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.
Starting from the finding that methyl 2,3,4,6-tetra-O-sulfonato-β-D-glucopyranoside (3) existed in a conformational equilibrium of the two chair conformers, the effect of sulfation on conformational equilibria was further investigated using a number of sulfated saccharides. Three sulfate groups on positions 3,4, and 6 or two on positions 2 and 3 were not sufficient to induce the conformational change as shown with methyl 2-amino-2-deoxy-3,4,6-tri-O-sulfonato-β-D-glucopyranoside. N-Sulfation of the amino group of the latter compound furnished an equilibrium of chair conformers with less 1 C 4 conformer content than for 3. The presence of persulfated methyl β-D-galactopyranoside in the usual 4 C 1 conformation suggested the involvement of the 4-O-sulfate in the effect. Methyl 2,3,4-tri-O-sulfonato-β-D-xylopyranoside was found to prefer the “all-axial” 1 C 4 conformation demonstrating that O-sulfates facilitate 1,3-O/O-diaxial interactions better than ester groups and in particular benzoates. Also, sulfated 1,5-anhydro-D-glucitol occurred as a conformational mixture, the influence of the anomeric effect may thus have been overestimated in the previous discussion of this conformational effect.  相似文献   

2.
Abstract

The oxidation of methyl 4,6-O-isopropylidene-α-d-glucopyranoside (1) via various routes to the dicarboxylate 2 is described. This reaction is used as a model for the oxidation of starch to dicarboxylic starch, a material with very promising properties as a cobuilder in detergents. The best oxidant found for C2-C3 cleavage was RuO4, prepared in situ by oxidation of a catalytic amount of RuIII with NaOCl.  相似文献   

3.
1-C-Nitroglycals. Preparation and Reaction with Some Nitrogen Nucleophiles Acetylation of the 1-deoxy-1-nitromannopyranoses 2 and 6 was accompagnied by spontanous β-elimination to give the 1-C-nitroglucals 3 and 7 , respectively, while acetylation of the gluco- and galacto-configurated 1-deoxy-1-nitropyranoses 8 and 14 gave the acetates 9 and 15 , respectively (Scheme 1). The acetylation of the ribo- and arabino-configurated 1-deoxy-1-nitrofuranoses 19 and 21 also occurred without β-elimination to give the acetates 20 and 22 , respectively (Scheme 2). Mild base treatment of the previously described O-acetylnitro-β-D -glucose 4 , the O-acetylnitro-β-D -pyranoses 9 and 15 , and the O-acetylnitro-β-D -furanoses 17 , 20 , and 22 gave the 1-C-nitroglycals 3 , 10 , 16 , 18 and 23 , respectively (Scheme 1 and 2). The previously obtained 1-C-nitroglucal 3 was deacetylated by treatment with MeOH in the presence of KCN or sodium m-nitrophenolate to give the free nitroglucal 5 . Deacetylation of the benzylidene protected 1-C-nitroglucal 10 (MeOH, NaOMe) gave the 4,6-O-benzylidene-1-C-nitroglucal 11 and traces of the 2-O-methyl-1-C-nitromannoses 12 and 13 . The UV, IR, 1H-NMR and 13C-NMR spectra of the 1-C-nitroglycals are discussed. In solution, the 1-C-nitroglycals 1 , 5 , 7 , 10 , 11 , and 16 adopt approximately a 4H5? and 3 a flattened 4H5 conformation. The structure of 5 was established by X-ray analysis. In the solid state, 5 adopts a sofa conformation, which is stabilized by an intramolecular H-bond. The β-addition of NH3 to the 1-C-nitroglucals 7 and 10 was followed by an O→ N acetyl migration to give exclusively anomeric pairs of the N-acetyl-1-nitromannosamine derivatives 24 / 25 and 26/27 , respectively (Scheme 3). The β-addition of methylamine, octadecylamine, and tryptamine to the 1-C-nitroglucal 11 also stereoelectronically controlled and gave the crystalline N-alkyl-1-nitromannosamines 28 , 29 , and 30 , respectively. The stereoelectronically controlled β-addition of NH3 to the 1-C-nitrogalactal 16 , followed by acetylation, yielded exclusively the talosamine derivative 31 , while the reversible β-addition of azide ions to 16 gave the anomeric 2-azido-1-nitrogalactoses 32 and 33 . The β-addition of azide ions to the 1-C-nitroglucal 1 led to the 2-azido-1-nitromannose 34 . In the presence of excess formaldehyde, this addition was followed by a Henry reaction. Chromatography of the crude product was accompagnied by solvolytic removal of the NO2 group to give the 3-azidomannoheptulose 35 in high yields (Scheme 4).  相似文献   

4.

Six secondary metabolites from the methanolic extract of Sweetia panamensis (Fabaceae) bark were isolated and characterised. Along with the pyrones desmethylangonine β-d-O-glucopyranoside and desmethylangonine β-d-O-glucopyranosyl-(1→6)-O-β-d-glucopyranoside, already reported in this species, 5-O-caffeoylquinic acid (chlorogenic acid), 4-O-caffeoylquinic acid, 3-O-caffeoylquinic acid and the isoflavonoid 5-O-methylgenistein 7-O-β-d-glucopyranoside were isolated for the first time from S. panamensis. Additionally, an LC-ESI-MS qualitative analysis was performed and an ultra performance liquid chromatography (UPLC) method was developed and validated for the determination of these compounds. The UPLC method was applied to the quantitative analysis of plant samples. Pyrones and caffeoylquinic acids resulted to be the main compounds in the extract; in particular desmethylangonine β-d-O-glucopyranosyl-(1→6)-O-β-d-glucopyranoside was the most abundant compound.

  相似文献   

5.
Twelve compounds were isolated from Chrozophora tinctoria (L.) Raf. They were identified as kaempferol, kaempferol 3-O-β-glucopyranoside, kaempferol 3-O-(6″-α-rhamnopyranosyl)-β-glucopyranoside, quercetin, quercetin 3-O-β-glucopyranoside, quercetin 3-O-(6″-α-rhamnopyranosyl)-β-glucopyranoside, apigenin, apigenin 7-O-β-glucopyranoside, acacetin, gallic acid, methyl gallate and β-sitosterol-3-O-β-glucopyranoside. Their structures were elucidated by chemical and spectral methods. Furthermore, chemosystematics of the isolated compounds is briefly discussed. It was indicated that C. tinctoria is the only species of Chrozophora that has the capability to synthesis kaempferol aglycone and their glycosides, and the finding is supported by its distinct morphological and anatomical aspects.  相似文献   

6.
The configurations of 1,4:3,6-dianhydro-2,5-di-O-mesyl-D -mannitol, 1,4:3,6-dianhydro-2,5-di-O-mesyl-D -glucitol, 1,4:3,6-dianhydro-2,5-di-O-mesyl-L -iditol, 1,4:3:6-dianhtydro-2-deoxy-2-iodo-5-O-mesyl-D -mannitol, 1,4:3:6-dianhtydro-2-deoxy-2-iodo-5-O-mesyl-D -glucitol, 1,4:3,6-dianhydro-2-deoxy-2-iodo-5-O-mesyl-L -iditol, 1,4:3,6-dianhydro-2,5-dideoxy-2,5-diiodo-D -glucitol and 1,4:3,6-dianhydro-2,5-dideoxy-2,5-diiodo-L -iditol were determined by 13C NMR spectroscopy, by invoking the field-effect.  相似文献   

7.
Glucose oxidase from Aspergillus niger, the specific enzyme for β-d-glucose oxidation, can also oxidize other related saccharides at very slow or negligible rates. The present study aimed to compare the kinetics of d-glucose oxidation using immobilized glucose oxidase on bead cellulose for the oxidation of related saccharides using the same biocatalyst. The significant differences were observed between the reaction rates for d-glucose and other saccharides examined. As a result, k cat/K M ratio for d-glucose was determined to be 42 times higher than d-mannose, 61.6 times higher than d-galactose, 279 times higher than d-xylose, and 254 times higher than for d-fructose and d-cellobiose. On the basis of these differences, the ability of immobilized glucose oxidase to remove d-glucose from d-cellobiose, d-glucose from d-xylose, and d-xylose from d-lyxose was examined. Immobilized catalase on Eupergit and mixed with immobilized glucose oxidase on bead cellulose or co-immobilized with glucose oxidase on bead cellulose was used for elimination of hydrogen peroxide from the reaction mixture. The accelerated elimination of d-glucose and d-xylose in the presence of co-immobilized catalase was observed. The co-immobilized glucose oxidase and catalase were able to decrease d-glucose or d-xylose content to 0–0.005% of their initial concentrations, while a minimum decrease of low oxidized saccharides d-xylose, d-cellobiose, and d-lyxose, respectively, was observed.  相似文献   

8.
Dried leaves of Marsdenia erecta R. Br. gave over 6% of a crude glycoside mixture, the main portion of which consisted of weakly polar material, soluble in ether or chloroform. By mild acid hydrolysis it yielded crude sugars and aglycones. The following four crist. sugars were isolated: D -cymarose, D -oleandrose and two bioses: pachybiose and marsectobiose (C14H24O8, new). By PC. and TLC. the presence of digitoxose, canarose, thevetose and 3-O-methyl-6-deoxyallose could be demonstrated. The crude acyl-genin mixture contained β-sitosteryl-β-D -glucopyranoside and three highly hydroxylated pregnane derivatives: drevogenin-P, 17 β-marsdenin (C21H32O6, new) and marsectohexol (C21H34O6, new), all partly esterified with acetic, tiglic and benzoic acid. Five crist. acyl-genins (A1–A5) were isolated by chromatography, but most of them still were mixtures. After alkaline hydrolysis of the crude acyl-genins 6 acyl-free compounds were obtained. 4 of them were identical with the above mentioned substances, the other two: 17-iso-drevogenin-P and marsdenin (17 α-marsdenin) are formed from drevogenin-P and 17 β-marsdenin by isomerisation.  相似文献   

9.
High-speed counter-current chromatography (HSCCC) was applied to the preparative isolation and purification of peonidin 3-O-(6-O-(E)-caffeoyl-2-O-β-D -glucopyranosyl-β-D -glucopyranoside)-5-O-β-D -glucoside ( 1 ), cyanidin 3-O-(6-O-p-coumaroyl)-β-D -glucopyranoside ( 2 ), peonidin 3-O-(2-O-(6-O-(E)-caffeoyl-β-D -glucopyranosyl)-6-O-(E)-caffeoyl-β-D -glucopyranoside)-5-O-β-D -glucopyranoside ( 3 ), peonidin 3-O-(2-O-(6-O-(E)-feruloyl-β-D -glucopyranosyl)-6-O-(E)-caffeoyl-β-D -glucopyranoside)-5-O-β-D -glucopyranoside ( 4 ) from purple sweet potato. Separation of crude extracts (200 mg) from the roots of purple sweet potato using methyl tert-butyl ether/n-butanol/acetonitrile/water/trifluoroacetic acid (1:4:1:5:0.01, v/v) as the two-phase solvent system yielded 1 (15 mg), 2 (7 mg), 3 (10 mg), and 4 (12 mg). The purities of 1 – 4 were 95.5%, 95.0%, 97.8%, and 96.3%, respectively, as determined by HPLC. Compound 2 was isolated from purple sweet potato for the first time. The chemical structures of these components were identified by 1H NMR, 13C NMR and ESI-MSn.  相似文献   

10.
Two new quercetin glycoside derivatives named quercetin-3-O-[2-O-trans-caffeoyl-α-l-rhamnopyranosyl-(1 → 6)-β-d-glucopyranoside] (1) and quercetin-3-O-[2-O-trans-caffeoyl-β-l-rhamnopyranosyl-(1 → 6)-β-d-glucopyranoside] (2) along with three known flavonoids, 5-hydroxy-6,7,3′,4′,5′-pentamethoxyflavone (3), 5,7-dihydroxy-8-methoxyflavone (4) and kaempferol 3-O-β-d-glucopyranoside (5), were isolated from the fruits of Gardenia jasminoides var. radicans. The structures of the new compounds were determined by means of extensive spectroscopic analysis (1D, 2D NMR and HR-ESI-MS), glycoside hydrolysis and sugar HPLC analysis after derivatisation. This is the first report on the isolation of a pair of compounds with α or β-l-rhamnopyranosyl configuration from plant and the first detail assignment of their NMR data.  相似文献   

11.
12.
End-functionalized (1→6)-2,5-anhydro-3,4-di-O-methyl-D-glucitols ( 3a–c ) were synthesized by the anionic cyclopolymerization of 1,2:5,6-dianhydro-3,4-di-O-methyl-D-mannitol ( 1 ), followed by treatment with a terminating agent such as 4-vinylbenzyl ( 2a ), oxetanyl ( 2b ), and methacryloyl group ( 2c ). The end-functionalization proceeded in a high efficiency at 73–98%. The radical copolymerization of styrene with 3a yielded a polymer ( 5a ) whose GPC trace exhibited a unimodal peak. 5a was polystyrene with (1→6)-2,5-anhydro-3,4-di-O-methyl-D -glucitol as pendant groups whose structure was confirmed by the 1H NMR spectrum.  相似文献   

13.
Isolation and Structure Elucidation of Neapolitanose (O-β-D -Glucopyranosyl-(1→2)-O-[β-D -glucopyranosyl-(1→6)]-D -glucose), New Trisaccharide from the Stigmas of Garden Crocusses (Crocus neapolitanus var.) From the stigmas of Crocus neapolitanus var. ‘Blue Bird’ two new crocetin glycosyl esters were isolated. They contained a hitherto unknown trisaccharide. For the structure elucidation a homonuclear 2D-1H-NMR-shift-correlation experiment was carried out with the peracetate of the isolated trisaccharide. This experiment revealed that the carbohydrate is O-β-D -glucopyranosyl-(1→2)-O-[β-D -glucopyranosyl-(1→6))]-D -glucose, for which we suggest the name ‘neapolitanose’. The two new C20-carotenoids from Crocus neapolitanus are crocetin (β-gentiobiosyl) (β-neapolitanosyl) ester ( 4 ) and crocetin di(β-neapolitanosyl) ester ( 5 ).  相似文献   

14.
In an effort to introduce N-chloroarylsulfonamides of different oxidizing strengths, nine sodium salts of mono- and di-substituted N-chloroarylsulfonamides are employed as oxidants for studying the kinetics of oxidation of D -fructose and D -glucose in aqueous alkaline medium. The results are analyzed along with those by the sodium salts of N-chlorobenzenesulfonamide and N-chloro-4-methylbenzenesulfonamide. The reactions show first-order kinetics each in [oxidant], [Fru/Glu], and [OH]. The rates slightly increase with increase in ionic strength of the medium. Further, the rate of oxidation of fructose is higher by 4 to 5 times than that of the glucose oxidation, by the same oxidant. Similarly, Ea values for glucose oxidations are higher by about 1.5 times the Ea values for fructose oxidations. The results have been explained by a plausible mechanism, and the related rate law deduced. The significant changes in the kinetics and thermodynamic data are observed with change of substituent in the benzene ring. It is because Cl+ is the effective oxidizing species in the reactions of N-chloroarylsulfonamides. The oxidative strengths of the latter therefore depend on the ease with which Cl+ is released from them. The ease with which Cl+ is released from N-chloroarylsulfonamides depends on the electron density of the nitrogen atom of the sulfonamide group, which in turn depends on the nature of the substituent in the benzene ring. The following Hammett equations are valid for the oxidation of fructose and glucose, log kobs = −3.13 + 0.54 σp and log kobs = −3.81 + 0.28 σp, respectively. The enthalpies and entropies of activations for oxidations by all the N-chloroarylsulfonamides correlate well with isokinetic temperatures of 301 K and 299 K, for fructose and glucose oxidations, respectively. The effect of substitution in the oxidants on the Ea and log A for the oxidations is also considered. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 572–582, 2005  相似文献   

15.
Reactions of Some 3-Deoxy-3-trifluoroacetamido-D -ribofuranosyl Halides with Mercuric Cyanide The expected 2,5-anhydro-D -allononitrile 11 was obtained when the 2-O-benzoyl-3-deoxy-3-trifluoroacetamido-D -ribofuranosyl bromide reacted with mercuric cyanide, whereas the isomeric 1,2-O (1-cyanoethylidene) derivative was formed from the 2-O-acetylated halide; no reaction occurred when the halide was 2-O-p-nitrobenzoylated.  相似文献   

16.
The triphenylstannyl β-D -glucopyranoside 4 was synthesized in one step from the 1,2-anhydro-α-D -glucopyranose 3 with (triphenylstannyl)lithium (Scheme 1). Transmetallation of 4 with excess BuLi, followed by quenching the dianion 7 with CD3OD gave (1S)-1,5-anhydro-3,4,6-tri-O-benzyl-[1-2H]-D - glucitol ( 8 ) in 81% yield (Scheme 2). Trapping of 7 with benzaldehyde, isobutyraldehyde, or acroleine gave the expected β-D -configurated products 11, 12 , and 13 in good yields. Preparation of C-acyl glycosides from acid chlorides, such as acetyl or benzoyl chloride was not practicable, but addition of benzonitrile to 7 yielded 84% of the benzoylated product 14 . Treatment of 7 with MeI led to 15 (30%) along with 40% of 18 , C-alkylation being accompanied by halogen-metal exchange. Prior addition of lithium 2-thienylcyanocuprate increased the yield of 15 to 50% and using dimethyl sulfate instead of MeI led to 77% of 15 . No α-D -anomers could be detected, except with allyl bromide as the electrophile, which yielded in a 1:1 mixture of the anomers 16 and 17 .  相似文献   

17.
C28H36O10. KSCN is monoclinic, space groupP21 withZ=2,a=10.390(3),b=8.959(7),c=16.377(7) Å, =92.49(5)°. FinalR=0.053 for 1437 reflections measured at room temperature. The K ion lies on the least-squares plane formed by the six oxygen atoms in the macrocyclic ring. The SCN ion was found on the same face of the macrocycle as the chiral glucopyranoside moiety.Methyl-4,6-O-benzylidene-2,3-O-(1,2-bis(ethoxyethoxy)benzenediyl)--d-glucopyranoside.  相似文献   

18.
The isolation and identification of fifteen crystalline components from the whole herb of Clinopodium umbrosum (Bieb.) C. Koch (Labiatae) are described. Their structures were determined on the basis of spectral evidence and chemical transformation. These compounds include five steroids (α-spinasterone, β-sitosterol, stigmasterol, α-spinasterol, and α-spinasteryl-3-O-β-glucopyranoside), four triterpenoids (3β-hydroxyurs-11-en-28,13-olide, betulinic acid, oleanolic acid, ursolic acid), four flavonoids (luteolin, luteolin-7-O-β-glucopyranoside, apigenin-7-O-β-glucuronide, and apigenin-7-O-β-methylglucuronate), and two lignolic acids [3-(3,4-dihydroxyphenyl)- lactic acid and rosmarinic acid].  相似文献   

19.
In continual study on the heartwood of Rhamnus nakaharai, a new alaternin-8-O-glucoside, namely 1,2,6,8-tetrahydroxy-3-methylanthraquinone-8-O-β-glucopyranoside (1), together with some known compounds were further isolated and characterised by 1-D, 2-D NMR and other spectral evidences. The free radical scavenging and antityrosinase activities of the isolates, including alaternin (1a), emodin (2a), emodin-8-O-β-glucopyranoside (2), 6-methoxysorigenin-8-O-β-glucopyranoside (3) and 6-methoxysorigenin (3a) were tested. Alaternin (1a) exhibited to be mild DPPH radical scavenger with half as potent as vitamin C, while both alaternin (1a) and emodin-8-O-β-glucopyranoside (2) exhibited stronger SOD-like activity than that of BHA. 6-Methoxysorigenin (3a), a reported potential antioxidant, and its 8-O-glucoside (3) both performed significant inhibitory effect on mushroom tyrosinase with about twice as potent as kojic acid, the positive control.  相似文献   

20.
The ethanolic extract of the stems of Viburnum fordiae Hance showed insecticidal and α-glucosidase inhibitory activities and then was fractionated by bioactivity-guided fractionation to obtain a rare C13-norisoprenoid (1), together with a new phenolic glycoside (2), and seven known compounds, alangionoside C (3), pisumionoside (4), koaburaside (5), 3,5-dimethoxy-benzyl alcohol 4-O-β-d-glucopyranoside (6), 3,4,5-trimethoxybenzyl-β-d-glucopyranoside (7), arbutin (8), and salidroside (9). The previously undescribed compounds were elucidated as (3R,9R)-3-hydroxy-7,8-didehydro-β-ionyl 9-O-α-d-arabinopyranosyl-(1→6)-β-d-glucopyranoside (1) and 2-(4-O-β-d-glucopyranosyl)syringylpropane-1,3-diol (2) by spectroscopic data (1H and 13C NMR, HSQC, HMBC, 1H-1H COSY, HSQC-TOCSY, HRESIMS, IR and ORD) and chemical methods. Compound 1 showed potent insecticidal effect against Mythimna separata with LD50 value of 140 μg g?1. Compounds 2, 5, 6, 8 and 9 showed varying α-glucosidase inhibitory activity with IC50 values ranging from 148.2 to 230.9 μM.  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