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.
The sodium salts of the 6-sulfate 7, the 4-sulfate 10, and the 4,6-disulfate 12 of benzyl 3-O-(beta-D-glucopyranosyl uronate)-beta-D-galactopyranoside (5) have been synthesized. Methyl (2,3,4-tri-O-acetyl-1-bromo-1-deoxy-alpha-d-glucopyran)uronate (1) was coupled with benzyl 2-O-benzoyl-4,6-O-benzylidene-beta-D-galactopyranoside (2) to yield 3. The benzylidene acetal of 3 was hydrolyzed to give benzyl 2-O-benzoyl-3-O-[methyl (2,3,4-tri-O-acetyl-beta-D-glucopyranosyl)uronate]-beta-D-galactopyra noside (4). Compound 4 was utilized as a key intermediate to prepare the sulfated disaccharides 7,10, and 12. Direct sulfation of 4 with sulfur trioxide-trimethylamine for 2 days yielded the 6-sulfate 6. The 4,6-disulfate 11 was accessible by running the reaction under the same conditions for 14 days. The 4-sulfate 9 was obtained after protecting the 6-OH group of 4 by reaction with benzoyl imidazole to give the 6-benzoate 8, followed by sulfation under vigorous conditions. Treatment of the protected compounds 4, 6, 9, and 11 with aqueous sodium hydroxide in tetrahydrofuran gave the unprotected 5, 7, 10, and 12, respectively.  相似文献   

2.
New analogues of acetal-type phospholipids were obtained on the basis of ethriol (2-hydroxymethy1-2-ethyl-1,3-propanediol). The starting triol originally was condensed with decanal or dodecanal to form acetals, which were then phosphorylated with tetraethyldiamidophosphorous acid chloride. The amidophosphites were further oxidized with iodosobenzene or sulfurized to the corresponding acetal-type phospholipids and their thio analogues.  相似文献   

3.
The 4,6-O-(1-methoxycarbonylethylidene), -(hydroxyisopropylidene), and -(methoxyisopropylidene) acetals of methyl 2,3-di-O-methyl-alpha-D-glucopyranoside were subjected to reductive cleavage in the presence of triethylsilane and trimethylsilyl methanesulfonate-boron trifluoride etherate (Me3SiOMs-BF3.Et2O), BF3.Et2O, or trimethylsilyl trifluoromethanesulfonate (Me3SiOSO2CF3) and the mole fractions of products were determined as a function of reaction time. The 4,6-(1-methoxycarbonylethylidene) acetal was quite stable to reductive-cleavage conditions but isomerization of the initial R,S mixture of diastereomers to the more-stable S diastereoisomer was noted. In addition, a slow, regiospecific, reductive ring-opening of the acetal was observed to give 6-O-[1-(methoxycarbonyl)ethyl] derivatives. The 4,6-(hydroxyisopropylidene) acetal was very unstable under reductive-cleavage conditions. Both Me3SiOMs-BF3.Et2O and Me3SiOSO2CF3 catalyzed complete removal of the group, via the intermediate 6-[1-(hydroxymethyl)ethyl] ether, but BF3.Et2O gave a mixture of products. The 4,6-(methoxyisopropylidene) acetal was also very labile under reductive-cleavage conditions; Me3SiOMs-BF3.Et2O catalyzed complete removal of the acetal, via the intermediate 6-[1-(methoxymethyl)ethyl]ether, but the intermediate ether was quite stable in the presence of either BF3.Et2O or Me3SiOSO2CF3. It is concluded from these studies that polysaccharides bearing 4,6-O-(1-carboxyethylidene) substituents can be analyzed directly by sequential permethylation and reductive cleavage. It is proposed that the identity of the substituted monomer and the positions of substitution of the acetal can be determined by sequential permethylation, ester reduction, and reductive cleavage.  相似文献   

4.
The acidic reagent formed in situ from anhydrous AlCl(3) and H(2)O in 3:1 ratio is much more efficient for the reductive openings of the cyclic benzylidene acetals with Me(3)N x BH(3) in tetrahydrofurane than the AlCl(3) alone. Under proposed conditions, the dioxane-type 4,6-O-bezylidene acetals of hexopyranosides give regioselectively the corresponding 4-hydroxy,6-O-benzyl derivatives in excellent yields. Reductive openings of the dioxolane-type 3,4-O-benzylidene acetals of galactopyranoside are also very efficient and regioselective and give either 3-O-benzyl derivative (from 3,4-O-exo-benzylidene acetal) or 4-O-benzyl derivative (from 3,4-O-endo-benzylidene acetal) depending on the configuration of the acetal carbon atom.  相似文献   

5.
The facile one step synthesis of 4-hydroxy-2(E)-nonenal and its dimethyl acetal via a cross-metathesis reaction between commercially available octen-3-ol and acrolein or its dimethyl acetal is reported. The method was extended to the synthesis of C6 and C12 4-hydroxy-2(E)-enals, their dimethyl acetal and of the 4-hydroxy-2(E)-nonenoic acid (4-HNA).  相似文献   

6.
Twelve bacterial polysaccharides of known structure containing a representative range of pyruvated monosaccharides, were methanolysed, trimethylsilylated, and analysed by g.l.c. and g.l.c.-m.s. Except for 3,4-O-(1-carboxyethylidene)-L-rhamnose, which was unusually labile, the pyruvic acid substituents were largely retained during methanolysis and the Me3Si derivatives of the resulting pyruvated methyl glycosides gave distinctive g.l.c. peaks with characteristic mass spectra. The pyranose rings of 4,6-O-(1-carboxyethylidene)-D-glucose, 4,6-O-(1-carboxyethylidene)-D-mannose, 4,6-O-(1-carboxyethylidene)-D-galactose, and 3,4-O-(1-carboxyethylidene)-D-galactose survived the methanolysis, but that of 2,3-O-(1-carboxyethylidene)-D-glucuronic acid was cleaved to give the methyl ester of 2,3-O-(1-carboxyethylidene)-aldehydo-D-glucuronic acid dimethyl acetal. In the case of 2,3-O-(1-carboxyethylidene)-D-galactose, cleavage of the pyranose ring was less complete; under the conditions used in these experiments two-thirds of the pyranose rings were intact while one-third were cleaved to give the methyl ester of 2,3-O-(1-carboxyethylidene)-aldehydo-D-galactose dimethyl acetal. A very small amount of 3,4-O-(1-carboxyethylidene)-L-rhamnose from one polysaccharide retained its pyruvic acid substituent after gentle methanolysis to give the methyl ester of 3,4-O-(1-carboxyethylidene)-aldehydo-L-rhamnose dimethyl acetal. Susceptibility to cleavage of the pyranose ring during methanolysis appears to be a property of pyruvated monosaccharides with trans-fused 1,3-dioxolane rings.  相似文献   

7.
Thiodiglycolaldehyde (2,2′-thiobisacetaldehyde, 1a) reacted severally with methanol, ethanol, and 2-propanol to give mixtures variously of thiodiglycoaldehyde bis(dialkyl acetals) (3a,b), cis-2,6-dialkoxy-1,4-oxathianes (5b-d), and trans-2,6-dialkoxy-1,4-oxathianes (7a-c). Thiodiglycolaldehyde bis(di-isopropyl acetal) (3c) was not formed in the reaction of 1a and 2-propanol, but 3c was obtained after bromoacetaldehyde di-isopropyl acetal was treated with sodium sulfide. The stereoisomers corresponding to 2,6-dimethoxy-1,4-oxathiane (5b, 7a) were obtained from the acyclic dimethyl acetal 3a. The reaction between 1a and thiols in acid media have been studied. With ethanethiol, thiodiglycolaldehyde bis(diethyl dithioacetal) was the only product, but a mixture of thiodiglycolaldehyde bis(di-tert-butyl dithioacetal), cis-2,6-bis(tert-butylthio)-1,4-dithiane, and trans-2,6-bis(tert-butylthio)-1,4-dithiane was obtained from 2-methyl-2-propanethiol. On oxidation to sulfones of the stereoisomers of 2,6-dialkoxy-1,4-oxathiane and 2,6-bis(alkylthio)-1,4-dioxane with hydrogen peroxide, the configurations were retained, but the stereoisomers of 2,6-bis(tert-butylthio)-1,4-dithiane were transformed into the same oxidation product.  相似文献   

8.
A novel asymmetric oxy-Michael addition reaction was developed. In the presence of a catalytic amount of base, chiral ketones 1 and 2, derived from D-glucose and D-fructose, respectively, reacted with omega-hydroxy enones or enoates 3a-e, 17 and 21 to form the hemiacetal-derived alkoxide which underwent stereoselective intramolecular Michael addition to give cyclic acetals. Although the stereoselectivities in the formation of the five-membered acetal rings were modest, six-membered ring formation proceeded with high stereoselectivity and the utility of the reaction was demonstrated by a simple syntheses of natural products.  相似文献   

9.
Acetalization of a steroidal ketone possessing an isopropylidenedioxy moiety in the other part of the molecule by acetal exchange with 2-ethyl-2-methyl-l,3-dioxolane was found to give an acetal with a sec-butylidenedioxy moiety. By this additional exchange reaction, a new chiral center was generated in the sec-butylidenedioxy moiety, and hence an unnecessary stereochemical complexity was added. To circumvent the problem, 2,2-dimethyl-l,3-dioxolane was employed for the acetalization. By adopting this procedure, several intermediates in Mori’s brassinolide synthesis could be obtained in pure and crystalline states. The present improved synthesis furnished 30 g of brassinolide in a 7.0% overall yield from stigmasterol, in contrast to 3.0% by the original procedure.  相似文献   

10.
Aldose diethyl dithioacetals react with 1.2 equivalents of 1-ethoxycyclohexene or 2-methoxypropene in N,N-dimethylformamide at 0° with p-toluenesulfonic acid as catalyst to give the five-membered ring acetal attached to the two terminal oxygen atoms as the major product in every case. In most instances, a small proportion of the terminal, six-membered ring acetal was also obtained, and in a few cases, terminal seven-membered ring acetals were also isolated. Cyclohexylidenation at room temperature gave the same products, but isopropyl-idenation at room temperature resulted in certain cases in partial rearrangement. Cyclohexylidenation reactions gave smaller proportions of the minor six- and seven-membered ring products. Structures were established from 13C-n.m.r. and mass spectra. The 13C-n.m.r. spectra of model cyclohexylidene derivatives were found very similar to those of isopropylidene derivatives previously studied. Two new features useful for structure determination were noted when the spectra of the precursor diols were compared with those of both types of derived acetals; the chemical shift of C-2 of a 1,3-propanediol derivative was shifted upfield by 6–9 p.p.m. on acetalation and the shifts of the diol carbon atoms attached to oxygen were affected according to the type of acetal and ring-size formed. Similar observations were made for methylene acetals.  相似文献   

11.
Hunter syndrome (mucopolysaccharidosis-II) is caused by deficiency of the lysosomal enzyme iduronate-2-sulfatase. The assay of this sulfatase requires the use of α-l-iduronate glycosides containing a sulfate at the 2-position. We report a simple, three-step procedure for the introduction of sulfate at the 2-position starting with the methyl ester of α-l-iduronate glycosides. The procedure involves protection of the 2- and 4-hydroxyl groups of the iduronate moiety as the dibutyl stannylene acetal, selective sulfation with sulfur trioxide-trimethylamine, and deprotection of the methyl ester to afford the desired 2-sulfate in 61% overall yield.  相似文献   

12.
Symmetrical polyols can be converted into benzyl ethers with one free hydroxyl group in good yield by reaction of the monodibutylstannylene acetal with excess benzyl bromide in the presence of tetrabutylammonium bromide and diisopropylethylamine in xylene. The reaction pathway involves initial benzylation of the dibutylstannylene acetal to give benzyl and bromodibutylstannyl ethers; if a hydroxyl group remains unsubstituted, the latter ether ring closes and reacts further.  相似文献   

13.
A novel galactosylalkylglycerol modified with a long-chain cyclic acetal at the sugar moiety, 3-O-(4'6'-plasmalogalactosyl) 1-O-alkylglycerol, was isolated from equine brain. The presence of cyclic acetal linkage, its linked position, and the length of the acetal chain of the natural plasmalo lipid were determined by proton NMR spectroscopy and fast-atom bombardment;-mass spectrometry, as well as gas chromatography;-mass spectrometry and gas;-liquid chromatography. To identify the isomeric stereostructure of the natural product, the plasmalo derivative was chemically synthesized from 3-O-galactosyl 2-O-acyl 1-O-alkyl glyceride through acetalization after deacylation. As a result, the direction and position of the acetal chain of the natural plasmalo lipid were characterized as an "endo"-type 4',6'-O-acetal derivative linked to galactoside by comparison with the NMR data of the synthesized product. The chain lengths of alkyl and acetal groups were C(14) for the former and C(16) and C(18) for the latter, and those for the latter group were mostly similar to those of plasmalogalactosyl ceramide, which was previously isolated from equine brain.  相似文献   

14.
Glutamate release activates signaling pathways important for learning and memory, and over-stimulation of these pathways during seizures leads to aberrant synaptic plasticity associated with hyper-excitable, seizure-prone states. Seizures induce rapid accumulation of membrane lipid-derived fatty acids at the synapses which, evidence suggests, regulate maladaptive connectivity. Here we give an overview of the significance of the arachidonyl- and inositol-derived messengers, prostaglandins (PGs) and diacylglycerol (DAG), in experimental models of epilepsy. We use studies conducted in our own laboratory to highlight the pro-epileptogenic role of cyclooxygenase-2 (COX-2) and its products, the PGs, and we discuss the possible mechanisms by which PGs may regulate membrane excitability and synaptic transmission at the cellular level. We conclude with a discussion of AA-DAG signaling in synaptic plasticity and seizure susceptibility with an emphasis on recent studies in our laboratory involving DAG kinase ε (DGKε)-knockout mice.  相似文献   

15.
Wang X  Harada T  Iwai H  Oku A 《Chirality》2003,15(1):28-30
In the presence of 2,6-diisopropylphenol and t-BuOMe (1 equiv. each), asymmetric Mukaiyama-Michael reaction of a variety of benzalacetone derivatives 2 with silyl ketene acetal 3 is efficiently catalyzed by O-(2-naphthoyl)-N-tosyl-(L)-allo-threonine-derived B-phenyloxazaborolidinone 1 (10 mol%) to give the corresponding adducts 5 in 60-90% ee.  相似文献   

16.
Abstract

Dimethylformamide dimethyl acetal (DMF-DMA) reacts with nucleosides under mild conditions to give N-dimethylaminomethylene (N-DMAM) derivatives. Silylation provides the DMAM-O-trialkylsilyl mixed derivatives which have good chromatographic and mass spectral properties.  相似文献   

17.
Feasibility of elemental sulfur reduction by Desulfovibrio desulfuricans in anaerobic conditions in a stirred reactor was studied. Hydrogen was used as energy source, whereas the carbonated species were bicarbonate and yeast extract. Attention was paid to reactor engineering aspects, biofilm formation on the sulfur surface, hydrogen sulfide formation rate and kinetics limitations of the sulfur reduction. D. desulfuricans formed stable biofilms on the sulfur surface. It was found that active sulfur surface availability limits the reaction rate. The reaction rate was first order with respect to sulfur and hydrogen velocity had no effect in the reaction rate for the range 8.2 x 10(-2) to 4.1 x 10(-1) Nm(3) m(-2) min(-1). At a superficial gas velocity (u(G)) = 3.1 x 10(-2) Nm(3) m(-2) min(-1), H(2)S(g) production rate decreased due to a deficient H(2)S stripping. A maximum H(2)S(g) production rate of 2.1 g H(2)S L(-1) d(-1) was achieved during 5 days with an initial sulfur density of 4.7% (w/v).  相似文献   

18.
In the presence of phosphate ions, molybdic ions (Mo6+) were reduced enzymatically with elemental sulfur by washed intact cells of Thiobacillus ferrooxidans to give molybdenum blue. The whole-cell activity that reduced Mo6+ was totally due to cellular sulfur:ferric ion oxidoreductase (SFORase) (T. Sugio, W. Mizunashi, K. Inagaki, and T. Tano, J. Bacteriol. 169:4916-4922, 1987). The activity of M06+ reduction with elemental sulfur was competitively inhibited by Fe3+, Cu2+, and Co2+. The Michaelis constant of SFORase for Mo6+ was 7.6 mM, and the inhibition constants for Fe3+, Cu2+, and Co2+ were 0.084, 0.015, and 0.17 mM, respectively, suggesting that SFORase can reduce not only Fe3+ and Mo6+ but also Cu2+ and Co2+ with elemental sulfur.  相似文献   

19.
The transformations of 1,2,3,4-tetrahydrodibenzothiophene (THDBT) were investigated with pure cultures of hydrocarbon-degrading bacteria. Metabolites were extracted from cultures with dichloromethane (DCM) and analyzed by gas chromatography (GC) with flame photometric, mass, and Fourier transform infrared detectors. Three 1-methylnaphthalene (1-MN)-utilizing Pseudomonas strains oxidized the sulfur atom of THDBT to give the sulfoxide and sulfone. They also degraded the benzene ring to yield 3-hydroxy-2-formyl-4,5,6,7-tetrahydrobenzothiophene. A cell suspension of a cyclohexane-degrading bacterium oxidized the alicyclic ring to give a hydroxy-substituted THDBT and a ketone, and it oxidized the aromatic ring to give a phenol, but no ring cleavage products were detected. GC analyses with an atomic emission detector, using the sulfur-selective mode, were used to quantify the transformation products from THDBT and dibenzothiophene (DBT). The cyclohexane degrader oxidized 19% of the THDBT to three metabolites. The cometabolism of THDBT and DBT by the three 1-MN-grown Pseudomonas strains resulted in a much greater depletion of the condensed thiophenes than could be accounted for in the metabolites detected by GC analysis, but there was no evidence of sulfate release from DBT. These 1-MN-grown strains transiently accumulated 3-hydroxy-2-formylbenzothiophene (HFBT) from DBT, but it was subsequently degraded. On the other hand, Pseudomonas strain BT1d, which was maintained on DBT as a sole carbon source, accumulated 52% of the sulfur from DBT as HFBT over 7 days, and, in total, 82% of the sulfur from DBT was accounted for by the GC method used. Lyophilization of cultures grown on 1-MN with DBT and methyl esterification of the residues gave improved recoveries of total sulfur over that obtained by DCM extraction and GC analysis. This suggested that the further degradation of HFBT by these cultures leads to the formation of organosulfur compounds that are too polar to be extracted with DCM. We believe that this is the first attempt to quantify the products of DBT degradation by the so-called Kodama pathway.  相似文献   

20.
Ethyl isopropenyl ether reacts with D-glucose in N,N-dimethylformamide containing a trace of p-toluenesulfonic acid to give crystalline 4,6-O-isopropylidene-α,β-D-glucopyranose (2) in near-quantitative yield. The structure of 2 was established by n.m.r. spectroscopy of it and of its β-triacetate 3, and by conversion of 3 through deacetonation and subsequent acetylation into β-D-glucopyranose pentaacetate (5). The acetonation reagent operates under kinetic control, with favored attack at primary hydroxyl groups, instead of by the thermodynamic control associated with conventional acetonation methods. The reagents converts methyl α-D-glucopyranoside (7) into the 4,6-isopropylidene acetal 8, and D-mannitol (9) into a 2:1 mixture of the 1,2:5,6-di-isopropylidene acetal 10 and the 1,2:3,4:5,6-tri-isopropylidene acetal 11.  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