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.
De-etherification of 6,6′-di-O-tritylsucrose hexa-acetate (2) with boiling, aqueous acetic acid caused 4→6 acetyl migration and gave a syrupy hexa-acetate 14, characterised as the 4,6′-dimethanesulphonate 15. Reaction of 2,3,3′4′,6-penta-O-acetylsucrose (5) with trityl chloride in pyridine gave a mixture containing the 1′,6′-diether 6 the 6′-ether 9, confirming the lower reactivity of HO-1′ to tritylation. Subsequent mesylation, detritylation, acetylation afforded the corresponding 4-methanesulphonate 8 1′,4-dimethanesulphonate 11. Reaction of these sulphonates with benzoate, azide, bromide, and chloride anions afforded derivatives of β-D-fructofuranosyl α-D-galactopyranoside (29) by inversion of configuration at C-4. Treatment of the 4,6′-diol 14 the 1,′4,6′-triol 5, the 4-hydroxy 1′,6′-diether 6 with sulphuryl chloride effected replacement of the free hydroxyl groups and gave the corresponding, crystalline chlorodeoxy derivatives. The same 4-chloro-4-deoxy derivative was isolated when the 4-hydroxy-1′,6′-diether 6 was treated with mesyl chloride in N,N-dimethylformamide.  相似文献   

2.
The influence substituents on the hydrolysis of substituted phenyl α-D-mannopyranosides by α-D-mannosidase from Medicago sativa L. has been investigated. As indicated by structure-activity relations, the electronic effect of the substituent has an influence on the rate of formation of the intermediate mannosyl-enzyme complex. This effect depends not only on the nature of the substituent, but also on its position (meta or para) and on the temperature of the experiment. Hammett-type linear free energy relationships show that the reaction constant p changes its sign at ~27°. Substrates with strong electron-withdrawing groups show values of log V that are linearly related to 1/T, whereas the Arrhenius plots for other substrates are severely curved. This complex behaviour is tentatively explained by assuming that some meta-substituents have an unusual, temperature- and substituent-dependent influence on the formation of the Michaelis—Menten complex.  相似文献   

3.
Condensation of dimeric 3,4,6-tri-O-acetyl-2-deoxy-2-nitroso-α-D-glucopyranosyl chloride (1) with 1,2-O-isopropylidene-α-D-glucofuranurono-6,3-lactone (2) gave 1,2-O-isopropylidene-5-O-(3,4,6-tri-O-acetyl-2-deoxy-2-hydroxyimino-α-D-arabino-hexopyranosyl)-α-D-glucofuranurono-6,3-lactone (3). Benzoylation of the hydroxyimino group with benzoyl cyanide in acetonitrile gave 1,2-O-isopropylidene-5-O-(3,4,6-tri-O-acetyl-2-benzoyloxyimino-2-deoxy-α-D-arabino-hexopyranosyl)-α-D-glucofuranurono-6,3-lactone (4). Compound 4 was reduced with borane in tetrahydrofuran, yielding 5-O-(2-amino-2-deoxy-α-D-glucopyranosyl)-1,2-O-isopropylidene-α-D-glucofuranose (5), which was isolated as the crystalline N-acetyl derivative (6). After removal of the isopropylidene acetal, the pure, crystalline title compound (10) was obtained.  相似文献   

4.
Conversion of benzyl αβ-D-galactofuranoside into the 5,6-O-[α-(dimethyl-amino)benzylidene] derivative, followed by acetylation of HO-2 and HO-3, and selective ring opening or the acetal, gave benzyl 2,3-di-O-acetyl-6-O-benzoyl-αβ-D-galactofuranoside(4). The title disaccharide was synthesised from4 by reaction with 3,4,6-tri-O-acetyl-α-D-galactofuranose 1,2-(methyl orthoacetate) followed by removal of protecting groups  相似文献   

5.
Reaction of 2,3-di-O-acetyl-1,6-anhydro-β-D-galactopyranose (2) with 2,3,4,6-tetra- O-acetyl-α-D-galactopyranosyl bromide in the presence of mercuric cyanide and subsequent acetolysis gave 1,2,3,6-tetra-O-acetyl-4-O-(2,3,4,6-tetra-O-acetyl-α-D-galactopyranosyl)-α-D-galactopyranose (4, 40%) and 1,2,3,6-tetra-O-acetyl-4-O-(2,3,4,6-tetra-O-acetyl-β-D-galactopyranosyl)-α-D-galactopyranose (5, 30%). Similarly, reaction of 2,4-di-O-acetyl-1,6-anhydro-β-D-galactopyranose (3) gave 1,2,4,6-tetra-O-acetyl-3-O-(2,3,4,6-tetra-O-acetyl-α-D-galactopyranosyl)-α-D-galactopyranose (6, 46%) and 1,2,4,6-tetra-O-acetyl-3-O-(2,3,4,6-tetra-O-acetyl-β-D-galactopyranosyl)-α-D-galactopyranose (7, 14%). The anomeric configurations of 4-7 were assigned by n.m.r. spectroscopy. Deacetylation of 4-7 afforded 4-O-α-D-galactopyranosyl-D-galactose (8), 4-O-β-D-galactopyranosyl-D-galactose (9), 3-O-α-D-galactopyranosyl-D-galactose (10), and 3-O-β-D-galactopyranosyl-D-galactose (11), respectively.  相似文献   

6.
Quaternary ammonium and triphenylphosphonium salts of 2,3,4-tri-O-benzyl-6-O-(N-phenylcarbamoyl)-D-glucopyranosyl bromide were readily prepared by reaction with tertiary amines and triphenylphosphine under anhydrous conditions. Methanolysis of these salts was studied to determine the conditions of solvent and temperature that would produce the highest yields of α-D-glucosides. The quaternary ammonium salts gave the highest yields with solvents of low dielectric constant and room temperature. The phosphonium salts gave moderate yields with diethyl ether at 50°. The synthesis of methyl 2,3,4-tri-O-benzyl-6-O-(2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl)-α-D-glucopyranoside by treatment of the quaternary ammonium salt of 2,3,4,6-tetra-O-benzyl-α-D-glucopyranosyl bromide with methyl 2,3,4-tri-O-benzyl-α-D-glucopyranoside was studied as a model for the synthesis of oligosaccharides. The anomeric composition of the disaccharide product could be easily determined from the optical rotation since the specific rotations of both the final product and of the gentiobioside analog are known. Under the best conditions, the yield of disaccharide was low (50%) and the reactions were not completely stereoselective.  相似文献   

7.
The title disaccharide (16) has been synthesized in 50% overall yield by way of condensation of 4,6-di-O-acetyl-2,3-O-carbonyl-α-D-mannopyranosyl bromide 5 with methyl 2,3-O-isopropylidene-α-L-rhamnopyranoside (1) in chloroform solution, in the presence of silver oxide. The disaccharide was characterized as the crystalline isopropyl alcoholate of methyl 4-O-β-D-mannopyranosyl-α-L-rhamnopyranoside (11) and as 1,2,3-tri-O acetyl-4-O- (2,3,4,6-tetra-O-acetyl-β-D-mannopyranosyl)-α-L-rhamnopyranose (15). Methyl β-D-mannopyranoside isopropyl alcoholate 7 was readily obtained in 85% yield via the reaction of bromide 5 with methanol.Reduction of 2,3-di-O-methyl-L-rhamnose with sodium borohydride, followed by acetylation, may result in the formation of an appreciable proportion of a boric ester, namely 1,5-di-O-acetyl-4-deoxy-2,3-di-O-methyl-L-rhamnitol-4-yl dimethyl borate, depending on the procedure used.  相似文献   

8.
Methyl α-d-mannopyranoside (1 mole) reacts with 2,2-dimethoxypropane (1 mole), to give the 4,6-O-isopropylidene derivative (2) which rearranges to the 2,3-O-isopropylidene derivative (4). Compound4 can also be prepared by graded hydrolysis of methyl 2,3:4,6-di-O-isopropylidene-α-d-mannopyranoside. Successive benzoylation, oxidation, and reduction of4 provides a useful route to a number ofd-talopyranoside compounds. Methyl α-d-mannofuranoside (1 mole) reacts with 1–2 moles of 2,2-dimethoxypropane to give the 5,6-O-isopropylidene derivative (16) in 90% yield.  相似文献   

9.
The crystal structure of α-D-Manp-(1→3)-β-D-Manp-(1→4)-α-D-GlcNAcp has been determined by the direct method using the multi-solution, tangent formula, and “magic integer” procedures. The space group is P22, and 2 molecules are in the unit cell with a  9.894 (5), b  10.372 (6), c  11.816 (6) Å, and β  95.03° (6). The structure was refined to R 0.059 for 2099 reflections measured with Mo Kα radiation. Difference synthesis showed all the hydrogen atoms, and indicated a partial (~30%) substitution of the α-anomer molecules by the β-anomer molecules. The D-mannopyranose and the D-glucopyranose have the normal 4C1 conformation; an intramolecular hydrogen-bond O-3″-H.....O-5′ (2.703 Å) stabilises the GlcNAc in relation to β-D-mannopyranose.  相似文献   

10.
Rate coefficients and activation parameters were determined for the hydrochloric acid-catalysed hydrolysis of substituted phenyl α-D-galactopyranosides. Application of the Hammett—Zucker and the Bunnett criteria leads to contradictory conclusions about the mechanism. Substituents have only a small influence on the reaction. Under comparable conditions, the phenyl α-D-galactopyranosides hydrolyse faster than the corresponding β anomers. Most probably, these α anomers hydrolyse via the cyclic mechanism with protonation of the exocyclic oxygen atom.  相似文献   

11.
3-Azido-2,4,6-tri-O-benzyl-3-deoxy-α-D-glucopyranosyl chloride (7), prepared conventionally from the azido precursor 2, was coupled with “diisopropylidene-D-pinitol” (8) to give the α-D-glucoside 9 in good yield, together with some β anomer. Removal of the O-benzyl groups from 9 and reduction of the azido group to ?NH2 were accomplished simultaneously. Further deprotection yielded 11, a 3-amino-3-deoxy-α-D-glucoside of D-pinitol (1a). Compound 11 was converted into the (impure) 3-acetamidino hydrochloride 12. The synthesis of 3,6-epimino-D-glucosides was accomplished by ring closure of the 3-N-tosyl-6-O-tosyl intermediates 17 and 13. The products, after deprotection, were methyl 3,6-dideoxy-3,6-epimino-β-D-glucopyranaside (20) and the novel 3,6-epimino analog 15 of the pinitol D-glucoside 11.  相似文献   

12.
Methods for the synthesis of 3-O-(α-d-mannopyranosyl)-d-mannose and 2-(4-aminophenyl)ethyl 3-O-(α-d-mannopyranosyl)-α-d-mannopyranoside have been investigated by a number of sequences. Glycosidations with 2,3-di-O-acetyl-4,6-di-O-benzyl-d-mannopyranosyl and 2-O-benzoyl-3,4,6-tri-O-benzyl-d-mannopyranosyl p-toluenesulfonates were found to give better yields than the Helferich modification, the use of a peracylated d-mannopyranosyl halide, or the use of triflyl leaving group. Only the α anomer was obtained. Factors influencing glycosidation reactions are discussed. A mercury(II) complex was used for selective 2-O-acylation of 4,6-di-O-benzyl-α-d-mannopyranosides. A disaccharide—protein conjugate was prepared by the isothiocyanate method.  相似文献   

13.
Syntheses of 2-Se-(1,2,3,4-tetra-O-acetyl-β-D-glucopyranosyl)-3-N,N-dimethyl-selenopseudourea hydroiodide (3), 1,2,3,4-tetra-O-acetyl-6-S-dimethylarsino-6-thio-β-D-glucopyranose (4), 1,2,3,4-tetra-O-acetyl-6-Se-dimethylarsino-6-seleno-β-D-glucopyranose (7), 6-S-dimethylarsino-6-thio-β-D-glucopyranose (5), and 6-Se-dimethylarsino-6-seleno-β-D-glucopyranose (9) are described. Various spectral properties of the compounds are given. The relative rates of alkaline hydrolysis of 5 and 9 are compared.  相似文献   

14.
Methyl 4,6-O-benzylidene-3-deoxy-3-phenylazo-α-D-glucopyranoside (1) has been oxidised with the Pfitzner—Moffat reagent to the 2,3-diulose 3-phenylhydrazone derivative (2) which has been characterised as the phenylosazone (3) and oxime (4). An unstable 2-imino derivative (10) of the same diulose has been produced by base-catalysed elimination of nitrogen from methyl 2-azido-4,6-O-benzylidene-2-deoxy-α-D-ribo-hexopyranosid-3-ulose (8). The imino intermediate was trapped as a quinoxaline derivative (9). The base-catalysed reactions of certain other hydrazone derivatives of methyl hexosiduloses have also been examined.  相似文献   

15.
The koenigs-Knorr glycosylation of 4,6-O-ethylidene-1,2-O-isopropylidene-3-O-(2,3-O-isopropylidene-α-l-rhamnopyranosyl)-α-d-galactopyranose (3) by 4,6-di-O-acetyl-2,3-O-carbonyl-α-d-mannopyranosyl bromide (10), as well as Helferich glycosylations of 3 by tetra-O-acetyl-α-d-mannopyranosyl and -α-d-glucopyranosyl bromides, proceeded smoothly to give high yields of trisaccharide derivatives (12, 16, and 17). An efficient procedure for the transformation of 12, 16, and 17 into the α-deca-acetates of the respective trisaccharides has been developed. Zemplén de-acetylation then afforded the title trisaccharides in yields of 53, 52, and 62 %, respectively, from 3. A new route to 1,4,6-tri-O-acetyl-2,3-O-carbonyl-α-d-mannopyranose is suggested.  相似文献   

16.
A homogeneous glucan has been isolated from the fruiting bodies of Cyttaria harioti Fischer. Partial acid hydrolysis produced major amounts of isomaltose, whereas acetolysis gave maltose and maltotriose. Enzymic hydrolysis with amylo-glucosidase and pullulanase indicated a structure based on maltotriose residues connected by (1→6)-α-D linkages. This conclusion was supported by periodate-oxidation data which also showed that 3–7% of the glucose resisted oxidation. Methylation analysis confirmed the presence of (1→6) and (1→4) linkages in the ratio 1:2.4.  相似文献   

17.
Hydrogenation, severally, of methyl 3-azido-2,3,6-trideoxy-β-D-erythro-hex-5-enopyranoside, its 3-benzamido analogue, and methyl 2,6-dideoxy-β-D-threo-hex-5-enopyranoside in the presence of palladium-on-barium sulphate gave the corresponding 6-deoxy-β-D-hexopyranoside derivatives. Stereoselective addition of hydrogen was observed in each case. Methyl 2,6-dideoxy-β-D-arabino-hexopyranoside was also prepared by reductive dehalogenation of methyl 3,4-di-O-benzoyl-6-bromo-2,6-dideoxy-β-D-arabino-hexopyranoside.  相似文献   

18.
19.
1-O-Tosyl-d-glucopyranose derivatives having a nonparticipating benzyl group at O-2 have been shown to react rapidly in various solvents with low concentrations of alcohols, either methanol or methyl 2,3,4-tri-O-benzyl-α-d-glucopyranoside. The stereospecificity of the glucoside-forming reaction could be varied from 80% of β to 100% of α anomer by changing the solvent or modifying the substituents on the 1-O-tosyl-d-glucopyranose derivative. 2,3,4-Tri-O-benzyl-6-O-(N-phenylcarbamoyl)-1-O-tosyl-α-d-glucopyranose in diethyl ether gave a high yield of α-d-glucoside. Kinetic measurements of reaction with various alcohols (methanol, 2-propanol, and cyclohexanol) show a high rate even at low concentrations of alcohol, and give some insight into the reaction mechanism. The high rate and stereoselectivity of their reaction suggest that the 1-O-tosyl-d-glucopyranose derivatives may be used as reagents for oligosaccharide synthesis.  相似文献   

20.
The effect of phenols on the hydrolysis of substituted phenyl β-d-gluco- and β-d-xylo-pyranosides by β-d-glucosidase from Stachybotrys atra has been investigated. Depending on the glycon part of the substrate and on the phenol substituent, the hydrolysis is either inhibited or activated. With aryl β-d-xylopyranosides, transfer of the xylosyl residue to the phenol, with the formation of new phenyl β-d-xylopyranosides, is observed. With aryl β-d-glucopyranosides, such transfer does not occur when phenols are used as acceptors, but it does occur with anilines. A two-step mechanism, in which the first step is partially reversible, is proposed to explain these observations. A qualitative analysis of the various factors determining the overall effect of the phenol is given.  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