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.
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   

2.
Stortz CA 《Carbohydrate research》2002,337(21-23):2311-2323
The adiabatic conformational surfaces of several beta-linked disaccharides, which correspond to the repeating structures of carrageenans, were calculated using the MM3 force-field. The studies were carried out on the disaccharide beta-D-Galp-(1 --> 4)-alpha-D-Galp and eight sulfated derivatives, as well as on carrabiose (beta-D-Galp-(1 --> 4)-3,6-An-alpha-D-Galp) and five sulfated derivatives. The presence of 3,6-anhydrogalactose does not change the main features of the maps, although it increases the flexibility of the glycosidic linkage. Sulfation neither produces a striking effect on the map shape, nor a shift on the global minimum, which always remains with psi (theta(C-1'-O-4-C-4C-5)) in trans orientation, and phi (theta(O-5'-C-1'-O-4-C-4)) with a value close to -80 degrees. This effect differs from that occurring on the alpha linkage of equivalent disaccharides, for which the sulfation pattern on the beta-galactose unit shifts the global minima to different positions. A reduction in the flexibility (originated in a deepening of the global minimum well) is observed by sulfation on position 2 of the beta-D-galactose unit, and by sulfation of position 6 of the alpha-D-galactose unit (when the beta-D-galactose unit is 4-sulfated). Within the compounds containing 3,6-anhydrogalactose, the effect of sulfation is even less noticeable. The calculated low-energy regions on carrabiose derivatives agree with X-ray diffraction data on carrageenan fibers and on peracetylated carrabiose dimethyl acetal, and with NOE calculations carried out on kappa-carrabiose.  相似文献   

3.
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   

4.
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   

5.
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   

6.
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   

7.
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   

8.
The synthesis of the trisaccharide allyl 2-O-(alpha-L-arabinofuranosyl)-6-O-(alpha-D-mannopyranosyl)-beta-D-mannopyra-noside is reported. Stereoselective glycosylation at C-6 of a non-protected allyl beta-mannoside with the acetylated alpha-D-mannosyl bromide gave the alpha-(1 --> 6)-disaccharide as the main product and the crystalline 3,6-branched trisaccharide as minor compound. Further glycosylation of the 2,3 diol (1 --> 6) disaccharide with L-arabinofuranosyl bromide furnished a mixture of 3-O- and 2-O-alpha-L-Ara-trisaccharides from which the title compound was isolated.  相似文献   

9.
The polyagglutinable erythrocytes NOR contain unusual neutral glycolipids reactive with anti-NOR antibodies. The disaccharide alpha-D-Galp-(1-->4)-D-GalpNAc and the trisaccharide alpha-D-Galp-(1-->4)-beta-D-GalpNAc-(1-->3)-D-Gal corresponding to the non-reducing end of a NOR glycolipid (NOR1) were chemically synthesized. The syntheses were based on a common (1-->4)-beta-D-GalNAc precursor, and utilized benzyl glycoside and benzyl ether functions for persistent blocking of hydroxyls. The alpha-D-Galp-(1-->4)-beta-D-GalpNAc structural element has been found only recently in Nature, and derivatives thereof have not been synthesized before. Both the synthesized oligosaccharides inhibited specifically human anti-NOR antibodies, the trisaccharide being 300 times more active than the disaccharide.  相似文献   

10.
Zhu Y  Chen L  Kong F 《Carbohydrate research》2002,337(3):207-215
A highly concise and effective synthesis of the mannose octasaccharide of the N-linked glycan in the adhesion domain of human CD2 was achieved via TMSOTf-promoted selective 6-glycosylation of a trisaccharide 4,6-diol acceptor with a pentasaccharide donor, followed by deprotection. The pentasaccharide was constructed by selective 3,6-diglycosylation of 1,2-O-ethylidene-beta-D-mannopyranose with 2-O-acetyl-3,4,6-tri-O-benzoyl-alpha-D-mannopyranosyl-(1-->2)-3,4,6-tri-O-benzoyl-alpha-D-mannopyranosyl trichloroacetimidate, while the trisaccharide was obtained by selective 3-O-glycosylation of allyl 4,6-O-benzylidene-alpha-D-mannopyranoside with the same disaccharide trichloroacetimidate, followed by debenzylidenation. The mannose hexasaccharide antigenic factor 13b was synthesized by condensation of a trisaccharide donor, 2-O-acetyl-3,4,6-tri-O-benzoyl-alpha-D-mannopyranosyl-(1-->2)-3,4,6-tri-O-benzoyl-alpha-D-mannopyranosyl-(1-->3)-4,6-di-O-acetyl-2-O-benzoyl-alpha-D-mannopyranosyl trichloroacetimidate, with a trisaccharide acceptor, methyl 3,4,6-tri-O-benzoyl-alpha-D-mannopyranosyl-(1-->2)-3,4,6-tri-O-benzoyl-alpha-D-mannopyranosyl-(1-->2)-3,4,6-tri-O-benzoyl-alpha-D-mannopyranoside, followed by deprotection.  相似文献   

11.
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   

12.
Heparin and heparan sulfate fragments, obtained by bacterial heparinase and heparitinases, bearing an unsaturation at C4-C5 of the uronic acid moiety, are able to produce up to 80% reduction of the cytosolic calcium of smooth muscle cell lines. Unsaturated disaccharides from chondroitin sulfate, dermatan sulfate, and hyaluronic acid are inactive, indicating that, besides the unsaturation of the uronic acid, a vicinal 1 --> 4 glycosidic linkage is needed. An inverse correlation between the molecular weight and activity is observed. Thus, the ED(50) of the N-acetylated disaccharide derived from heparan sulfate (430 Da) is 88 microm compared with 250 microm of the trisulfated disaccharide (650 Da) derived from heparin. Except for enoxaparin (which contains an unsaturation at the non-reducing end and 1 --> 4 glycosidic linkage), other low molecular weight heparins and native heparin are practically inactive in reducing the cytosolic calcium levels. Thapsigargin (sarcoplasmic reticulum Ca(2+)-ATPase inhibitor), vanadate (cytoplasmic membrane Ca(2+)-ATPase inhibitor), and nifedipine and verapamil (Ca(2+) channel antagonists) do not interfere with the effect of the trisulfated disaccharide upon the decrease of the intracellular calcium. A significant decrease of the activity of the trisulfated disaccharide is observed by reducing extracellular sodium, suggesting that the fragments might act upon the Na(+)/Ca(2+) exchanger promoting the extrusion of Ca(2+). This was further substantiated by binding experiments and circular dichroism analysis with the exchanger inhibitor peptide.  相似文献   

13.
An acidic O-specific polysaccharide containing D-glucuronic acid (D-GlcA), 2,3-diacetamido-2,3-dideoxy-D-glucuronic acid (D-GlcNAc3NAcA), 2,3-diacetamido-2,3-dideoxy-D-mannuronoyl-L-alanine (D-ManNAc3NAcA6Ala), and 2-acetamido-2,4, 6-trideoxy-4-[(S)-3-hydroxybutyramido]-D-glucose (D-QuiNAc4NAcyl) was obtained by mild acid degradation of the lipopolysaccharide of the bacterium Pseudoalteromonas sp. KMM 634 followed by gel-permeation chromatography. The polysaccharide was cleaved selectively with a new solvolytic agent, trifluoromethanesulfonic acid, to give a disaccharide and a trisaccharide with D-GlcNAc3NAcA at the reducing end. The borohydride-reduced oligosaccharides and the initial polysaccharide were studied by GLC-MS and 1H- and 13C-NMR spectroscopy, and the following structure of the linear tetrasaccharide repeating unit of the polysaccharide was established: -->3)-alpha-D-QuipNAc4Ac4NAcyl-(1-->4)-beta-D-ManpNAc3NAcA6Ala+ ++-(1-->4)-b eta-D-GlcpNAc3NAc3NAcA-(1-->4)-beta-D-GlcpA-(1-->.  相似文献   

14.
The chemical synthesis of the bacterial O-antigen from Salmonella serogroup E1, 3-O-(4-O-beta-D-mannopyranosyl-alpha-L-rhamnopyranosyl)-alpha-D-galactos e, presents a particular challenge because it contains a beta-(1-->4) mannosidic linkage to L-rhamnose. We report a chemoenzymatic synthesis of this crucial antigenic material which culminates in the enzymatic formation of the critical beta-mannosyl connection catalyzed by Salmonella GDP-alpha-D-Man:alpha Rha1-->3 alpha Gal-PP-Und beta-(1-->4)-mannosyltransferase (ManT beta 4). In comparison with previous synthetic routes, this method is advantageous since it utilizes intermediates, available in significant yield, which can be readily derivatized from the reducing end to present flexibility for analog construction, while the enzymatic construction of the Man1-->4Rha glycosidic bond is both rapid and occurs in high yield. Furthermore, the reported spectroscopic and enzymatic structural characterization of the trisaccharide product furnishes the first indisputable functional link between wbaO and ManT beta 4 and clearly sets the stage for the future mechanistic study and exploitation of this fascinating glycocatalyst.  相似文献   

15.
Deprotection of the fully blocked disacharide allyl O-(2-amino-4,6-O-benzylidene-3-O-[(R)-1-carboxyethyl]-2-deoxy-beta-D-glucopyranosyl-1',2-lactam)-(1-->4)-2-acetamido-3,6-di-O-benzyl-2-deoxy-beta-D-glucopyranoside by selective de-O-allylation and parallel removal of the benzylidene and O-benzyl groups is described. The resulting beta-muramyl lactam-(1-->4)-GlcNAc disaccharide is characterised as the per-O-acetylated derivative by 1H and 13C NMR spectroscopy and X-ray structure analysis. Conformational analysis about glycosidic bond of repeating units of bacterial spore cortex is based on experimental data and molecular modelling.  相似文献   

16.
(1-->3)-alpha-glucanases catalyze the hydrolysis of fungal cell wall (1-->3)-alpha-glucan, and function during cell division of yeasts containing this cell wall component or act in mycoparasitic processes. Here, we characterize the mechanism of action of the (1-->3)-alpha-glucanase MutAp from the mycoparasitic fungus Trichoderma harzianum. We observed that MutAp releases predominantly beta-glucose upon hydrolysis of crystalline (1-->3)-alpha-glucan, indicating inversion of the anomeric configuration. After having identified (1-->3)-alpha-glucan tetrasaccharide as the minimal substrate for MutAp, we showed that reduced (1-->3)-alpha-glucan pentasaccharide is cleaved into a trisaccharide and a reduced disaccharide, demonstrating that MutAp displays endo-hydrolytic activity. We propose a model for the catalytic mechanism of MutAp, whereby the enzyme breaks an intrachain glycosidic linkage of (1-->3)-alpha-glucan, and then continues its hydrolysis towards the non-reducing end by releasing beta-glucose residues in a processive manner.  相似文献   

17.
The production of an adiabatic map for a di- or trisaccharide requires the generation of many relaxed maps, ideally 59,049 for a disaccharide or 4,782,969 for a trisaccharide composed by hexose residues, due to a combination of exocyclic angle torsions. As the production of this amount of maps is usually ruled out for time considerations, different approaches were exploited. When working at low dielectric constants, starting points originated in cooperative hydrogen bonds through the rings are usually sufficient to produce an adiabatic map, but at higher dielectric constants those circuits are meaningless, and many low-energy conformers appear in each energy well. Herein, different conformations of four disaccharides (beta-4-linked mannobiose, and three galactobioses, linked alpha-(1-->3), alpha-(1-->4), and beta-(1-->4)) and one trisaccharide (beta-4-linked mannotriose) were minimized using mm3 at epsilon = 80, and the difference in energy produced by changes in torsional angles was recorded. A remarkable additive effect was found to occur when the exocyclics were gathered in groupings of two or three neighboring angles. Thus, in most cases, each grouping can be studied separately, and the minimum energy conformers can be predicted without the need of resorting to thousands of calculations. In some cases where two protons of different groups show steric interactions in some specific conformations, small deviations of the additivity were encountered. Anyway, a complex system with many variables can be transformed in one with many fewer variables, thus simplifying further studies. An attempt to calculate the same effect at epsilon = 3 shows that hydrogen bonding and electrostatic interactions make impossible to find those additive effects, thus precluding its utilization at such low dielectric constants.  相似文献   

18.
Mild acid degradation of lipopolysaccharides from Pseudomonas aeruginosa O10a and O10a,b (Lányi classification) resulted in O-specific polysaccharides built up of trisaccharide repeating units containing 2-acetamido-2,6-dideoxy-D-glucose (N-acetylquinovosamine, DQuiNAc), 2-acetamido-2,6-dideoxy-D-galactose (N-acetylfucosamine, DFucNAc), and 5-acetamido-3,5,7,9-tetradeoxy-7-[(R)-3-hydroxybutyramido] -L-glycero-L-manno-nonulosonic acid. The latter is a di-N-acyl derivative of a new sialic-acid-like sugar which was called by us pseudaminic acid (PseN2). A 3-hydroxybutyric acid residue was also found in natural carbohydrates for the first time. In the O10a,b polysaccharide pseudaminic acid carried an O-acetyl group at position 4. For selective cleavage of the O10a polysaccharide, solvolysis with hydrogen fluoride was employed which, owing to the relatively high stability of the glycosidic linkage of pseudaminic acid, led to the disaccharide with this sugar on the non-reducing terminus. Performing the solvolysis in methanol afforded the methyl glycoside of this disaccharide which proved to be more advantageous for further analysis. Carboxyl-reduction made the glycosidic linkage of pseudaminic acid extremely labile, and mild acid hydrolysis of the carboxyl-reduced 010a polysaccharide afforded the trisaccharide with a ketose derivative on the reducing terminus. Establishing the structure of the oligosaccharide fragments obtained and interpreting the 13C nuclear resonance spectra of the polysaccharides allowed to determine the following structure for their repeating units: (formula: see text) In the polysaccharides the N-acetylquinovosamine residue is attached not to pseudaminic acid itself, but to its N-acyl substituent, 3-hydroxybutyryl group, and thus the monomers are linked via both glycosidic and amidic linkages.  相似文献   

19.
Zhao W  Kong F 《Carbohydrate research》2005,340(10):1673-1681
Beta-D-Xylp-(1-->2)-alpha-D-Manp-(1-->3)-[beta-D-Xylp-(1-->2)][beta-D-Xylp-(1-->4)]-alpha-D-Manp-(1-->3)-[beta-D-Xylp-(1-->4)]-alpha-D-Manp, the fragment of the exopolysaccharide from Cryptococcus neoformans serovar C, was synthesized as its methyl glycoside. Thus, chloroacetylation of allyl 3-O-acetyl-4,6-O-benzylidene-alpha-D-mannopyranoside (1) followed by debenzylidenation and selective 6-O-benzoylation afforded allyl 2-O-chloroacetyl-3-O-acetyl-6-O-benzoyl-alpha-D-mannopyranoside (4). Glycosylation of 4 with 2,3,4-tri-O-benzoyl-D-xylopyranosyl trichloroacetimidate (5) furnished the beta-(1-->4)-linked disaccharide 6. Dechloroacetylation gave the disaccharide acceptor 7 and subsequent coupling with 5 produced the trisaccharide 8. Deacetylation of 8 gave the trisaccharide acceptor 9 and subsequent coupling with a disaccharide 10 produced the pentasaccharide 11. Reiteration of deallylation and trichloroacetimidate formation from 11 yielded the pentasaccharide donor 12. Coupling of a disaccharide acceptor 13 with 12 afforded the heptasaccharide 14. Subsequent deprotection gave the heptaoside 16, while selective 2-O-deacetylation of 14 gave the heptasaccharide acceptor 15. Condensation of 15 with glucopyranosyluronate imidate 17 did not yield the expected octaoside, instead, an orthoester product 18 was obtained. Rearrangement of 18 did not give the target octaoside; but produced 15. Meanwhile, there was no reaction between 15 and the glycosyl bromide donor 19.  相似文献   

20.
Alternansucrase (EC 2.4.1.140) is a d-glucansucrase that synthesizes an alternating alpha-(1-->3), (1-->6)-linked d-glucan from sucrose. It also synthesizes oligosaccharides via d-glucopyranosyl transfer to various acceptor sugars. Two of the more efficient monosaccharide acceptors are D-tagatose and L-glucose. In the presence of d-tagatose, alternansucrase produced the disaccharide alpha-d-glucopyranosyl-(1-->1)-beta-D-tagatopyranose via glucosyl transfer. This disaccharide is analogous to trehalulose. We were unable to isolate a disaccharide product from L-glucose, but the trisaccharide alpha-D-glucopyranosyl-(1-->6)-alpha-d-glucopyranosyl-(1-->4)-l-glucose was isolated and identified. This is analogous to panose, one of the structural units of pullulan, in which the reducing-end D-glucose residue has been replaced by its L-enantiomer. The putative L-glucose disaccharide product, produced by glucoamylase hydrolysis of the trisaccharide, was found to be an acceptor for alternansucrase. The disaccharide, alpha-D-glucopyranosyl-(1-->4)-L-glucose, was a better acceptor than maltose, previously the best known acceptor for alternansucrase. A structure comparison of alpha-D-glucopyranosyl-(1-->4)-L-glucose and maltose was performed through computer modeling to identify common features, which may be important in acceptor affinity by alternansucrase.  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