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 46 毫秒
1.
The 13C NMR spectra of some polysaccharides and their methyl derivatives have been analysed. The numbers and positions of the assigned 13C NMR signals give some information about the structure of the monomer unit and the positions of the glycosidic linkage but no information about the anomeric configuration. In this case the 1J(C-1, H) coupling constants make it possible to identify the anomeric configuration, because the mean differences of the J values for the α- and β-anomers are 12 Hz (at least 5 Hz) with the higher values for the α-anomers.  相似文献   

2.
29Si NMR spectra of the O-trimethylsilyl (OTMS) derivatives of various methyl α- and β-D -galactopyranosides have been recorded. The effect of changes in the anomeric configuration provides a means of assigning the resonance of the 2-OTMS substituent. Whereas the signal of the OTMS group attached at the 6-position can be assigned readily, those of the OTMS group at the 3- or 4-position cannot be assigned unequivocally.  相似文献   

3.
ABSTRACT

3,4-Di-O-acetyl-2-deoxy-2-nitroso-α-D-xylo-pentopyranosyl chloride (2) reacts with pyrazole to afford 1-[3,4-di-O-acetyl-2-deoxy-2-(Z)-hydroxyimino-α- (3) and β-D-threo-pentopyranosyl]pyrazole (4). The products of condensation were modified at C-2 or C-3 to give pyrazole derivatives with 3-azido-2,3-dideoxy-2-hydroxyimino-pentopyranosyl (5,7,8,9,10), 2-acetoxyimino-2,3-dideoxy-β-D-glycero-pentopyranosyl (12,13), β-D-lyxo- (14), β-D-xylopentopyranosyl (15) structures and 2,3-dihydro-2-pyrazol-1-yl-6H-pyran-3-one oximes (6,11). The conformation of the sugar residue and configuration at the anomeric centre and of the hydroxyimino group were established on the basis of 1H NMR and polarimetric data.  相似文献   

4.
2′-Deoxy-D-ribonucleoside analogs of biologically active 2-β-D-ribofuranosylthiazole-4-carboxamide were synthesized and the structure of the β anomer was determined by X-ray crystallography and ′1H nmr. 2′-Methylene protons of α- and β-deoxyribonucleosides were observed to exhibit characteristic patterns in the 1H nmr which was used to distinguish between the two anomers. The method could be used to determine the anomeric configuration of both N- and C-2′-deoxyribonucleosides.  相似文献   

5.
The regio- and stereoselectivity of the glycosidation of the partially protected mono-alcohols 3 and 7 , the diols 2 and 8 , and the triol 4 by the diazirine 1 have been investigated. Glycosidation of the α-D -diol 2 (Scheme 2) gave regioselectively the 1,3-linked disaccharides 11 and 12 (80%, α-D /β-D 9:1), whereas the analogous reaction with the βD -anomer 8 led to a mixture of the anomeric 1,3- and 1,4-linked disaccharides 13 (12.5%), 14 (16%), 15 (13%), and 16 (20.5%; Table 2). Protonation of the carbene by OH–C(4) of 2 is evidenced by the observation that the α-D -mono-alcohol 3 did not react with 1 under otherwise identical conditions, and that the β-D -alcohol 7 yielded predominantly the β-D -glucoside 18 (52%) besides 14% of 17 . Similarly as for the glycosidation of the diol 2 , the influence of the H-bond of HO? C(4) on the direction of approach of the carbene, the role of HO? C(4) in protonating the carbene, and the stereoelectronic control in the interception of the ensuring oxycarbenium cation are evidenced by the reaction of the triol 4 with 1 (Scheme 3), leading mostly to the α-D -configurated 1,3-linked disaccharide 19 (41%), besides its anomer 20 (16%), and some 4-substituted β-D -glucoside 21 (9%). No 1,6-linked disaccharides could be detected. In agreement with the observed reactivity, the 1H-NMR and IR spectra reveal a strong H-bond between HO? C(3) and the phthalimido group in the α-D -, but not in the β-D -allosides. The different H-bonds in the anomeric phthalimides are in keeping with the results of molecular-mechanics calculations.  相似文献   

6.
Model reactions of 0-(α-D-glucopyranosyl)trichloroacetimidate 2α with methanol and choTesterol under various conditions demon-strated that stereocontrolled glucosyl transfer with inversion of configuration at the anomeric center is best carried out in di-chloromethane at low temperatures with boron trifluoride-ether as a catalyst. Under these conditions β-glucoside 4β and β-disaccha-rides 5β- 9β were obtained in good to excellent yields.

With Brtosnsted acids, fast glucosyl transfer to the acid anion was mainly observed and required no further acidic catalysis. With strong acids formation of the thermodynamically more stable product dominated. However, with the weaker carboxylic acids highly diastereoselective inversion of configuration at the anomeric center led, for instance, to β-1-O-acyl derivatives 11β - 18β, revealing a convenient method for the synthesis of O-glycosyl-carb-oxylates. This method was also applied to resolution of racemic carboxylic acids.

Similar results were obtained with N-nucleophiles. Hydrazoic acid gave exclusively α-azide 19a. Nitrogen heterocycles gave with boron trifluoride-ether catalysis mainly β-nucleosides 20β - 23β. Reaction of trichloroacetimidate 2α with O-nucleophiles in aceto-nitrile as solvent led to different products due to competition  相似文献   

7.
Abstract

Partially protected benzyl α-and β-pyranosides undergo Wittig rearrangement reactions on treatment with strong bases in tetrahydrofuran to give hydroxymethylphenyl C-glycosyl derivatives. Two products were generally obtained and all Wittig rearrangement products retained the configuration at the migrating anomeric center. In benzene, benzyl 4,6-O-isopropylidene-β-D-glucopyranoside reacted with n-butyl lithium to give addition to the anomeric center accompanied by ring opening with loss of the aglycone. Allyl glycosides do not give Wittig rearrangement products.  相似文献   

8.
Solid-liquid phase-transfer glycosylation (KOH, tris[2-(2-methoxyethoxy)ethye]amine ( = TDA-1), MeCN) of pyrrolo[2,3-d]pyrimidines such as 3a and 3b with an equimolar amount of 5-O-[(1,1 -dimethylethyl)dimethylsilyl]-2,3-O-(1-methylethylidene)-α-D -ribofuranosyl chloride (1) [6] gave the protected β-D -nucleosides 4a and 4b , respectively, stereoselectively (Scheme). The β-D -anomer 2 [6] yielded the corresponding α-D -nucleosides 5a and 5b with traces of the β-D -compounds. The 6-substituted 7-deazapurine nucleosides 6a , 7a , and 8 were converted into tubercidin (10) or its α-D -anomer (11) . Spin-lattice relaxation measurements of anomeric ribonucleosides revealed that T1 values of H? C(8) in the α-D -series are significantly increased compared to H? C(8) in the β-D -series while the opposite is true for T1 of H? C(1′). 15N-NMR data of 6-substituted 7-deazapurine D -ribofuranosides were assigned and compared with those of 2′-deoxy compounds. Furthermore, it was shown that 7-deaza-2′deoxyadenosine ( = 2′-deoxytubercidin; 12 ) is protonated at N(1), whereas the protonation site of 7-deaza-2′-deoxyguanosine ( 20 ) is N(3).  相似文献   

9.
The disaccharide methyl (4-O-benzoyl-3-O-benzyl-2-O-acetyl-α-L-rhamno pyranosyl)-(1→4)-(allyl 2,3-di-O-benzyl-β-D-galactopyranosid)uronate (13) was obtained in an excellent yield of 88% using methyl (allyl 2,3-di-O-benzyl-β-D-galactopyranosid)uronate ((12) as the glycosyl acceptor and a slight excess of the 1,2-di-O-acetyl-rhamnoglycosyl donor 5a. Disaccharide 13 is a key intermediate that can be used either as a glycosyl acceptor or glycosyl donor for the preparation of rhamnogalacturonan fragments. Here, introduction of the trichloroacetimidate function at the anomeric center gave the disaccharide glycosyl donor 28, which could be applied in a blockwise glycosylation reaction to form the L-Rha-α(1→4)-D-GalA-α(1→4)-D-GalA trisaccharide 29. Generally, on condition that no neighboring group effect influenced the reaction at the anomeric center of the α-trichloroacetimidate galacturonate glycosyl donors (20–22, 28), α-glycosidic linkages were nearly exclusively formed, except in the case of the 4-O-methylgalactopyranosyluronate 22.  相似文献   

10.
The two anomeric 1-(2′-deoxy-D -ribofuranosyl)-2(1H)-pyridones 6 and 7 were synthesized from 2-pyridone and 3,5-di-(O-p-toluoyl)-2-deoxy-D -ribofuranosyl chloride ( 2 ) via the di-O-p-toluoyl derivatives 3 and 4 using the mercuric halide procedure. Phosphorylation of the nucleosides 6 and 7 by bis-(2,2,2-trichloroethyl)-chlorophosphate gave the phosphate esters 8 and 9 together with some 2-(bis-[2,2,2-trichloroethyl]-phosphoryloxy)-pyridine 10 , which proved to be very labile. Structure and configuration of compounds 6 to 9 were established by spectral methods, the configurations being derived from the chemical shifts of the sugar protons and the splitting patterns of the anomeric protons (‘triplet-quartet rule’). The specific rotations of 3 , 4 , 6 , 7 , 8 and 9 show that the three pairs of anomers represent exceptions to Hudson's rule of isorotation. Reductive removal of the trichloroethyl groups in 8 and 9 with zinc proceeds stepwise, yielding the phosphoro-diesters 13 and 14 and the two desired anomeric 5′-nucleotides 15 and 16 . These latter were purified and characterised as the ammonium salts. Enzymatic cleavage by the 5′-nucleotidase of Crotalus adamanteus venom took place only in the ‘natural’ β-series. The ‘unnatural’ α-anomers were resistent to the enzyme. The structure of 10 was established by spectral methods and confirmed by synthesis.  相似文献   

11.
2,3;5,6-Di-O-isopropylidene-α-d-mannofuranosyl chloride reacted with allylmagnesium bromide with preferential inversion of the anomeric configuration to furnish a mixture of the 1-allyl-1-deoxy-β- and α-d-mannofuranoses. Separation of β and α derivatives was possible only after conversion to the 1-allyl-1-deoxylyxofuranoses 2 and 3. The β configuration of the predominant product 2 was proved using the NOE method.  相似文献   

12.
Anion recognition properties of d-ribose-based receptors α- and β-1 were measured by 1H NMR in CDCl3 and MeCN-d3. Receptor β-1 showed effective binding with anions by cooperative hydrogen bonds of cis-diol. The anomeric isomer α-1 is a less effective anion receptor which has similar cis-diol as a recognition site, indicating that the stereo configuration of the anomeric position is of significant influence on the anion recognition ability.  相似文献   

13.
Abstract

N-Phthaloyl, N-acetyl, N-benzyl, N-acetyl-N-methyl, N,N-dimethyl, N-benzoyl, and N,N-dibenzoyl derivatives of the trisaccharide β-D-glucopyranosyl-(1-3)-O-(2-deoxy-2-amino-β-D-glucopyranosyl)-(1-4)-O-β-D-glucopyranose were synthesized and analyzed by FAB MS. The intensity ratios of the peaks resulting from cleavage of the anomeric bond of the glucosamine residue and the respective molecular ion peaks turned out to be high for the N-acyl derivatives and up to two orders of magnitude lower for the N-alkyl compounds. These results show that fragmentation at the anomeric carbon of the amino sugar may be assisted by the carbonyl group and the resulting cation is stabilized by delocalization of the positive charge.  相似文献   

14.
The synthesis of new deoxyribose nucleosides by coupling chloropurines with modified D -ribose derivatives is reported. The methyl 2-deoxy-N-methyl-3-O-(p-toluoyl)-α-D -ribofuranosiduronamide (α-D - 8 ) and the corresponding anomer β-D - 8 were synthesized starting from the commercially available 2-deoxy-D -ribose ( 1 ) (Scheme 1). Reaction of α-D - 8 with the silylated derivative of 2,6-dichloro-9H-purine ( 9 ) afforded regioselectively the N9-(2′-deoxyribonucleoside) 10 as anomeric mixture (Scheme 2), whereas β-D - 8 did not react. Glycosylation of 9 or of 6-chloro-9H-purine ( 17 ) with 1,2-di-O-acetyl-3-deoxy-N-methyl-β-D -ribofuranuronamide ( 13 ) yielded only the protected β-D -anomers 14 and 18 , respectively (Scheme 3). Subsequent deacetylation and dechlorination afforded the desired nucleosides β-D - 11 , β-D - 12,15 , and 16 . The 3′-deoxy-2-chloroadenosine derivative 15 showed the highest affinity and selectivity for adenotin binding site vs. A1 and A2A adenosine receptor subtypes.  相似文献   

15.
Abstract

(3-Oxo)-butyl mannopyranosides undergo Norrish type II photocyclization giving hydroxyspiroketals of which the structures have been established by NMR spectroscopy and X-ray analysis. The influence of the anomeric and C-2 configurations have been studied and compared to the gluco derivatives. The favored photocyclization of β-mannoside occurs with retention of configuration and that of α-mannoside with inversion of configuration.  相似文献   

16.
《合成通讯》2013,43(16):2499-2506
2,3-Dibromo-2-methyl-N-(1-adamantyl)propanamide (4), a precursor equally suited for the preparation of an α-lactam and a β-lactam, upon treatment with sodium tert-butoxide ether gives no α-lactam (5), but an excellent yield of the isomeric β-lactam, 1-(1-adamantyl)-3-bromo-3-methylazetidinone (6) as the only product. Repeating the experiment using a large excess of sodium tert-butoxide still leads to β-lactam 6 in 76.1% yield, but now accompanied by its dehydrobrominated derivative, β-lactam 7, in 17.4% yield, and no trace of α-lactam 5  相似文献   

17.
We describe the synthesis of 2′-deoxy-3′,5′-ethano-D -ribonucleosides 1 – 8 (= (5′,8′-dihydroxy-2′-oxabicyclo-[3.3.0]oct-3′-yl)purines or -pyrimidines) of the nucleobases adenine, thymine, cytosine, and guanine. They differ from natural 2′-deoxyribonucleosides only by an additional ethylene bridge between the centers C(3′) and C(5′). The configuration at these centers (3S,5′R) was chosen as to match the geometry of a repeating nucleoside unit in duplex DNA as close as possible. These nucleosides were designed to confer, as constituents of an oligonucleotide chain, a higher degree of preorganization of a single strand for duplex formation with respect to natural DNA, thus leading to an entropic advantage for the pairing process. The synthesis of these ‘bicyclonucleosides’ was achieved by construction of an enantiomerically pure carbohydrate precursor 18 / 19 (Schemes 1), which was then converted to the corresponding nucleosides by known methods in nucleoside synthesis (Schemes 2 and 3). In all cases, both anomeric forms of the nucleosides were obtained in pure crystalline form, the relative configuration of which was established by 1H-NMR-NOE spectroscopy. A conformational analysis of the nucleosides with β-configuration at the anomeric center by means of X-ray and 1H-NMR (including NOE) spectroscopy show the furanose part of the molecules to adopt uniformly a 1′exo-conformation with the base substituents preferentially in the anti-range in the pyrimidine nucleosides (anti/syn ca. 2:1) distribution in the purine nucleosides (in solution).  相似文献   

18.
Nucleoside configuration (α-d vs. β-d ), nucleobase substituents, and the helical DNA environment of silver-mediated 5-aza-7-deazaguanine-cytosine base pairs have a strong impact on DNA stability. This has been demonstrated by investigations on oligonucleotide duplexes with silver-mediated base pairs of α-d and β-d anomeric 5-aza-7-deaza-2′-deoxyguanosines and anomeric 2′-deoxycytidines incorporated in 12-mer duplexes. To this end, a new synthetic protocol has been developed to access the pure anomers of 5-aza-7-deaza-2′-deoxyguanosine by glycosylation of either the protected nucleobase or its salt followed by separation of the glycosylation products by crystallization and chromatography. Thermal stability measurements were performed on duplexes with α-d /α-d and β-d /β-d homo base pairs or α-d /β-d and β-d /α-d hybrid pairs within two sequence environments, positions 6 or 7, of oligonucleotide duplexes. The respective Tm stability increases observed after silver ion addition differ significantly. Homo base pairs with β-d /β-d or α-d /α-d nucleoside combinations are more stable than α-d /β-d hybrid base pairs. The positional switch of silver-ion-mediated base pairs has a significant impact on stability. Nucleobase substituents introduced at the 5-position of the dC site of silver-mediated base pairs affect base pair stability to a minor extent. Our investigation might lead to applications in the construction of bioinspired nanodevices, in DNA diagnostics, or metal-DNA hybrid materials.  相似文献   

19.
ABSTRACT

The direct coupling of 3-n-alkyl catechols to the acetate or trichloroacetimidate derivatives of β-D- or α-D-glycosides (glucose, galactose, xylose, mannose and maltose) catalyzed by BF3Ot2 has been studied. β-Glycosides with an equatorial acetate group at position 2 formed exclusively β adducts with yields of 60–80%. α-Glycosides with an equatorial acetate group at position 2 formed β adducts, while β-glycosides with an axial acetate group formed α adducts when activated as trichloroacetimidates, with yields of 70–85%. This was applied to the coupling of 3-n-alkylcatechols of increasing chain length (up to C15) to sugar derivatives. The coupling position of glycosides on the catechol was determined either by differential NOE experiments and by the regioselective synthesis of 1-(O-β-D-glucopyranosyl)-3-pentadecylcatechol, a water soluble analogue of the poison ivy skin allergen. 1H NMR of acetylated and deprotected compounds were investigated and the conformational preferences of the C6 side chain determined using molecular modeling.  相似文献   

20.
Positive-ion fast atom bombardment mass spectrometry appears to be a useful method for the differentiation of anomeric C-glycosides. The mass-analysed ion kinetic energy (MIKE) and collision-activated dissociation (CAD) MIKE spectra of selected positive ions can be used as fingerprints of the α- or β-anomers. The main fragmentation routes and particularly the formation of the [M ? H]+ ion and the [M + H ? PhCH2OH]+ ion were traced for each anomer.  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