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 9 毫秒
1.
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   

2.
3.
Cell wall specimens from nonendospermic tissues of six grasses and representatives of nine other monocot families were treated with a specific glucanase in order to liberate wall-bound noncellulosic β-d-glucans. Gel filtration chromatography profiles of the oligosaccharides released from all grass species indicated the presence of a mixed linkage β-(1→3):(1→4)-glucan. The results also indicate that this glucan was not present in the other monocots examined. The evidence from this and previous studies indicates that the mixed linkage glucan may be restricted in the monocots to the Gramineae.  相似文献   

4.
Reaction of 1,2-O-cyclopentylidene-α-d-glucofuranurono-6,3-lactone (2) with 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl bromide (1) gave 1,2-O-cyclopentylidene- 5-O-(2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl)-α-d-glucofuranurono-6,3-lactone (3, 45%) and 1,2-O-cyclopentylidene-5-O-(2,3,4,6-tetra-O-acetyl-β-d-glucopyranosyl)-α-d-glucofuranurono-6,3-lactone (4, 38%). Reduction of 3 and 4 with lithium aluminium hydride, followed by removal of the cyclopentylidene group, afforded 5-O-α-(9) and -β-d-glucopyranosyl-d-glucofuranose (12), respectively. Base-catalysed isomerization of 9 yielded crystalline 5-O-α-d-glucopyranosyl-d-fructopyranose (leucrose, 53%).  相似文献   

5.
The rate constants for the hydrolysis of six alkyl and four aryl β-d-xylofuranosides in aqueous perchloric acid at various temperatures have been measured. The effects of varying the aglycon structure on the hydrolysis rate are interpreted in terms of two concurrent reactions. Either, the substrate, protonated on the glycosidic oxygen atom, undergoes a rate-limiting heterolysis to form a cyclic oxocarbonium ion, or, an initial rapid protonation of the ring oxygen is followed by a unimolecular cleavage of the five-membered ring, all subsequent reactions being fast. It is suggested that xylofuranosides having strongly electron-attracting aglycon groups react mainly by the former pathway, whereas the latter is more favourable for substrates having electron-repelling aglycon groups. The negative entropies of activation obtained with the latter compounds are attributed to the rate-limiting opening of the five-membered ring. The rate variations of the hydrolyses of alkyl β-d-xylofuranosides in aqueous perchloric acid-methyl sulfoxide mixtures are interpreted as lending further support for the suggested chance in mechanism.  相似文献   

6.
The crystal structures of α-d-glucopyranosyl β-d-psicofuranoside and α-d-galactopyranosyl β-d-psicofuranoside were determined by a single-crystal X-ray diffraction analysis, refined to R1 = 0.0307 and 0.0438, respectively. Both disaccharides have a similar molecular structure, in which psicofuranose rings adopt an intermediate form between 4E and 4T3. Unique molecular packing of the disaccharides was found in crystals, with the molecules forming a layered structure stacked along the y-axis.  相似文献   

7.
Hydrolysis of purin-6-yl 2-deoxy-1-thio-β-d-arabino-hexopyranoside (2) to 6-mercaptopurine and 2-deoxy-d-glucose is catalyzed by hydronium ion and almond β-d-glucosidase. The dependence of rate on acidity in water and deuterium oxide indicates that 2 and its conjugate acid undergo hydrolysis via a mechanism that involves a partially rate-limiting proton transfer. Although 2 is ≈103 more reactive than 6-purinyl β-d-glucothiopyranoside (1) in dilute aqueous acid, 1 is a better substrate for almond β-d-glucosidase.  相似文献   

8.
Structural features of noncellulosic β-d-glucans of Zea mays, Hordeum vulgare, Triticum vulgare, Secale cereale, and Sorghum bicolor were compared. Treatment of cell walls derived from these species with specific Bacillus subtilis or Rhizopus glucanases yields virtually identical profiles upon Bio-Gel P-2 fractionation of the liberated oligosaccharides. The two predominant reaction products, a trisaccharide and tetrasaccharide, were identified as 3-O-β-cellobiosyl-d-glucose and 3-O-β-cellotriosyl-d-glucose respectively by virtue of the specificity of these enzymes and by paper chromatography and electrophoresis. The similarity of the reaction product profiles indicates a rather regular repeating sequence in all β-d-glucans examined. The ratios of 3-O-β-cellobiosyl-d-glucose to 3-O-β-cellotriosyl-d-glucose indicates that 30.4–30.9% of the β-glucosyl linkages in the intact molecule are 1 → 3. The yields of wall glucan as estimated from the quantity of oligosaccharides released, range from 41 μg/mg wall in Hordeum to 97 μg/mg wall from Sorghum.  相似文献   

9.
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   

10.
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   

11.
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   

12.
2-Deoxy-β-d-lyxo-hexose (2-deoxy-β-d-galactose, C6H12O5), Mr = 164.16, is monoclinic, P21 with a = 9.811(1), b = 6.953(1), c = 5.315(1) Å, β = 91.58(2)°, V = 362.5(1) Å3, Z = 2, and Dx = 1.504 g.cm?3. The structure was solved by direct methods (MULTAN 79) and refined to R = 0.032 for 800 observed reflections. Each hydroxyl oxygen, acting both as donor and acceptor, is involved in a hydrogen-bonding system, which consists of infinite helical chains around the crystallographic screw axes. Moreover, weak interactions allow the incorporation of the ring-oxygen atoms into an interconnected network.  相似文献   

13.
The intestinal transport of three actively transported sugars has been studied in order to determine mechanistic features that, (a) can be attributed to stereospecific affinity and (b) are common.The apparent affinity constants at the brush-border indicate that sugars are selected in the order, β-methyl glucose >d-galactose > 3-O-methyl glucose, (the Km values are 1.23, 5.0 and 18.1 mM, respectively.) At low substrate concentrations the Kt values for Na+ activation of sugar entry across the brush-border are: 27.25, and 140 mequiv. for β-methyl glucose, galactose and 3-O-methyl glucose, respectively. These kinetic parameters suggest that Na+, water, sugar and membrane-binding groups are all factors which determine selective affinity.In spite of these differences in operational affinity, all three sugars show a reciprocal change in brush-border entry and exit permeability as Ringer [Na] or [sugar] is increased. Estimates of the changes in convective velocity and in the diffusive velocity when the sugar concentration in the Ringer is raised reveal that with all three sugars, the fractional reduction in convective velocity is approximately equal to the (reduction of diffusive velocity)2. This is consistent with the view that the sugars move via pores in the brush-border by convective diffusion.Theophylline reduces the serosal border permeability to β-methyl glucose and to 3-O-methyl glucose relatively by the same extent and consequently, increases the intracellular accumulation of these sugars.The permeability of the serosal border to β-methyl glucose entry is lower than permeability of the serosal border to β-methyl glucose exit, which suggests that β-methyl glucose may be convected out of the cell across the lateral serosal border.  相似文献   

14.
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   

15.
The formation of 1,6-anhydro-β-d-glucopyranose and several d-glucosyl oligosaccharides has been observed during the action of a purified, fungal glucosyltransferase (EC 2.4.1.24) on maltose. Such products are synthesized by a transglucosylation mechanism involving the formation of a d-glucosyl-enzyme complex and the displacement of the d-glucosyl group by appropriate acceptor-substrates. The formation of the 1,6-anhydro bond is a novel type of transfer reaction and occurs by displacement of the enzyme from the d-glucosyl-enzyme complex by the proton of the primary hydroxyl group of the same glucosyl group. This reaction is characterized by inversion of configuration at the position of glucosidic bond-cleavage of the substrate. Synthesis of the d-glucosyl oligosaccharides occurs by displacement of the d-glucosyl groups from the enzyme by suitable acceptor-substrates. In these cases, the reactions are characterized by retention of configuration of the d-glucosidic bonds of the substrate. The list of oligosaccharides produced from maltose includes nigerose, kojibiose, isomaltose, maltotriose, panose, isomaltotriose, and 6-O-d-glucosyl-panose. The identity of these compounds has been established by methylation analysis and enzymic hydrolysis. d-Glucose is also a product of the reaction and arises from both the reducing and the non-reducing groups of maltose.  相似文献   

16.
3- O-(2-Acetamido-2-deoxy-β-d-glucopyranosyl)-α-d-galactopyranose (10, “Lacto-N-biose II”) was synthesized by treatment of benzyl 6-O-allyl-2,4-di-O-benzyl-β-d-galactopyranoside with 2-methyl-(3,4,6-tri-O-acetyl-1,2-dideoxy-α-d-glucopyrano)[2,1-d]-2-oxazoline (5), followed by selective O-deallylation, O-deacetylation, and catalytic hydrogenolysis. Condensation of 5 with benzyl 6-O-allyl-2-O-benzyl-α-d-galactopyranoside, followed by removal of the protecting groups, gave 10 and a new, branched trisaccharide, 3,4-di-O-(2-acetamido-2-deoxy-β-d-glucopyranosyl)-d-galactopyranose (27).  相似文献   

17.
The structure of neoschaftoside is shown for the first time to be 6-C-β-d-glucopyranosyl-8-C-β-l-arabinopyranosylapigenin. A variety of chemical and spectroscopic techniques are involved.  相似文献   

18.
2,6-Anhydro-3-deoxy-aldehydo-d-lyxo-hept-2-enose (7) 2,6-anhydro-3-deoxy-d-lyxo-hept-2-enitol (8) were synthesized as half-chair analogs of d-galactal (1). As 1 is a strong inhibitor of, as well as a substrate for, β-d-galactosidase from Escherichia coli, the same properties were expected for 7 and 8; however, both were ineffective. This result, together with those of other authors, allows speculative conclusions on the tight binding of 1 to the enzyme only, when water or an alcohol is bound as a co-substrate.  相似文献   

19.
Enzymic analyses of aqueous extracts of barley obtained at 25–100° demonstrated an exponential relationship between β-d-glucan content and temperature. Purified β-d-glucans, in particular those from extracts made at 40° and 100°, were compared by the Smith-degradation method followed by g.l.c. analysis of the methyl and the trimethylsilyl ethers of the products. All extracts yielded qualitatively the same products, including oligosaccharides which were characteristic of sequences of adjacent (1→3)-linkages. The material extracted at 100° contained relatively more of these sequences, and also showed a much higher specific viscosity, than did the material extracted at low temperatures. The structural implications of these findings are discussed.  相似文献   

20.
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   

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