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 11 毫秒
1.
Attempted cyclization of 2,3,4-tri-O-methyl-5-seleno-L-arabinose dimethyl acetal in acidic solution gave the corresponding diselenide. Intramolecular attack by the selenobenzyl group at C-5 of 5-O-p-tolylsulfonyl-L-arabinose dibenzyl diseleno-acetal resulted in the formation of benzyl 1,5-diseleno-L-arabinopyranoside. Similarly, 2,3,5-tri-O-methyl-4-O-p-tolylsulfonyl-D-xylose dibenzyl diselenoacetal gave benzyl 2,3,5-tri-O-methyl-1,4-diseleno-L-arabinofuranoside, and 2,3,4-tri-O-acetyl-5-O-p-tolylsulfonyl-D-xylose (or ribose) dibenzyl diselenoacetal gave benzyl 2,3,4-tri-O-acetyl-1,5-diseleno-D-xylo- (or ribo-)pyranoside. The glycosylic benzylseleno group was removed from the pyranoside with mercuric acetate, but attempted deacetylation of the product led to decomposition and not to the expected 5-seleno-D-xylopyranose.  相似文献   

2.
Laminarabiose, cellobiose, and gentiobiose were acetonated with 2,2-dimethoxy-propane under various conditions. Two isopropylidene acetals in which the reducing D-glucose residue had the furanoid form were obtained from laminarabiose, and two, in which the reducing D-glucose residue formed the acyclic dimethyl acetal, from cellobiose. Gentiobiose gave both types of isopropylidene compound.  相似文献   

3.
4.
The rates of formation and dissociation of concanavalin A with some 4-methylumbelliferyl and p-nitrophenyl derivatives of α- and β-D-mannopyranosides and glucopyranosides were measured by fluorescence and spectral stopped-flow methods. All process examined were uniphasic. The second-order formation rate constants varied only from 6.8 · 104 to 12.8 · 104 M?. s?1, whereas the first-order dissociation rate constants ranged from 4.1. to 220 s?1, all at ph 5.0, I = 0.3 M, and 25°C. Dissociation rates thus controlled the value of binding constant. The effect of temperature on these reactions was examined, from which enthalpies and entropies of activation and of reaction could be calculated. The effects of pH at 25°C on the reaction rates of 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside with concanavalin A were examined. The value of the binding constant Kap (derived from the kinetics) at any pH could be related to the intrinsic binding constant K by the expression Kap = KaK(Ka + [H+])?1. The values of Ka, the ionization constant of the protein segment responsive to sugar binding, were 3 · 10?4 M and 1 · 10?4 M for 4-methylumbelliferyl α-D-mannopyranoside and 4-methylumbelliferyl α-D-glucopyranoside, respectively. The binding constant of p-nitrophenyl α-D-mannopyranoside is surprisingly much less sensitive to a pH change from 5.0 to 2.7. Ionic strength had little effect on the binding characteristics of 4-methylumbelliferyl α-D-mannopyranoside to concanavalin A at pH 5.2 and 25°C.  相似文献   

5.
Allyl 4-O-(4-O-acetyl-2-O-benzoyl-3,6-di-O-benzyl-β-d-galactopyranosyl)-2-O-benzoyl-3,6-di-O-benzyl-α-d- galactopyranoside was O-deallylated to give the 1-hydroxy derivative, and this was converted into the corresponding 1-O-(N-phenylcarbamoyl) derivative, treatment of which with dry HCl produced the α-d-galactopyranosyl chloride. This was converted into the corresponding 2,2,2-trifluoroethanesulfonate, which was coupled to allyl 2-O-benzoyl-3,6-di-O-benzyl-α-d-galactopyranoside, to give crystalline allyl 4-O-[4-O-(4-O-acetyl-2-O-benzoyl-3,6-di-O-benzyl-β-d-galactopyranosyl)-2-O-benzoyl-3,6-di- O-benzyl-β-d-galactopyranosyl]-2-O-benzoyl-3,6-di-O-benzyl-α-d-galactopyranoside (15) in 85% yield, no trace of the α anomer being found. The trisaccharide derivative 15 was de-esterified with 2% KCN in 95% ethanol, and the product O-debenzylated with H2-Pd, to give the unprotected trisaccharide. Alternative sequences are discussed.  相似文献   

6.
The structure of d-threo-2,5-hexodiulosonic acid (1) and various derivatives in solution was determined by 13C-n.m.r. spectroscopy to be a hydrated, pyranose form. The structures of the methyl ester of 1 and of its 5-(dimethyl acetal) were confirmed by chemical means and by X-ray structure analysis.  相似文献   

7.
Two l-arabino-d-galactan-containing glycoproteins having a potent inhibitory activity against eel anti-H agglutinin were isolated from the hot saline extracts of mature radish leaves and characterized to have a similar monosaccharide composition that consists of l-arabinose, d-galactose, l-fucose, 4-O-methyl-d-glucuronic acid, and d-glucuronic acid residues. The chemical structure features of the carbohydrate components were investigated by carboxyl group reduction, methylation, periodate oxidation, partial acid hydrolysis, and digestion with exo- and endo-glycosidases, which indicated a backbone chain of (1→3)-linked β-d-galactosyl residues, to which side chains consisting of α-(1→6)-linked d-galactosyl residues were attached. The α-l-arabinofuranosyl residues were attached as single nonreducing groups and as O-2- or O-3-linked residues to O-3 of the β-d-galactosyl residues of the side chains. Single α-l-fucopyranosyl end groups were linked to O-2 of the l-arabinofuranosyl residues, and the 4-O-methyl-β-d-glucopyranosyluronic acid end groups were linked to d-galactosyl residues. The O-α-l-fucopyranosyl-(1→2)-α-l-arabinofuranosyl end-groups were shown to be responsible for the serological, H-like activity of the l-arabino-d-galactan glycoproteins. Reductive alkaline degradation of the glycoconjugates showed that a large proportion of the polysaccharide chains is conjugated with the polypeptide backbone through a 3-O-d-galactosylserine linkage.  相似文献   

8.
2,3-O-Isopropylidene-d-ribose diethyl dithioacetal, prepared from d-ribose, was converted in three steps into the corresponding dimethyl acetal, which was monotosylated at O-5, and the ester oxidized at C-4 with pyridinium chlorochromate; addition of methyl phenylphosphinate to the resulting pentos-4-ulose derivative then provided (4R,S)-4,5-anhydro-2,3-O-isopropylidene-4-C-[(R,S)-(methoxy)phenylphosphinyl]-d-erythro-pentose dimethyl acetal. Hydrogenation of this compound in the presence of Raney Ni, followed by reduction with SDMA, hydrolysis, and acetylation, yielded the title compounds (seven kinds), the structures of which were established on the basis of their 400-MHz, 1H-n.m.r. and mass spectra. A general dependence of the 2JPH and 3JPH values on the OPCH and PCCH dihedral angles provided an effective method for the assignment of the configurations and conformations of these 4-deoxy-4-phosphinyl-pentofuranoses.  相似文献   

9.
A variety of 1,3-diamino and 1,4-diaminocyclitols, manoaminocyclitols, and triaminocyclohexanol have been synthesized starting with the chiral ketone intermediate, 2, derived from l-quinic acid. Reduction of 2 with lithium borohydride afforded two epimeric diols (4 and 5), both of which were transformed by straight-forward but distinctly different chemical procedures into potentially useful aglycons for preparing novel tupes of bioactive, aminocyclitol glycoside antibiotics. The disposition of the substituents at C-1, C-3, C-4, and C-5 in 19 and 37 is identical with that present in the 2-deoxystreptamine nucleus in the naturally occurring antibiotics  相似文献   

10.
A partition chromatographic procedure utilizing a cationic exchange resin column in the Li+ form and 90% ethanol as the mobile phase was employed to quantify 3-deoxy-d-manno-octulosonic acid (KDO) and l-glycero-d-manno-heptose in the lipopolysaccharides (LPS) of Re and RdP? rough mutants of Salmonella minnesota. In a standard mixture of monosaccharides, KDO eluted shortly after the void volume and heptose eluted after the neutral hexoses. Mild acid treatment of either the Re or RdP? LPS with 0.16 n methanesulfonic acid in the presence of Dowex 50-X8 resin (H+ form) released more than 80% of the KDO residues within 15 min. The heptose of the RdP? LPS, first detected after 90 min of hydrolysis, increased gradually to a maximum level at 12 h. A secondary gradual increase in KDO became apparent during the heptose release. The weight contents of these two monosaccharides based upon aheir maximum values detected during hydrolysis were 20.3 ± 0.6% KDO, for the Re LPS, and 13.8 ± 0.4% KDO and 12.0 ± 0.4% heptose, for the RdP? LPS. The relationship between the kinetics of release of KDO and heptose and the nature of the linkages involving these two monosaccharides are discussed.  相似文献   

11.
A method has been studied for the determination of the position of the linkage of the 2-acetamido-2-deoxy-D-galactose and 2-acetamido-2-deoxy-D-glucose residues in oligosaccharides and glycoproteins that is based on the borohydride reduction of the reducing terminal residues to the corresponding alditol derivatives periodate oxidation, borohydride reduction, hydrolysis (eventually followed by borohydride reduction), separation of the fragments as per-O-(trimethylsilyl) or per-O-(trifluoroacetyl) derivatives, and identification of the fragments as derivatives of 2-acetamido-2-deoxyglycerol, 2-acetamido-2-deoxy-L-threitol, 2-acetamido-2-deoxy-L-arabinitol, 2-acetamido-2-deoxy-D-xylitol, 2-acetamido-2-deoxy-D-galactitol, and 2-acetamido-2-deoxy-D-glucitol by gas-liquid chromatography-mass spectrometry. New syntheses for the standard compounds 2-acetamido-2-deoxy-L-threitol and 2-acetamido-2-deoxy-D-xylitol are described.  相似文献   

12.
1,2,4,6-Tetra-O-acetyl-3-O-benzyl-α-D-mannopyranose (7) was obtained in good yield from 3,4,6-tri-O-benzyl-1,2-O-(1-methoxyethylidene)-β-D-mannopyranose (1) by acetolysis. Hydrogenolysis of 7 afforded 1,2,4,6-tetra-O-acetyl-α-D-mannopyranose which is a versatile intermediate for the preparation of other 3-O-substituted D-mannoses, such as 3-O-methyl-D-mannose and 3-O-α-D-mannopyranosyl-D-mannose. 3,4-Di-O-methyl-D-mannose was readily prepared from 1,2,6-tri-O-acetyl-3,4-di-O-benzyl-α-D-mannopyranose, which was also obtained from 1 by controlled acetolysis.  相似文献   

13.
A method is described for the synthesis of purine d-arabinonucleosides that uses purine bases and 2,2′-anhydro-(1-β-d-arabinofuranosylcytosine), AraC-an, as the starting materials. AraC-an was chosen as the precursor to the d-arabinosyl donor, because it is more readily available than any of the products that may be sequentially derived from it, namely, 1-β-d-arabinofuranosylcytosine (AraC), 1-β-d-arabinofuranosyluracil (AraU), and α-d-arabinofuranosyl-1-phosphate (Araf 1-P), a d-arabinofuranosyl donor. Four reactions were involved in the overall process; (a) AraC-an was nonenzymically hydrolyzed at alkaline pH to AraC which was then (b) deaminated by cytidine deaminase to AraU, a nucleoside, (c) phosphorylyzed by uridine phosphorylase to Araf 1-P, and (d) this ester caused to react with a purine base to afford a purine d-arabinonucleoside, the reaction being catalyzed by purine nucleoside phosphorylase. All four reactions occurred in situ, the first and second being performed sequentially, whereas the third and fourth were combined in a single step. The three enzyme catalysts were purified from Escherichia coli. The efficiency of the method is exemplified by the synthesis of the d-arabinonucleosides of 2,6-diaminopurine and adenine; the overall yields, based on AraC-an, were 60 and 80%, respectively.  相似文献   

14.
Capillary g.l.c. on SE-30 of the trimethylsilylated (-)-2-butyl glycosides of d and l monosaccharides gives multiple peak patterns, which can be used for the assignment of the absolute configurations. (-)-2-Butyl glycosides can be prepared from monosaccharides or their methyl glycosides; consequently, for the analysis of oligo- or poly-saccharides, hydrolysis as well as methanolysis can be applied. Provided that the peaks of the (-)-2-butyl glycosides do not completely overlap, mixtures of monosaccharides can be analysed directly, as illustrated for the constituents of the cell-wall lipopolysaccharide from Salmonella typhimurium LT-2.  相似文献   

15.
The kinetics of D-xylose transport were studied in Rhodotorula glutinis. Analysis of the saturation isotherm revealed the presence of at least two carriers for d-xylose in the Rhodotorula plasma membrane. These two carriers exhibited Km values differing by more than an order of magnitude. The low Km carrier was repressed in rapidly growing cells and depressed by starvation of the cells.Several hexoses were observed to inhibit d-xylose transport. In the studies reported here, the inhibitions produced by d-galactose and 2-deoxy-d-glucose were examined in some detail in order to define the interactions of these sugars with the d-xylose carriers. 2-Deoxy-d-glucose competitively inhibited both of the d-xylose carriers. In contrast, only the low-Km carrier was competitively inhibited by d-galactose.  相似文献   

16.
17.
Extraction of germinating-lupin cotyledons, followed by ion-exchange and gel chromatography, gave two α-l-arabinofuranosidases and three β-d-galactopyranosidases. Some fractions were further purified by using Concanavalin A-Sepharose. The changes in the activities of the enzymes during germination have been determined. Some kinetic and physical properties of these enzymes are described, and their role in the modification of cell-wall polysaccharides is discussed.  相似文献   

18.
Fusion or β-l-fucopyranose tetraacetate with phosphoric acid for 1 min at 50° gives a 9:1 anomeric mixture of the α-and β-pyranosyl phosphates. Longer fusion times give the α-anomer exclusively. The l-fucofuranose tetraacetates were synthesized for the first time by acetolysis or methyl-2,3,5-tri-O-acetyl-β-l-fucofuranoside. Fusion or the furanose tetraacetates with phosphoric acid gave a mixture or the fucofuranosyl phosphates in which the β-anomer predominated (β/α = 2.4). Anomeric pairs in the fucofuranose series appear to be distinguishable by the chemical shift of the C-6 methyl protons, as already shown by Sinclair and Sleeter in the pyranose series.  相似文献   

19.
A series of aroyl- and aryl-hydrazide derivatives was prepared from d-glycero-d-gulo-heptono-1,4-lactone (1). The reactivity of the NH proton in these hydrazides, in terms of their dissociation constants (pKa), was determined from their electronic spectra, and correlated to the Hammett σ values of the substituents. Comparable reactivities of the NH protons for the compounds, and the effect of the substituent, were studied by n.m.r. spectroscopy. Decomposition of the aroylhydrazides with copper(II) sulfate or nitrous acid resulted in the regeneration of 1.  相似文献   

20.
The transport of d-glucose by brush border membranes isolated from the rabbit renal cortex was studied. At concentrations less than 2 mM, the rate of d-glucose uptake increased linearly with the concentration of the sugar. No evidence was found for a “high-affinity” (μM) saturable site. Saturation was indicated at concentrations of d-glucose greater than 5 mM. The uptake of d-glucose was stereospecific and selectively inhibited by d-galactose and other sugars. Phlorizin inhibited the uptake of d-glucose in the presence and absence of Na+. The glycoside was a potent inhibitor of the efflux of d-glucose. Preloading the brush border membrane vesicles with d-glucose, but not with l-glucose, accelerated exchange diffusion of d-glucose. These results demonstrate that the uptake of d-glucose by renal brush borders represents transport into an intravesicular space rather than solely binding. The rate of d-glucose uptake was increased when the Na+ in the extravesicular medium was high and the membranes were preloaded with a Na+-free medium. The rate of d-glucose uptake was inhibited by preloading the brush border membranes with Na+. These results are consistent with the Na+ gradient hypothesis for d-glucose transport in the kidney. Thus, the presence of a Na+-dependent facilitated transport of d-glucose in isolated renal brush border membranes is indicated. This finding is consistent with what is known of the transport of the sugar in more physiologically intact preparations and suggests that the membranes serve as an effective model system in examining the mechanism of d-glucose transport in the kidney.  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