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 0 毫秒
1.
The polymerization of 1,2-anhydro-3,4,6-tri-O-benzyl-β-D -mannopyranose proceeds in the presence of Lewis acids, cationic coordination catalysts, and strong bases. Debenzylation of the products yields oligomeric saccharides or low polymers. Polymerization in toluene by means of potassium alkoxide complexed with crown ethers leads to essentially stereoregular (1 → 2)-α-D -mannopyranan. The original derivatives have been characterized by optical rotation, viscosity, molecular weight, gel permeation chromatography, and spectrometry. The free polysaccharides have been characterized by optical rotation, molecular weight, and 1H- and 13C-nmr spectrometry and compared to yeast mannan hydrolysate oligomers.  相似文献   

2.
Poly-α-(1 → 6)-anhydro-D -glucopyranose has been synthesized by the phosphorus pentafluoride-catalyzed polymerization of l,6-anhydro-2,3,4-tri-O-benzyl-β-D -glucopyranose and subsequent debenzylation. Physical characterization establishes its high polymeric character and stereoregularity.  相似文献   

3.
The physical dimensions of several (1 → 6) branched (1 → 3) -β-D -glucan samples obtained from different organisms and their derivatives have been studied by electron microscopy, light scattering measurements, viscometry, and gel permeation chromatography. The electron micrographs indicate that in most samples these biopolymers are adequately described as linear worm-like coils. A sample reconstituted from alkaline media appeared as a blend of the linear, circular, and aggregated polymer morphologies. The average mass per unit length, ML = Mw/Lw for the macroscopically linear samples, was estimated to be 2100 ± 200 g mol?1 nm?1. The parameter mL was determined from the contour lengths obtained by electron microscopy and the molecular weight by light scattering measurements. The observed ML was consistent with the triple-helical structure reported from x-ray diffraction studies and observed degree of side-chain substitution. From the molecular snapshots shown in the electron micrographs, the persistence lengths of these β-D -glucans were determined to be 140 ± 30 nm. The experimentally determined intrinsic viscosities were consistent with these estimates of ML and persistence length. Comparison of the molecular weight distributions obtained from gel permeation chromatography and those deduced from the electron micrographs indicates that number and weight average contour lengths are more reliable than z and z + 1 averages. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
Condensation of 2,4,6-tri-O-acetyl-3-deoxy-3-fluoro-α- -galactopyranosyl bromide (3) with methyl 2,3,4-tri-O-acetyl-β- -galactopyranoside (4) gave a fully acetylated (1→6)-β- -galactobiose fluorinated at the 3′-position which was deacetylated to give the title disaccharide. The corresponding trisaccharide was obtained by reaction of 4 with 2,3,4-tri-O-acetyl-6-O-chloroacetyl-α- -galactopyranosyl bromide (5), dechloroacetylation of the formed methyl O-(2,3,4-tri-O-acetyl-6-O-chloroacetyl-β- -galactopyranosyl)-(1→6)- 2,3,4-tri-O-acetyl-β- -galactopyranoside to give methyl O-(2,3,4-tri-O-acetyl-β- -galactopyranosyl)-(1→6)-2,3,4-tri-O-acetyl-β- -galactopyranoside (14), condensation with 3, and deacetylation. Dechloroacetylation of methyl O-(2,3,4-tri-O-acetyl-6-O-chloroacetyl-β- -galactopyranosyl)-(1→6)-O-(2,3,4-tri-O-acetyl- β- -galactopyranosyl)-(1→6)-2,3,4-tri-O-acetyl-β- -galactopyranoside, obtained by condensation of disaccharide 14 with bromide 5, was accompanied by extensive acetyl migration giving a mixture of products. These were deacetylated to give, crystalline for the first time, the methyl β-glycoside of (1→6)-β- -galactotriose in high yield. The structures of the target compounds were confirmed by 500-MHz, 2D, 1H- and conventional 13C- and 19F-n.m.r. spectroscopy.  相似文献   

5.
6.
Non-cellulosic β-glucans are now recognized as potent immunological activators, and some are used clinically in China and Japan. These β-glucans consist of a backbone of glucose residues linked by β-(1→3)-glycosidic bonds, often with attached side-chain glucose residues joined by β-(1→6) linkages. The frequency of branching varies. The literature suggests β-glucans are effective in treating diseases like cancer, a range of microbial infections, hypercholesterolaemia, and diabetes. Their mechanisms of action involve them being recognized as non-self molecules, so the immune system is stimulated by their presence. Several receptors have been identified, which include: dectin-1, located on macrophages, which mediates β-glucan activation of phagocytosis and production of cytokines, a response co-ordinated by the toll-like receptor-2. Activated complement receptors on natural killer cells, neutrophils, and lymphocytes, may also be associated with tumour cytotoxicity. Two other receptors, scavenger and lactosylceramide, bind β-glucans and mediate a series of signal pathways leading to immunological activation. Structurally different β-glucans appear to have different affinities toward these receptors and thus generate markedly different host responses. However, the published data are not always easy to interpret as many of the earlier studies used crude β-glucan preparations with, for the most part, unknown chemical structures. Careful choice of β-glucan products is essential if their benefits are to be optimized, and a better understanding of how β-glucans bind to receptors should enable more efficient use of their biological activities.  相似文献   

7.
8.
A (1 → 3)-β-glucan 3-glucanohydrolase (EC 3.2.1.39) has been purified approx. 190-fold from extracts of germinating barley. The enzyme has an apparent Mr 32 000, a pI of 8.6, and a pH optimum of 5.6. Analysis of hydrolysis products released from the (1 → 3)-β-glucan, laminarin, shows that the enzyme is an endohydrolase. Sequence analysis of the 46 NH2-terminal amino acids of the (1 → 3)-β-glucanase reveals 54% positional identity with barley (1 → 3,1 → 4)-β-glucanases (EC 3.2.1.73) and suggests a common evolutionary origin for these two classes of β-glucan endohydrolases. The barley (1 → 3)-β-glucanase also exhibits significant similarity with a (1 → 3)-β-glucanase from tobacco.  相似文献   

9.
Sequential tritylation, benzoylation, and detritylation of methyl 3-deoxy-3-fluoro-β- -galactopyranoside gave crystalline methyl 2,4-di-O-benzoyl-3-deoxy-3-fluoro-β- -galactopyranoside (9), which was used as the initial nucleophile in the synthesis of the target oligosaccharide (16). Treatment of 9 with 2,3,4-tri-O-benzoyl-6-O-bromoacetyl-α- -galactopyranosyl bromide gave the corresponding disaccharide derivative 13, having a selectively removable blocking group at O-6′. Debromoacetylation of 13 afforded the disaccharide nucleophile 14 which, when treated with 2,4,6-tri-O-benzoyl-3-deoxy-3-fluoro-α- -galactopyranosyl bromide, gave the fully protected trisaccharide 15. Debenzoylation of 15 gave the title glycoside 16. Condensation reactions were performed with silver trifluoromethane-sulfonate as a promoter in the presence of sym-collidine under base-deficient conditions, and gave excellent yields of the desired β-(trans)-products. Analyses of the 1H- and 13C-n.m.r. spectra, as well as determination of the JCF and JHF coupling constants, were made by using various one- and two-dimensional n.m.r. techniques.  相似文献   

10.
A chiroptical method of conformational analysis is applied to linear (1 → 3)-β-D -glucans and the dimeric analogues β- and α-laminaribioside. The method is based on a recently developed semiempirical calculational model for saccharide optical activity. We conclude that disaccharide conformational energy maps in the literature represent the effective potential energy surface in aqueous solution well. The positive optical rotation observed with long chains in dilute alkaline solution is not characteristic of any single–chain conformation, and must reflect chain association.  相似文献   

11.
13C-nmr has been employed to probe the molecular conformation and crystal structure of (1 → 6)-β-D -glucan (pustulan) in the solution, gel, and solid states. CP/MAS 13C-nmr spectra recorded for partially crystalline solid pustulan display a resonance near 82 ppm that is absent in solution spectra. The intensity and peak width of this resonance were found to depend on relative crystallinity as determined by x-ray diffraction. CP/MAS spectra of aqueous pustulan gels also exhibit the 82-ppm resonance, suggesting that the gelation mechanism may involve microcrystalline junction zones. Since the 82-ppm resonance is absent in the CP/MAS spectrum of the (1 → 6)-β-linked dimer gentiobiose, we tentatively conclude the crystal structure of this dimer does not adequately model the yet undetermined structure of pustulan.  相似文献   

12.
O-α- -Rhamnopyranosyl-(1→3)- -rhamnopyranose (19) and O-α- -rhamnopyranosyl-(1→2)- -rhamnopyranose were obtained by reaction of benzyl 2,4- (7) and 3,4-di-O-benzyl-α- -rhamnopyranoside (8) with 2,3,4-tri-O-acetyl-α- -rhamnopyranosyl bromide, followed by deprotection. The per-O-acetyl α-bromide (18) of 19 yielded, by reaction with 8 and 7, the protected derivatives of the title trisaccharides (25 and 23, respectively), from which 25 and 23 were obtained by Zemplén deacetylation and catalytic hydrogenolysis, With benzyl 2,3,4-tri-O-benzyl-β- -galactopyranoside, compound 18 gave an ≈3:2 mixture of benzyl 2,3,4-tri-O-benzyl-6-O-[2,4-di-O-acetyl-3-O-(2,3,4-tri-O-acetyl-α- -rhamnopyranosyl)-α- -rhamnopyranosyl]-β- -galactopyranoside and 4-O-acetyl-3-O-(2,3,4-tri-O-acetyl-α- -rhamnopyranosyl)-β- -rhamnopyranose 1,2-(1,2,3,4-tetra-O-benzyl-β- -galactopyranose-6-yl (orthoacetate). The downfield shift at the α-carbon atom induced by α- -rhamnopyranosylation at HO-2 or -3 of a free α- -rhamnopyranose is 7.4-8.2 p.p.m., ≈1 p.p.m. higher than when the (reducing-end) rhamnose residue is benzyl-protected (6.6-6.9 p.p.m.). α- -Rhamnopyranosylation of HO-6 of gb- -galactopyranose deshields the C-6 atom by 5.7 p.p.m. The 1 2-orthoester ring structure [O2,C(me)OR] gives characteristic resonances at 24.5 ±0.2 p.p.m. for the methyl, and at 124.0 ±0.5 p.p.m. for the quaternary, carbon atom.  相似文献   

13.
The enzymatic degradation of β(1 → 4) xylan single crystals with xylanases was investigated by electron microscopy and electron diffraction. The enzyme attack takes place at the edge of the crystals and progresses towards their centers. This is consistent with an endo-enzyme mechanism, where the enzyme interacts essentially with the accessible xylan chains located at the crystal periphery.  相似文献   

14.
The reaction of 2,3-di-O-acetyl-4-O-benzyl-α,β-d-xylopyranosyl bromide (2) with methyl 2,3-di-O-acetyl-β-d-xylopyranoside gave methyl O-(2,3-di-O-acetyl-4-O-benzyl-β-d-xylopyranosyl)-(1→4)-2,3-di-O-acetyl-β-d-xylopyranoside (22). Catalytic hydrogenolysis of 22 exposed HO-4′ which was then condensed with 2. This sequence of reactions was repeated three more times to afford, after complete removal of protecting groups, a homologous series of methyl β-glycosides of (1→4)-β-d-xylo-oligosaccharides. 13C-N.m.r. spectra of the synthetic methyl β-glycosides (di- to hexa-saccharide) are presented together with data for six other, variously substituted, homologous series of (1→4)-d-xylo-oligosaccharides.  相似文献   

15.
The crystal structures of (1→3)-α-d-glucan triacetates were studied by X-ray diffraction measurements on fibre diagrams. The oriented films annealed in water at high temperature were of higher crystallinity and occurred as two crystalline polymorphs (GTA I and GTA II) depending on the samples and also the annealing temperature. All reflections in GTA I were indexed with a pseudo-orthorhombic unit cell with a = 1·753, b = 3·018 and c(fibre axis) = 1·205 nm. From the fibre repeat data coupled with the density data and the presence of only the (003) reflection on the meridian, an extended three-fold helical structure was proposed. Although some reflections in GTA II split from the layer lines, the basic unit cell was a monoclinic system with a = 1·685, b = 3·878, c (fibre axis) = 1·210 nm and γ = 112·2°. A similar three-fold structure to GTA I was proposed from the almost identical fibre repeat and the conformational analysis on (1→3)-α-d-glucan. It was concluded that, on acetylation, the d-glucan structure changed from the fully extended two-fold helix to the extended three-fold accompanied by some extent of chain shrinking.  相似文献   

16.
Botryosphaeran, a (13;16)-β-d-glucan produced by Botryosphaeria rhodina MAMB-05, was found to be present in a triple helix conformation from helix–coil transition studies using Congo Red. The triple helix conformation was disrupted at increasing alkali concentrations. Conformational changes were also observed using phenanthrene as a fluorescent probe, and the fluorescence intensity decreased 80% in the presence of dimethyl sulfoxide. The results confirmed the triple helix conformation of botryosphaeran, an important property manifesting biological response modifying activity.  相似文献   

17.
The emergence of multi-drug resistant (MDR) strains of Mycobacterium tuberculosis (MTB) and the continuing pandemic of tuberculosis emphasizes the urgent need for the development of new anti-tubercular agents with novel drug targets. The recent structural elucidation of the mycobacterial cell wall highlights a large variety of structurally unique components that may be a basis for new drug development. This publication describes the synthesis, characterization, and screening of several octyl Galf(β,1→5)Galf and octyl Galf(β,1→6)Galf derivatives. A cell-free assay system has been utilized for galactosyltransferase activity using UDP[14C]Galf as the glycosyl donor, and in vitro inhibitory activity has been determined in a colorimetric broth microdilution assay system against MTB H37Ra and three clinical isolates of Mycobacterium avium complex (MAC). Certain derivatives showed moderate activities against MTB and MAC. The biological evaluation of these disaccharides suggests that more hydrophobic analogues with a blocked reducing end showed better activity as compared to totally deprotected disaccharides that more closely resemble the natural substrates in cell wall biosynthesis.  相似文献   

18.
The mutant strain M5N1 CS of Rhizobium meliloti produces, in a Rhizobium complete medium supplemented with fructose and sucrose, a partially acetylated homopolymer of D -glucuronic acid residues linked β-(1 → 4). This polysaccharide forms thermoreversible gels with monovalent salts and thermally stable gels with divalent salts. In order to define the different levels of structural characterization, modeling simulations were performed for both the regular (1 → 4)-β-D -glucuronan and the acetylated derivatives. This required the evaluation of the accessible conformational space for the 16 disaccharides. Detailed conformational analysis was accomplished using the flexible residue of the MM3 molecular mechanics procedure and the results were used to access the configurational statistics of representative polysaccharide chains. Within the potential energy surfaces calculated for each disaccharide, several low energy conformers can be identified. When these conformations are extrapolated to regular polysaccharide structures, they generate polymers with right- and left-handed chirality along with a 2-fold axis. This later arrangement (n = 2, h = 5.16 Å) closely corresponds to that derived from a fiber x-ray diffraction investigation. The insertion of acetyl groups induces changes in the helical features of the polymer. As for the simulation of the configurational properties of (1 → 4)-β-D -glucuronan, an extended disordered chain having a persistence length of 105 Å (corresponding to 22 monomers) is predicted. This agrees with previous conclusions derived from solution study. The inclusion of varying amounts of acetyl groups only slightly perturbs the calculated persistence length. © 1998 John Wiley & Sons, Inc. Biopoly 45: 165–175, 1998  相似文献   

19.
Cellotriosyl and cellotetraosyl residues, linked by single (1→3)-β-linkages, account for more than 90% of the 40°C water-soluble (1→3), (1→4)-β-d-glucan from barley flour. We have analysed their sequence dependence by treating the polymer as a two-state Markov chain with stationary distribution. Quantitation of the penultimate oligosaccharides released during hydrolysis of the (1→3), (1→4)-β-d-glucan with (1→3), (1→4)-β-d-glucan 4-glucanohydrolase (EC 3.2.1.73) by analytical gel filtration chromatography enabled the relative abundance of two adjacent cellotriosyl, two adjacent cellotetraosyl and adjacent cellotetraosyl/cellotriosyl residues to be estimated and the sequence dependence to be evaluated.Within the theoretical and practical constraints of the method it is concluded that the cellotriosyl and cellotetraosyl residues are arranged in an essentially independent (random) fashion. Thus, any mechanism proposed for the biosynthesis of the molecule should explain this apparently random distribution of cellotriosyl and cellotetraosyl residues as well as the presence, in relatively low frequency, of blocks of up to 10 or more adjacent (1→4)-linkages.  相似文献   

20.
Molecular mechanics and dynamics calculations were carried out on the disaccharides α-L-Rhap-(1 → 2)-α-L-Rhap-(1 → OMe) (1) and α-L-Rhap-(1 → 3)-α-L-Rhap-(1 OMe) (2), and the trisaccharide α-L-Rhap-(1 → 2)-α-L-Rhap-(1 → 3)-α-L-Rhap-(1 → OMe) (3). The semiflexible conformational behavior of these molecules was characterized by the occupation of a combination of different glycosidic linkage and side-chain conformational positions whose relative occupations were sensitive to dielectric screening. Molecular dynamics simulations of the trisaccharide 3 showed little difference between the linkage conformations in the trisaccharide and the component disaccharides 1 and 2. Experimental optical rotation data of 1 and 2 were obtained as a function of temperature in varying solvents. The molecular models were combined with the semiempirical theory of Stevens and Sathyanarayana to yield calculated optical rotations. Interpretation of the data of both 1 and 2 implied that a combination of conformations, both in glycosidic and side-chain positions, could explain the experimental data. Solvents effects were important in influencing the conformational mix and averaged optical rotation. Three-bond heteronuclear coupling constants 3JC, H were obtained for the glycosidic linkages of 1 and 2 in D2O and DMSO. Analysis of the coupling constants with a Karplus curve showed that small reductions in the glycosidic torsion angles of the conformations of the models used here of ca. 10°–15° in ϕ and 5°–10° in ψ were required to give better agreement with experiment; a combination of conformations for both 1 and 2 was consistent with the data. There was a negligible influence on the coupling constants of 1 on changing the solvent from D2O to DMSO. © 1997 John Wiley & Sons, Inc.  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