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.
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   

2.
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   

3.
The I-D-ribosyladenines have been obtained by treatment of 5-amino-4-cyanoimidazole (N-substituted or not) with ethyl N-(2,3-O-isopropylidene-D-ribofuranosyl)formimidate. The anomeric mixtures of the corresponding O-isoprophylidene nucleoside have been separated and the anomer fully characterized. In neutral aqueous medium, these compounds are transformed into an anomeric mixture of the corresponding 6-D-ribofuranosyladenine. In basic medium, however, anomerisation of the starting compounds to give an α,β equilibrium, in which the α anomer preponderates, takes place.  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.
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   

6.
Selective, double irradiation allows the assignment of most 13C-n.m.r. signals in a series of per-O-acetyl disaccharides composed of two D-glucose residues linked α-(1→3), β-(1→3), α-(1→4), β-(1→4), α-(1→6), β-(1→6), and α,α-(1→1). The main influences that affect the chemical shifts are discussed and the spectra of β-cellobiose octaacetate and β-maltose octaacetate are compared to those of cellulose and amylose triacetate, respectively, to show the possibilities and limitations of a disaccharide model for the interpretation of the 13C-spectrum of a polymer.  相似文献   

7.
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   

8.
Quaternary ammonium and phosphonium salts were readily obtained by treating 2,3,4,6-tetra-O-benzyl-α-D-galactopyranosyl bromide with tertiary amines and phosphines in various solvents under anhydrous conditions. Optical rotations and n.m.r. spectra of the hygroscopic syrups indicated that they exist mainly in the β-D configuration. Several dialkyl sulfides reacted very slowly with the galactosyl bromide and no conclusive evidence for sulfonium salt formation was obtained. 2,3,4,6-Tetra-O-benzyl-α-D-galactopyranosyl chloride failed to react with any of the nucleophiles.Methanolysis reactions of the phosphonium salts were too slow to be practical and were not studied extensively. Methanolyses of several quaternary ammonium salts in various solvents were not completely stereospecific, but gave good yields of methyl 2,3,4,6-tetra-O-benzyl-α-D-galactopyranoside. Attempted reactions of benzyl 2-O-benzoyl-4,6-O-benzylidene-β-D-galactopyranoside with quaternary ammonium salts derived from 2,3,4,6-tetra-O-benzyl-α-D-galactopyranosyl bromide failed to produce the corresponding derivative of 3-O-(α-D-galactopyranosyl)-D-galactose.  相似文献   

9.
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   

10.
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   

11.
Benzoylation of D-glycero-L-manno-heptono-1,4-lactone (1) with benzoyl chloride and pyridine for 2 h afforded crystalline penta-O-benzoyl-D-glycero-L-manno-heptono-1,4-lactone (2), but a large excess of reagent during 8 h also led to 2,5,6,7-tetra-O- benzoyl-3-deoxy-D-lyxo-hept-2-enono-1,4-lactone (3). Catalytic hydrogenation of 3 was stereoselective and gave 2,5,6,7-tetra-O-benzoyl-3-deoxy-D-galacto-heptono-1,4-lactone (4). Debenzoylation of 4 followed by oxidative decarboxylation with ceric sulfate in aqueous sulfuric acid gave 2-deoxy-D-lyxo-hexose (5). Application of the same reaction to 3-deoxy-D-gluco-heptono-1,4-lactone afforded 2-deoxy-D-arabino-hexose (6).  相似文献   

12.
Yves Lemoigne 《Geobios》1982,15(1):33-41
The study of numerous prints referred to the devonian genus Leptophloeum, more particularly from Kazakhstan and South Africa, has permitted to specify the morphology of the leaves and to explain the prints of the leaf cushions. The different species are revised. A comparison of the devonian genus Leptophloeum with the permian genus Lycopodiopsis leads us to consider the distinction of a new order: the Leptophloeodendrales regrouping these two genus.  相似文献   

13.
An acidic α-D-mannosidase has been isolated from the culture filtrate of a wood-rotting Basidiomycete, Pycnoporus sanguineus and the molecular and enzymatic properties of the enzyme determined. The extracellular mannosidase was homogeneous on PAGE at pH 9.4. The Mr as determined by SDS-polyacrylamide disc gel electrophoresis was 64000, and the pI was pH 4.7 using electrofocusing. The purified enzyme had a pH optimum of 4.5 with Baker's yeast mannan and had no activity towards p-nitrophenyl-α-mannoside. The Km and kcat values for Manα1-2Man at pH 4.5 and 30° were 0.9 mM and 1.9 sec. the enzyme had no activity towards Manα1-3Manα1-2Man, and it cleaved specifically the 1,2-α-linked side chain of yeast α-mannan, producing free α-D-mannose.  相似文献   

14.
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   

15.
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   

16.
In 1979 and 1980, batch culture experiments were conducted to observe the inhibitory effect of copper ion (concentrations of 10, 50, 100, 200 and 400 µg Cu · l–1) on the standing crops and photosynthesis of phytoplankton of the Saguenay River (for 124 hours) and in Chlorella vulgaris (for 8 days). These algal assays were carried out using the surface water of the Saguenay River. In natural populatoins of phytoplankton, it was found that photosynthesis was more sensitive than growth: at the lowest concentrations, such as 10 µg Cu · 1–1, copper seemed to increase the chlorophyll concentrations whereas the rates of primary production show a decrease of 60% with respect to the control. At higher concentrations of copper, the effect is weak in chlorophyll concentrations and more pronounced in the rates of primary production (decrease of 86 to 90%). The pennate diatoms are dominant (in all the samples) and these organisms are known as relatively resistant to copper. In Chlorella vulgaris, it was observed that with 100 µg Cu · 1–1, chlorophyll concentrations and rates of photosynthesis respectively decrease by 63 and 99% with respect to the control. At higher concentrations of copper, a maximum decrease of 70% and 99% respectively for chlorophyll concentrations and rates of primaryproduction are observed.
  相似文献   

17.
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   

18.
A new α-D-glucan, designated elsinan, has been isolated from the culture filtrate of Elsinoe leucospila grown in potato extract-sucrose medium. Acid hydrolysis of the methylated polysaccharide gave 2,3,6- and 2,4,6-tri-O-methyl-D-glucose, in the ratio of 2.5:1.0, together with small proportions of 2,3,4,6-tetra- (0.7%) and 2,4-di-O-methyl-D-glucose (0.5%), indicating that the glucan is an essentially linear polymer containing (1→4)- and (1→3)-α-D-glucosidic linkages. Periodate oxidation, followed by borohydride reduction and mild hydrolysis with acid (mild Smith degradation) yielded 2-O-α-D-glucosyl-D-erythritol and erythritol, in the molar ratio of 1.0:1.4, and a trace of glycerol. Partial acid hydrolysis, and also acetolysis, of elsinan gave nigerose, maltose, O-α-D-glucopyranosyl-(1→3)-O-α-D-glucopyranosyl (1→4)-D-glucopyranose, O-α-D-glucopyranosyl-(1→4)-O-α-D-glucopyranosyl-(1→3)-D-glucopyranose, maltotriose, and a small proportion of maltotetraose. It is concluded that elsinan is composed mainly of maltotriose residues joined by α-(1→3)-linkages, in the sequence →3)-α-D-Glcp-(1→4)-α-D-Glcp-(1→.The unique structural features of elsinan are discussed in comparison with other glucans.  相似文献   

19.
Irradiation of hydroxyalkyl glucosides in the presence of mercury(II) oxide and iodine gave glucosidic C-1-spiroorthoesters. Where the starting glycoside had CH2OH, the cyclization was regio- and stereo-specific by α-attack. The C-1-spiroorthoesters were readily available in a “one-pot synthesis” starting from glucosyl halides.  相似文献   

20.
Sterility is a potential toxic effect of chemotherapy. This risk is well established for alkylating agents, but is less clearly defined for anthracyclines, methotrexate and fluorouracil and poorly defined for alkaloids, platinum, etoposide and taxanes. The main predictive factors for ovarian toxicity are the additive effect of cytotoxic drugs, the cumulative dose of each drug and the patient’s age. This effect of chemotherapy is evaluated on menstrual cycles, hormonal assays and the number of pregnancies observed in patient cohorts. Chemotherapy induces destruction of oocytes and granulosa cells. In mice, it has been shown that adriamycin may induce oocyte apoptosis, which can be prevented by modulation of cycle cell signalling (dysregulation of Bax gene or, on the contrary, expression of its antagonist gene Bcl-2 or inhibition of apoptosis with sphingosine-1-phosphate or caspase inhibitors). Clinical data in the literature are usually based on retrospective studies and are somewhat confused: global fertility after MOPP chemotherapy for Hodgkin’s disease is about 20%, adjuvant chemotherapy with CMF, F(A)C or TAC for breast cancer induces amenorrhea in 50% to 70% of cases, PVB or BEP chemotherapy for ovarian germ cell tumors has little effect on fertility when the uterus and one ovary can be preserved, and the majority of women treated with methotrexate, actinomycin D or various combinations for persistent trophoblastic disease remain fertile. Preservation of fertility is a major goal for cancer patients receiving chemotherapy: in vitro fertilization could preserve the couple’s fertility, but is usually not feasible as it would delay initiation of chemotherapy until after stimulation of ovulation; oocyte or ovarian tissue cryopreservation is at the stage of research; oral contraceptives have not been demonstrated to be effective to preserve ovarian function; gonadotropin releasing hormone (GnRH) agonists prevent cyclophosphamide toxicity in rat and monkey ovaries, and a few pilot clinical studies suggest that chemotherapy-induced amenorrhea could be prevented by administration of GnRH analogues simultaneously to chemotherapy, but randomised studies are necessary.  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