure. This has implications for the preferred conformation and chain dynamics of alpha-galactosides. In addition, comparison of the relevant PMF surfaces reveals the substitution of a high dielectric constant for explicit water solution to be a valid approximation for reproducing the minimum energy conformation of this glycosidic linkage.  相似文献   

16.
Surfactant protein D (SP-D), a C-type lectin, is an important pulmonary host defense molecule. Carbohydrate binding is critical to its host defense properties, but the precise polysaccharide structures recognized by the protein are unknown. SP-D binding to Aspergillus fumigatus is strongly inhibited by a soluble beta-(1-->6)-linked but not by a soluble beta-(1-->3)-linked glucosyl homopolysaccharide (pustulan and laminarin, respectively), suggesting that SP-D recognizes only certain polysaccharide configurations, likely through differential binding to nonterminal glucosyl residues. In this study we have computationally docked alpha/beta-D-glucopyranose and alpha/beta-(1-->2)-, alpha/beta-(1-->3)-, alpha/beta-(1-->4)-, and alpha/beta-(1-->6)-linked glucosyl trisaccharides into the SP-D carbohydrate recognition domain. As with the mannose-binding proteins, we found significant hydrogen bonding between the protein and the vicinal, equatorial OH groups at the 3 and 4 positions on the sugar ring. Our docking studies predict that alpha/beta-(1-->2)-, alpha-(1-->4)-, and alpha/beta-(1-->6)-linked but not alpha/beta-(1-->3)-linked glucosyl trisaccharides can be bound by their internal glucosyl residues and that binding also occurs through interactions of the protein with the 2- and 3-equatorial OH groups on the glucosyl ring. By using various soluble glucosyl homopolysaccharides as inhibitors of SP-D carbohydrate binding, we confirmed the interactions predicted by our modeling studies. Given the sequence and structural similarity between SP-D and other C-type lectins, many of the predicted interactions should be applicable to this protein family.  相似文献   

17.
The solution conformation of methyl α-lactoside has been studied through nmr spectroscopy and molecular mechanics calculations using the assisted model building with energy refinement (AMBER) force field. The nmr data have included nuclear Overhauser effect (NOE) measurements hot It in the laboratory and rotating frames, longitudinal relaxation times, and homonuclear and heteronuclear coupling constants. The steady-state and transient NOEs have been interpreted in terms of an ensemble average distribution of conformers, making use of the complete relaxation matrix approach. The molecular mechanics calculations have been performed at two dielectric constants [ε = 1 * r and 80 Debyes (D)] in an exhaustive way, and have been complemented with specific calculations at intermediate ε values. Relaxed energy maps and adiabatic surfaces have been generated for the different dielectric constants. The probability distribution of conformers has been estimated from these steric energy maps. Molecular dynamics simulations in vacuo have also been performed. The experimental results indicate that the β(1 → 4)-glycosidic linkage shows some fluctuations among three low energy regions, although spends ca. 85% of its lime in the region close to the global minimum. It is shown that the over estimation of the electrostatic contributions in AMBER is responsible for the failure of this force field to explain the experimental results when Used at low dielectric constant (ε < 20 D). The matching between the expected and observed facts increases for ε > 40 D. Different conditions have been tested to perform temperature constant molecular dynamics simulations in vacuo, which have indicated that, when used without explicit solvent, this force field should only be employed in a qualitatively way when analyzing dynamical properties of oligosaccharides. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
The preferred conformations and conformational flexibilities of the trisaccharides alpha-D-Glcp-(1-->2)-beta-D-Glcp-(1-->3)-alpha-D-Glcp-OMe (I) and alpha-D-Glcp-(1-->3)[beta-D-Glcp-(1-->4)]-alpha-D-Glcp-OMe (II) in aqueous solution were determined using nuclear magnetic resonance (NMR) spectroscopy, molecular dynamics (MD) and Langevin dynamics (LD) simulations, and hydrodynamics calculations. Both trisaccharides have a vicinal substitution pattern in which long range (nonsequential) interactions may play an important role. LD simulation at 600 K indicated that the all-syn conformation predominated, though other conformations were apparent. NOE data and MD and LD simulations at 298 K all indicated that trisaccharide I is a single all-syn conformer in solution. Given that previous studies showed evidence of anti-conformers in beta-D-Glcp-(1-->2)-beta-D-Glcp-(1-->3)-alpha-D-Glcp-OMe, this result provides an example of how changing the anomeric configuration of one residue from beta to alpha can make an oligosaccharide more rigid. Discrepancies in inter-ring distances obtained by experiment and by simulation of the all-syn conformer suggest the presence of an anti-psi conformation at the beta-(1-->4)-linkage for II. A combined analysis of measured and calculated translational diffusion constants and (13)C T(1) relaxation times yield order parameters of 0.9 for each trisaccharide. This implies that any interconversion among conformations is significantly slower than tumbling. Anisotropies of approximately 1.6 and 1.3 calculated for I and II, respectively, are consistent with the observed relatively flat T(1) profiles because the tumbling is not in the motional narrowing regime.  相似文献   

19.
C Mukhopadhyay  C A Bush 《Biopolymers》1991,31(14):1737-1746
Molecular dynamics simulations without explicit inclusion of solvent molecules have been performed to study the motions of Lewisa and Lewisb blood group oligosaccharides, and two blood group A tetrasaccharides having type I and type II core chains. The blood group H trisaccharide has also been studied and compared with the blood group A type II core chain. The potential energy surface developed by Rasmussen and co-workers was used with the molecular mechanics code CHARMM. The lowest energy minima of the component disaccharide fragments were obtained from conformational energy mapping. The lowest energy minima of these disaccharide fragments were used to build the tri- and tetrasaccharides that were further minimized before the actual heating/equilibration and dynamics simulations. The trajectories of the disaccharide fragments, e.g., Fuc alpha- (1----4)GlcNAc, Gal beta-(1----4)GlcNAc, etc., show transitions among various minima. However, the oligosaccharides were found to be dynamically stable and no transitions to other minimum energy conformations were observed in the time series of the glycosidic dihedral angles even during trajectories as long as 300 ps. The stable conformations of the glycosidic linkages in the oligosaccharides are not necessarily the same as the minimum energy conformation of the corresponding isolated disaccharides. The average fluctuations of the glycosidic angles in the oligosaccharides were well within the range of +/- 15 degrees. The results of these trajectory calculations were consistent with the relatively rigid single-conformation models derived for these oligosaccharides from 1H-nmr data.  相似文献   

20.
The alpha-Gal trisaccharide Gal(alpha)(1-->3)Galbeta(1-->4)GlcNAc 11 was synthesized on a homogeneously soluble polymeric support (polyethylene glycol, PEG) by use of a multi-enzyme system consisting of beta-1,4-galactosyltransferase (EC 2.4.1.38), alpha-1,3-galactosyltransferase (EC 2.4.1.151), sucrose synthase (EC 2.4.1.13) and UDP-glucose-4-epimerase (EC 5.1.3.2). In addition workup was simplified by use of dia-ultrafiltration. Thus the advantages of classic chemistry/enzymology and solid-phase synthesis could be united in one. Subsequent hydrogenolytic cleavage afforded the free alpha-Gal trisaccharide.  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