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 859 毫秒
1.
Dehydration of d-altro-2-heptulose phenylosazone with methanolic sulfuric acid afforded two 3,6-anhydro-osazone derivatives (2 and 3). Compound 3 was obtained as the preponderant isomer, with inversion at C-1 (C-3 of the starting osazone), and 2 was obtained without inversion. Refluxing of 3 with copper sulfate afforded the C-nucleoside analog, namely, 2-phenyl-4-β-d-ribofuranosyl-1,2,3-osotriazole (4). Acetylation of 4 afforded the tri-O-acetyl derivative 5. The anomeric configuration was determined by c.d. and n.m.r. spectroscopy. The mass spectra of compounds 25 are discussed.  相似文献   

2.
Dehydration of 7-deoxy-l-manno-2-heptulose phenylosazone with methanolic sulfuric acid afforded two 3,6-anhydro-osazone derivatives (2 and 3). Refluxing the anhydro-osazones with copper sulfate gave two C-nucleoside analogs, namely, 4-(5-deoxy-α-l-arabinofuranosyl)-2-phenyl-ν-triazole (4) and 4-(5-deoxy-β-l-arabinofuranosyl)-2-phenyl-ν-triazole (5). The structure and anomeric configurations of 2, 4, and 5 were determined by n.m.r. spectroscopy. The preponderant conformation of 4 and 5, and the mass spectra of 2, 4, and 5 are discussed.  相似文献   

3.
Dehydration of the 3-epimeric 2-hexulose phenylosazones l-xylo-hexulose phenylosazone and l-lyxo-hexulose phenylosazone afforded 3,6-anhydro-l-lyxo-2-hexulose phenylosazone (2) as the preponderant isomer from both. The identity of 2 was obtained by t.l.c., and by acetylation followed by comparison of the products. Acetylation of 2 with acetic anhydride-pyridine afforded the di-O-acetyl derivative 4, and further acetylation gave the N-acetyldi-O-acetyl derivative 5. Refluxing of 2 with copper sulfate afforded a C-nucleoside analog, namely, 2-phenyl-4-α-l-threofuranosyl-1,2,3-osotriazole (6). The anomeric configuration was determined by n.m.r. spectroscopy. The stereochemical course of the dehydration process and the mass spectra of compounds 2, 4, 5, and 6 are discussed.  相似文献   

4.
Dehydration of the hydroxyalkyl chain of 1-phenyl-3-(d-arabino-tetritol-1-yl)pyrazolo[3,4-b]quinoxaline gave the C-nucleoside 3-β-d-erythrofuranosyl-1-phenyl-pyrazolo[3,4-b]quinoxaline (2) in 82% yield. The structure, and the configuration at the anomeric carbon atom, of 2 were elucidated by periodate oxidation, c.d. and n.m.r. spectroscopy, and mass spectrometry. N.m.r.-spectral and c.d. studies revealed that, due to the large size of the heterocyclic base, compound 2 is formed by inversion in the configuration or C-1 of the side chain. The mechanism of the dehydrative cyclization with inversion is discussed.  相似文献   

5.
A synthesis of l-dendroketose (5) has been achieved by microbiological oxidation by Acetobacter suboxydans of the branched-chain alditol 2-C-(hydroxymethyl)-d-erythro-pentitol (4). Treatment of the oxidation product with acetone, copper(II) sulfate, and sulfuric acid afforded the two di-O-isopropylidene-l-dendroketose derivatives 6 and 7. Assignment of configuration at the branching carbon atom (C-4) and at the anomeric center in 6 and 7 was made on the basis of the carbon-13 magnetic resonance spectra of these derivatives.  相似文献   

6.
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   

7.
The reaction of sucrose with a combination of 2,2-dimethoxypropane, N,N-dimethylformamide, and toluene-p-sulphonic acid (reagent A) gave, after acetylation followed by chromatography, 1′,2:4,6-di-O-isopropylidenesucrose tetra-acetate (1) in 15% yield. The structure of 1 was determined on the basis of p.m.r. and mass spectrometry, and by chemical transformations. Treatment of 1 with aqueous acetic acid afforded sucrose 3,3′,4′,6′-tetra-acetate 2. Reacetalation of 2 using reagent A gave 1 in 80% yield. The p.m.r. spectrum of 2 confirmed the presence of hydroxyl groups at C-2 and C-4. The following sequence of reactions showed that the remaining two hydroxyl groups were located at C-6 and C-1′. Selective tritylation of 2 gave 1′,6-di-O-tritylsucrose 3,3′,4′,6′-tetra-acetate (3) as the minor, and 6-O-tritylsucrose 3,3′,4′,6′-tetra-acetate (4) as the major, product. When tritylation was carried out under forcing conditions, 2 gave 3 as the major product. Acetylation of 4 afforded 6-O-tritylsucrose hepta-acetate. Mesylation of 2 gave the tetramethanesulphonate 5, which afforded the 6-dcoxy-6-iodo derivative 6 on treatment with a refluxing solution of sodium iodide in butanone. Treatment of 3 with methanesulphonyl chloride in pyridine gave the disulphonate 7, which on detritylation followed by acetylation gave 2,4-di-O-methanesulphonylsucrose hexa-acetate (9). Treatment of 9 with sodium benzoate in hexamethylphosphoric triamide displaced the 4-sulphonate, with inversion of configuration, to give the galacto derivative 10.  相似文献   

8.
Acetylation of benzyl 6-deoxy-3,4O-isopropylidene-β-L-galactopyranoside gave benzyl 2-O-acetyl-6-deoxy-3,4-O-isopropylidene-β-L-galactopyranoside (1). Removal of the isopropylidene group afforded benzyl 2-O-acetyl-6-deoxy-β-L-galactopyranoside (2), which was converted into benzyl 2-O-acetyl-6-deoxy-3,4-di-O-(methyl-sulfonyl)-β-L-galactopyranoside (3). Benzyl 2,3-anhydro-6-deoxy-4-O-(methyl-sulfonyl)-β-L-gulopyranoside (4) was obtained from 3 by treatment with alkali. Reaction of 4 with sodium azide in N,N-dimethylformamide gave a mixture of two isomeric benzyl 2,4-diazido-2,4,6-trideoxy hexoses, the syrupy diazido derivative 5 and the crystalline benzyl 2,4-diazido-2,4,6-trideoxy-β-L-idopyranoside (6). Acetylation of 6 afforded a compound whose n.m.r. spectrum was completely first order and in agreement with the structure of benzyl 3-O-acetyl-2,4-diazido-2,4,6-trideoxy-β-L-idopyranoside (7). Lithium aluminium hydride reduction of 5, followed by acetylation, afforded a crystalline product (8), shown by n.m.r. spectroscopy to be benzyl 2,4-diacetamido-3-O-acetyl-2,4,6-trideoxy-β-L-altropyranoside. Similar treatment of the diazido derivative 6 afforded benzyl 2,4-diacetamido-3-O-acetyl-2,4,6-trideoxy-β-L-idopyranoside (9). Compounds 8 and 9 could also be obtained from 4 by treatment of the crude diazido mixture with lithium aluminium hydride, with subsequent N-acetylation. The syrupy benzyl 2,4-diacetamido-2,4,6-trideoxy-β-L-altropyranoside (10) and the crystalline benzyl 2,4-diacetamido-2,4,6-trideoxy-β-L-idopyranoside (11) thus obtained were then O-acetylated to give 8 and 9 respectively. Benzyl 2,4-diacetamido-2,4,6-trideoxy-β-L-talopyranoside (15) was obtained from 11 by treatment with methanesulfonyl chloride and subsequent solvolysis. Compound 15 was O-acetylated to yield benzyl 2,4-diacetamido-3-O-acetyl-2,4,6-trideoxy-β-L-talopyranoside (16). the n.m.r. spectrum of which was in full agreement with the assigned structure. The mass spectra of compounds 8–11, 15, and 16 were also in agreement with their proposed structures. Removal of the benzyl groups from 10, 11 and 15 afforded the corresponding 2,4-diacetamido-2,4,6-trideoxyhexoses 12, 13, and 17, having the L-altro, L-ido, and L-talo configurations, respectively.  相似文献   

9.
Starting with cellobiosides, several different procedures were employed to prepare 6,6′-dichloro-6,6′-dideoxy, 6,6′-dibromo-6,6′-dideoxy, and 6,6′-dideoxy-6,6′-diiodo derivatives. Reduction with lithium aluminum hydride or nickel boride afforded peracetyl derivatives of methyl, phenyl, and benzyl 6-deoxy-4-O-(6-deoxy-β-D-glucopyranosyl)-β-D-glucopyranoside. Following acetolysis or hydrogenolysis, the glycosyl halide and the corresponding-glycal 40 were prepared. Iodomethoxylation of 40 and subsequent reduction gave the title compound. Alternatively, the halomethoxylation products of cellobial hexaacetate gave, by various procedures, the 2,6,6′-trideoxy-2,6,6′-trihalo derivatives, which, in turn, could be reduced to the title compound. The structures of the derivatives prepared were unequivocally assigned by n.m.r. spectroscopy. The various reaction sequences were compared with respect to the number of steps and the yields obtained.  相似文献   

10.
A general method for the preparation of 2′-azido-2′-deoxy- and 2′-amino-2′-deoxyarabinofuranosyl-adenine and -guanine nucleosides is described. Selective benzoylation of 3-azido-3-deoxy-1,2-O-isopropylidene-α-d-glucofuranose afforded 3-azido-6-O-benzoyl-3-deoxy-1,2-O-isopropylidene-α-d-glucofuranose (1). Acid hydrolysis of 1, followed by oxidation with sodium metaperiodate and hydrolysis by sodium hydrogencarbonate gave 2-azido-2-deoxy-5-O-benzoyl-d-arabinofuranose (3), which was acetylated to give 1,3-di-O-acetyl-2-azido-5-O-benzoyl-2-deoxy-d-arabinofuranose (4). Compound 4 was converted into the 1-chlorides 5 and 6, which were condensed with silylated derivatives of 6-chloropurine and 2-acetamido-hypoxanthine. The condensation reaction gave α and β anomers of both 7- and 9-substituted purine nucleosides. The structures of the nucleosides were determined by n.m.r. and u.v. spectroscopy, and by correlation of the c.d. spectra of the newly prepared nucleosides with those published for known purine nucleosides.  相似文献   

11.
Klebsiella K36 capsular polysaccharide has been investigated by methylation, Smith-periodate oxidation and partial hydrolysis techniques. The structure was found to consist of a hexasaccharide repeating unit as shown. The anomeric configurations of the sugar were determined by 1H and 13C n.m.r. spectroscopy on isolated oligomers obtained during the degradative studies and on the intact polysaccharide.  相似文献   

12.
The dilithio derivative of 2,4-di-O,N-trimethylsilylcytosine was condensed with 2,4:3,5-di-O-benzylidene-D-ribose to give a mixture of the protected, epimer at C-1′ pentitols 5 and 6; in addition, a compound substituted at N-3 or N-4, whose structure was not elucidated, was also obtained. The epimers were treated with acid to give 4-amino-2-hydroxy-5-(β-and α-D-ribofuranosyl)pyrimidine (10 and 12). The n.m.r. spectrum of 10 corresponds predominantly to the C-2endo structure. On the other hand, the n.m.r. spectrum of 12 presents couplings identical with those of the “α pseudo-uridine”. On nitric deamination, each isomer gave in a highly preponderant yield the corresponding pseudo-uridine at C-1′.  相似文献   

13.
Acid-catalyzed dehydration of methyl and ethyl 2-methyl-5-(d-arabino-tetrahydroxybutyl)-3-furoate (4a, b) takes place preferentially with inversion of configuration at C-1′, yielding the corresponding 5-(1,4-anhydro-d-ribo-tetrahydroxybutyl)-2-methyl-3-furoate (6a, b), and, to a much smaller extent, with retention of configuration giving the isomeric d-arabino anhydro-derivative (5a, b). The reaction is reversible, the equilibrium being set up when there is a high concentration of the thermodynamically more-stable d-ribo anhydro-derivative in the presence of the d-arabino isomer, the starting (d-arabino-tetrahydroxybutyl)furan (4a, db), and a compound thought to be methyl (or ethyl) 2-methyl-5-(d-ribo-tetrahydroxybutyl)-3-furoate (13). A mechanism is proposed for this reaction which involves the C-1′ carbonium ion 15 as the key intermediate. The anhydro derivatives of the d-ribo and d-arabino configurations can be distinguished by their optical rotations, the chemical shifts of H-1′, and the J1′,2′ coupling constants.  相似文献   

14.
The C-nucleoside analogs 6-chloro-3-β-d-erythrofuranosyl-l-phenylpyrazolo-[3,4-b]quinoxaline (5) and 3-β-d-erythrofuranosy]-l-p-tolylpyrazolo[3,4-b]quinoxaline (10) were prepared by dehydration of the polyhydroxyalkyl chain of 6(7)-chlorolo-phenyl-3-(d-arabino-tetritol-l-yl)-pyrazolo(3,4-b]quinoxaline and 3-(d-arbino-tetritol-l-yl)-l-p-tolylpyrazolo[3,4-b]quinoxaline, respectively. The structure and anomeric configuration of 5 and 10 were determined by high-resolution, n.m.r. spectroscopy. The mass spectra and biological activities of some of these compounds are discussed.  相似文献   

15.
Methanolysis of benzyl α-glycosides of N-acetylmuramic acid lactones with HO-6 free (2) and substituted (4, 7, 10, and 12) is catalysed by small amounts of silica gel to give, exclusively, the corresponding methyl esters with HO-4 unsubstituted (3, 5, 8, 11, 13); opening of the lactone ring proceeds with retention of the d-gluco configuration and can be followed by 1H-n.m.r. spectroscopy. Condensation of 2 with 2-methyl-(3,4,6-tri-O-acetyl-1,2-dideoxy-α-d-glucopyrano)-[2,1-d]-2-oxazoline (15) gave the β-(1→6)-linked disaccharide lactone 16 which, on methanolysis, yielded the disaccharide methyl ester 17, also obtained by condensation of 3 and 15. In the presence of imidazole, the lactones 2 and 4 underwent aminolysis with amino acid and peptide esters as nucleophiles to give the N-acetylmuramoylamide derivatives 19–24. The structures of methanolysis and aminolysis products were established by 1H-n.m.r. spectroscopy and independent syntheses.  相似文献   

16.
A new stereoselective preparation of N-aceyl-d-galactosamine (1b) starting from the known p-methoxyphenyl 3,4-O-isopropylidene-6-O-(1-methoxy-1-methylethyl)-β-d-galactopyranoside (10) is described using a simple strategy based on (a) epimerization at C-2 of 10 via oxidation-reduction to give the talo derivative 11, (b) amination with configurational inversion at C-2 of 11 via a SN2-type reaction on its 2-imidazylate, (c) anomeric deprotection of the p-methoxyphenyl β-d-galactosamine glycoside 14, (d) complete deprotection. Applying the same protocol to 2,3:5,6:3′,4′-tri-O-isopropylidene-6′-O-(1-methoxy-1-methylethyl)-lactose dimethyl acetal (4), directly obtained through acetonation of lactose, the disaccharide β-d-GalNAcp-(1→4)-d-Glcp (1a) was obtained with complete stereoselectivity in good (40%) overall yield from lactose.  相似文献   

17.
A conformational analysis of 2,3,4-tri-O-acetyl-D-xylono-1,5-lactone (5) has been performed by using 1H-n.m.r. spectral data. Evidence is presented that the C-3 and C-4 acetoxyl groups are anti-periplanar. The possible contribution of attractive 1,3- and 1,4-interactions between the electropositive lactone-ring oxygen and the endo-acetoxyl groups at C-3 and C-4 to the conformational stability of 5 is discussed.  相似文献   

18.
Controlled reaction of L-threo-2,3-hexodiulosono-1,4-lactone with substituted phenylhydrazines gave the 2-(monoarylhydrazones) (2), which underwent dehydrative acetylation to 4-(2-acetoxyethylidene)-4-hydroxy-2,3-dioxohutyro-1,4-lactone 2-(2-arylhydrazones) (3). The latter reacted with methylhydrazine to give 1-methyl-3-(1-methylpyrazolin-3-yl)-4,5-pyrazoledione 4-(2-arylhydrazones) (4). Reaction of the monoarythydrazones (2) with phenylhydrazine gave the mixed bishydrazones (5), which were rearranged by alkali and acidification to the pyrazolediones (6). Compounds 6 gave triacetyl (7) and tribenzoyl derivatives (8), and, on periodite oxidation, the aldehydes (9), which afforded the monohydrazones (10). The i.r.. n.m.r.. and mass-spectral data of some of the compounds were investigated.  相似文献   

19.
9-α-D-Mannofuranosyladenine (1) was quantitatively phosphorylated at O-5 by phosphoryl chloride in the presence of triethyl phosphate, giving phosphate 2. Treatment of 9-(2,3-O-isopropylidene-α-D-mannofuranosyl)adenine (3) with phosphoryl chloride-trimethyl phosphate, followed by hydrolysis at pH 1.5 to remove the protecting group, yielded mononucleotides 2 and 4 having the phosphate group at C-5′ and C-6′, respectively. These mononucleotides, chromatographically homogeneous in six solvent systems, were further characterized by their patterns of chromatography on Dowex ion-exchange resin, by their mass spectra, and by phosphorus n.m.r. spectroscopy. Both the 5′- and 6′-phosphates are noncompetitive inhibitors of adenylate kinase (for which a sensitive, accurate, and inexpensive, assay-system was developed). Of the two, the 6′-mononucleotide was the more potent inhibitor of adenylate kinase.  相似文献   

20.
《Carbohydrate research》1987,165(1):1-16
A 300-MHz 1H-n.m.r. study of methyl 2,3,4-tri-O-methyl-ga- (1) and β-d-galactopyranoside-6-(dimethyl phosphate) (3), using various solvents, shows that the gauche (gg) rotamer populations about the C-5C-6 bond are are the same in all solvents, whereas those of the gauche(trans) (gt) and trans(gauche) (tg, O-5 and O-6 trans) rotamers are solvent dependent. The tg population increases with decreasing polarity of the solvent, which is attributed to an increased electrostatic repulsion between O-5 and O-6 in apolar solvents. The tg population of 3 is larger than that of 1 and the same difference is observed in the corresponding compounds (2 and 4) which have a trigonal-bipyramidal five-coördinated phosphorus (Pv) at position 6 and which have a higher electron density at O-6. These differences in rotamer populations are due to an effect additional to that of the coulombic effect between O-5 and O-6. That these differences are caused by a combination of the gauche and anomeric effects is supported by the finding that the tg population increases with increasing pKa of the group at C-1. The results of the n.m.r. measurements (in CCl4) are reproduced fairly accurately by MNDO calculations on model systems. The solvent dependence of the rotamer population around the C-5ẋC-6 bond is a good criterion for the assignment of the H-6S,6R resonances since, for galactopyranosides, J5,6S increases and J5,6R decreases as the polarity of the solvent decreases.  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