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.
Three new glycosides, (3β,5α,8α,11α,12β,14β,17α,20R)‐3‐[(2,6‐dideoxy‐4‐O‐(6‐deoxy‐3‐O‐methyl‐β‐D ‐allopyranosyl)‐3‐O‐methyl‐β‐D ‐arabino‐hexopyranosyl)oxy]‐12‐O‐tigloyl‐8,20 : 11,20‐diepoxypregnane‐12,14‐diol ( 1 ), (3β,5α,8α,11α,12β,14β,17α,20R)‐3‐[(2,6‐dideoxy‐4‐O‐(6‐deoxy‐3‐O‐methyl‐β‐D ‐ allopyranosyl)‐3‐O‐methyl‐β‐D ‐arabino‐hexopyranosyl)oxy]‐12‐O‐(2‐methylbutanoyl)‐8,20 : 11,20‐diepoxypregnane‐12,14‐diol ( 2 ), and (3β,5α,11α,12β,14β,17α)‐12‐acetoxy‐3‐[(2,6‐dideoxy‐4‐O‐(6‐deoxy‐3‐O‐methyl‐β‐D ‐allopyranosyl)‐3‐O‐methyl‐β‐D ‐arabino‐hexopyranosyl)oxy]‐20‐oxo‐8,14‐epoxypregnan‐ 11‐yl isobutyrate ( 3 ) were isolated from the stems of Marsdenia tenacissima. The structures of the new compounds were elucidated by means of spectral data, including HR‐ESI‐MS, and 1D‐ and 2D‐NMR.  相似文献   

2.
The Oshima? Nozaki (Et2AlI) condensation of isolevoglucosenone ( 4 ) with 2,6‐anhydro‐3,4,5,7‐tetra‐O‐benzyl‐D ‐glycero‐D ‐gulo‐heptose ( 5 ) gave an enone 6 that was converted with high stereoselectivity to 3‐C‐[(1R)‐2,6‐anhydro‐D ‐glycero‐D ‐gulo‐heptitol‐1‐C‐yl]‐2,3‐dideoxy‐D ‐arabino‐hexose ( 1 ; 1 : 1 mixture of α‐ and β‐D ‐pyranose), and to 3‐C‐[(1R)‐2,6‐anhydro‐D ‐glycero‐D ‐gulo‐heptitol‐1‐C‐yl]‐2,3‐dideoxy‐D ‐lyxo‐hexose ( 2 ; 2.7 : 1.4 : 1.0 : 1.4 mixture of α‐D ‐furanose, β‐D ‐furanose, α‐D ‐pyranose, and β‐D ‐pyranose). The Oshima? Nozaki (Et2AlI) condensation of levoglucosenone ( 17 ) with aldehyde 5 gave an enone 18 that was converted with high stereoselectivity to 3‐C‐[(1R)‐2,6‐anhydro‐D ‐glycero‐D ‐gulo‐heptitol‐1‐C‐yl]‐3,4‐dideoxy‐α‐D ‐arabino‐hexopyranose ( 3 ; single anomer).  相似文献   

3.
Convergent syntheses of the 9‐(3‐X‐2,3‐dideoxy‐2‐fluoro‐β‐D ‐ribofuranosyl)adenines 5 (X=N3) and 7 (X=NH2), as well as of their respective α‐anomers 6 and 8 , are described, using methyl 2‐azido‐5‐O‐benzoyl‐2,3‐dideoxy‐2‐fluoro‐β‐D ‐ribofuranoside ( 4 ) as glycosylating agent. Methyl 5‐O‐benzoyl‐2,3‐dideoxy‐2,3‐difluoro‐β‐D ‐ribofuranoside ( 12 ) was prepared starting from two precursors, and coupled with silylated N6‐benzoyladenine to afford, after deprotection, 2′,3′‐dideoxy‐2′,3′‐difluoroadenosine ( 13 ). Condensation of 1‐O‐acetyl‐3,5‐di‐O‐benzoyl‐2‐deoxy‐2‐fluoro‐β‐D ‐ribofuranose ( 14 ) with silylated N2‐palmitoylguanine gave, after chromatographic separation and deacylation, the N7β‐anomer 17 as the main product, along with 2′‐deoxy‐2′‐fluoroguanosine ( 15 ) and its N9α‐anomer 16 in a ratio of ca. 42 : 24 : 10. An in‐depth conformational analysis of a number of 2,3‐dideoxy‐2‐fluoro‐3‐X‐D ‐ribofuranosides (X=F, N3, NH2, H) as well as of purine and pyrimidine 2‐deoxy‐2‐fluoro‐D ‐ribofuranosyl nucleosides was performed using the PSEUROT (version 6.3) software in combination with NMR studies.  相似文献   

4.
Nucleobase‐anion glycosylation of 2‐[(2‐methyl‐1‐oxopropyl)amino]imidazo[1,2‐a]‐1,3,5‐triazin‐4(8H)‐one ( 6 ) with 3,5‐di‐O‐benzoyl‐2‐deoxy‐2‐fluoro‐α‐D ‐arabinofuranosyl bromide ( 8 ) furnishes a mixture of the benzoyl‐protected anomeric 2‐amino‐8‐(2‐deoxy‐2‐fluoro‐D ‐arabinofuranosyl)imidazo[1,2‐a]‐1,3,5‐triazin‐4(8H)‐ones 9 / 10 in a ratio of ca. 1 : 1. After deprotection, the inseparable anomeric mixture 3 / 4 was silylated. The obtained 5‐O‐[(1,1‐dimethylethyl)diphenylsilyl] derivatives 11 and 12 were separated and desilylated affording the nucleoside 3 and its α‐D anomer 4 . Similar to 2′‐deoxy‐2′‐fluoroarabinoguanosine, the conformation of the sugar moiety is shifted from S towards N by the fluoro substituent in arabino configuration.  相似文献   

5.
The 2‐amino‐2‐deoxy‐α‐D ‐glucopyranosyl moiety (ring I) of paromomycin was replaced by a 2,4‐diamino‐2,4‐dideoxy‐α‐D ‐glucopyranosyl, 2,4‐diamino‐2,4‐dideoxy‐α‐D ‐galactopyranosyl, 2‐amino‐2‐deoxy‐α‐D ‐galactopyranosyl, or 3,4,5‐trideoxy‐4‐aza‐α‐D ‐erythro‐heptoseptanosyl moiety to investigate the effect of the substituent at C(4′) on the interaction with ribosomal RNA. The triflate 6 was prepared from the key intermediate pentaazido 3′,6′‐dibenzyl ether 5 , and the hexosulose 10 was obtained by oxidation of 5 with DessMartin's periodinane. Stereoselective reduction of 10 with NaBH4 gave the alcohol 11 that was transformed into the triflate 12 . The epimeric hexaazides 7 and 13 were obtained by treating the triflates 6 and 12 , respectively, with tetrabutylammonium azide. Periodate cleavage of glycol 2 yielded the dialdehyde 24 that was reductively aminated with aniline and benzylamine to give the 3,4,5‐trideoxy‐4‐aza‐α‐D ‐erythro‐heptoseptanosides 25 and 26 , respectively. Standard azide reduction and debenzylation yielded 9 (2,4‐diamino‐2,4‐dideoxy‐α‐D ‐galactopyranosyl ring I), 13 (2‐amino‐2‐deoxy‐α‐D ‐galactopyranosyl ring I), 17 (2,4‐diamino‐2,4‐dideoxy‐α‐D ‐glucopyranosyl ring I), and 27 and 28 (3,4,5‐trideoxy‐4‐aza‐α‐D ‐erythro‐heptoseptanosyl ring I). The derivatives 9 and 13 possessing a D ‐galacto‐configured ring I were less active than the corresponding D ‐gluco‐analogues 17 and paromomycin ( 1 ), respectively. The C(4′)‐aminodeoxy derivative 17 (D ‐gluco ring I) and the known 4′‐deoxyparomomycin ( 23 ), prepared by a new route, displayed slightly lower antibacterial activities than paromomycin ( 1 ). Cell‐wall permeability is not responsible for the unexpectedly low activity for 17 , as shown by cell‐free translation assays. The results evidence that the orientation of the substituent at C(4′) is more important than its nature for drug binding and activity.  相似文献   

6.
Enantiomerically pure (+)‐(1S,4S,5S,6S)‐6‐endo‐(benzyloxy)‐5‐exo‐{[(tert‐butyl)dimethylsilyl]oxy}‐7‐oxabicyclo[2.2.1]heptan‐2‐one ((+)‐ 5 ) and its enantiomer (−)‐ 5 , obtained readily from the Diels‐Alder addition of furan to 1‐cyanovinyl acetate, can be converted with high stereoselectivity into 8‐oxabicyclo[3.2.1]octane‐2,3,4,6,7‐pentol derivatives (see 23 – 28 in Scheme 2). A precursor of them, (1R,2S,4R,5S,6S,7R,8R)‐7‐endo‐(benzyloxy)‐8‐exo‐hydroxy‐3,9‐dioxatricyclo[4.2.1.02,4]non‐5‐endo‐yl benzoate ((−)‐ 19 ), is transformed into (1R,2R,5S, 6S,7R,8S)‐6‐exo,8‐endo‐bis(acetyloxy)‐2‐endo‐(benzyloxy)‐4‐oxo‐3,9‐dioxabicyclo[3.3.1]non‐7‐endo‐yl benzoate ((−)‐ 43 ) (see Scheme 5). The latter is the precursor of several protected 2,6‐anhydrohepturonic acid derivatives such as the diethyl dithioacetal (−)‐ 57 of methyl 3,5‐di‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐D ‐glycero‐D ‐galacto‐hepturonate (see Schemes 7 and 8). Hydrolysis of (−)‐ 57 provides methyl 3,5‐di‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐D ‐glycero‐D ‐galacto‐hepturonate 48 that undergoes highly diastereoselective Nozaki‐Oshima condensation with the aluminium enolate resulting from the conjugate addition of Me2AlSPh to (1S,5S,6S,7S)‐7‐endo‐(benzyloxy)‐6‐exo‐{[(tert‐butyl)dimethylsilyl]oxy}‐8‐oxabicyclo[3.2.1]oct‐3‐en‐2‐one ((−)‐ 13 ) derived from (+)‐ 5 (Scheme 12). This generates a β‐C‐mannopyranoside, i.e., methyl (7S)‐3,5‐di‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐7‐C‐[(1R,2S,3R,4S,5R,6S,7R)‐6‐endo‐(benzyloxy)‐7‐exo‐{[(tert‐butyl)dimethylsilyl]oxy}‐4‐endo‐hydroxy‐2‐exo‐(phenylthio)‐8‐oxabicyclo[3.2.1]oct‐3‐endo‐yl]‐L ‐glycero‐D ‐manno‐heptonate ((−)‐ 70 ; see Scheme 12), that is converted into the diethyl dithioacetal (−)‐ 75 of methyl 3‐O‐acetyl‐2,6‐anhydro‐4,5‐dideoxy‐4‐C‐{[methyl (7S)‐3,5,7‐tri‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐L ‐glycero‐D ‐manno‐heptonate]‐7‐C‐yl}‐5‐C‐(phenylsulfonyl)‐L ‐glycero‐D ‐galacto‐hepturonate ( 76 ; see Scheme 13). Repeating the Nozaki‐Oshima condensation to enone (−)‐ 13 and the aldehyde resulting from hydrolysis of (−)‐ 75 , a (1→3)‐C,C‐linked trisaccharide precursor (−)‐ 77 is obtained.  相似文献   

7.
Three new acacic acid derivatives, named coriariosides C, D, and E ( 1–3 ) were isolated from the roots of Albizia coriaria. Their structures were elucidated on the basis of extensive 1D‐ and 2D‐NMR studies and mass spectrometry as 3‐O‐[β‐D ‐xylopyranosyl‐(1 → 2)‐β‐D ‐fucopyranosyl‐(1 → 6)‐2‐(acetamido)‐2‐deoxy‐β‐D ‐glucopyranosyl]‐21‐O‐{(2E,6S)‐6‐O‐{4‐O‐[(2E,6S)‐2,6‐dimethyl‐ 6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐β‐D ‐quinovopyranosyl}‐2,6‐dimethylocta‐2,7‐dienoyl}acacic acid 28‐O‐β‐D ‐xylopyranosyl‐(1 → 4)‐α‐L ‐rhamnopyranosyl‐(1 → 2)‐β‐D ‐glucopyranosyl ester ( 1 ), 3‐O‐{β‐D ‐fucopyranosyl‐(1 → 6)‐[β‐D ‐glucopyranosyl‐(1 → 2)]‐β‐D ‐glucopyranosyl}‐21‐O‐{(2E,6S)‐6‐O‐{4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl]‐β‐D ‐quinovopyranosyl}‐2,6‐dimethylocta‐2,7‐dienoyl}acacic acid 28‐O‐α‐L ‐rhamno pyranosyl‐(1 → 2)‐β‐D ‐glucopyranosyl ester ( 2 ), and 3‐O‐[β‐D ‐fucopyranosyl‐(1 → 6)‐β‐D ‐glucopyranosyl]‐21‐O‐{(2E,6S)‐6‐O‐{4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐O‐(β‐D ‐quinovopyranosyl)octa‐2,7‐dienoyl)‐β‐D ‐quinovopyranosyl]octa‐2,7‐dienoyl}acacic acid 28‐O‐β‐D ‐glucopyranosyl ester ( 3 ). Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

8.
Four new triterpenoidal saponins acylated with monoterpenic acid, i.e., adianthifoliosides C, D, E, and F ( 1 – 4 ), besides the two known julibroside III and the monodesmonoterpenyl elliptoside A, were isolated from the roots of Albizia adianthifolia. Their structures were elucidated on the basis of extensive 1D‐ and 2D‐NMR studies and mass spectrometry as 3‐O‐{Oα‐L ‐arabinopyranosyl‐(1→2)‐Oβ‐d‐ fucopyranosyl‐(1→6)‐O‐[β‐d‐ glucopyranosyl‐(1→2)]‐β‐d‐ glucopyranosyl}‐21‐O‐{(2E,6S)‐6‐{{4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐(β‐D ‐quinovopyranosyloxy)octa‐2,7‐dienoyl]‐β‐d‐ quinovopyranosyl}oxy}‐2‐(hydroxymethyl)‐6‐methylocta‐2,7‐dienoyl}acacic acid 28‐{Oα‐L ‐arabinofuranosyl‐(1→4)‐O‐[β‐d‐ glucopyranosyl‐(1→3)]‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐d‐ glucopyranosyl} ester ( 1 ), 21‐O‐{(2E,6S)‐6‐{{4‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐(β‐d‐ quinovopyranosyloxy)octa‐2,7‐dienoyl]‐β‐d‐ quinovopyranosyl}oxy}‐2‐(hydroxymethyl)‐6‐methylocta‐2,7‐dienoyl}‐3‐O‐{Oβ‐D ‐xylopyranosyl‐(1→2)‐Oβ‐d‐ fucopyranosyl‐(1→6)‐2‐(acetylamino)‐2‐deoxy‐β‐d‐ glucopyranosyl}acacic acid 28‐{Oα‐L ‐arabinofuranosyl‐(1→4)‐O‐[β‐d‐ glucopyranosyl‐(1→3)]‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐d‐ glucopyranosyl} ester ( 2 ), 21‐O‐{(2E,6S)‐6‐{{3‐O‐[(2E,6S)‐2,6‐dimethyl‐6‐(β‐d‐ quinovopyranosyloxy)octa‐2,7‐dienoyl]‐β‐d‐ quinovopyranosyl}oxy}‐2,6‐dimethylocta‐2,7‐dienoyl}‐3‐O‐{Oβ‐D ‐xylopyranosyl‐(1→2)‐Oβ‐d‐ fucopyranosyl‐(1→6)‐2‐(acetylamino)‐2‐deoxy‐β‐d‐ glucopyranosyl}acacic acid 28‐{Oα‐L ‐arabinofuranosyl‐(1→4)‐O‐[β‐d‐ glucopyranosyl‐(1→3)]‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐d‐ glucopyranosyl} ester ( 3 ), and 3‐O‐{Oα‐L ‐arabinopyranosyl‐(1→2)‐Oβ‐d‐ fucopyranosyl‐(1→6)‐O‐[β‐d‐ glucopyranosyl‐(1→2)]‐β‐d‐ glucopyranosyl}‐21‐O‐{(2E,6S)‐2,6‐dimethyl‐6‐(β‐d‐ quinovopyranosyloxy)octa‐2,7‐dienoyl}acacic acid 28‐{Oα‐L ‐arabinofuranosyl‐(1→4)‐O‐[β‐d‐ glucopyranosyl‐(1→3)]‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐d‐ glucopyranosyl} ester ( 4 ).  相似文献   

9.
Practical syntheses of 2‐keto‐3‐deoxy‐D ‐xylonate (D ‐KDX) and 2‐keto‐3‐deoxy‐L ‐arabinonate (L ‐KDA) that rely on reaction of the anion of ethyl 2‐[(tert‐butyldimethylsilyl)oxy]‐2‐(dimethoxy phosphoryl) acetate with enantiopure glyceraldehyde acetonide, followed by global deprotection of the resultant O‐silyl‐enol esters, have been developed. This has enabled us to confirm that a 2‐keto‐3‐deoxy‐D ‐gluconate aldolase from the archaeon Sulfolobus solfataricus demonstrates good activity for catalysis of the retro‐aldol cleavage of both these enantiomers to afford pyruvate and glycolaldehyde. The stereochemical promiscuity of this aldolase towards these enantiomeric aldol substrates confirms that this organism employs a metabolically promiscuous pathway to catabolise the C5‐sugars D ‐xylose and L ‐arabinose.  相似文献   

10.
Two novel echinocystic acid (=(3β,16α)‐3,16‐dihydroxyolean‐12‐en‐28‐oic acid) glycosides, foetidissimosides C ( 1 ), and D ( 2 ), along with new cucurbitane glycosides, i.e., foetidissimosides E/F ( 3 / 4 ) as an 1 : 1 mixture of the (24R)/(24S) epimers, were obtained from the roots of Cucurbita foetidissima. Their structures were elucidated by means of a combination of homo‐ and heteronuclear 2D‐NMR techniques (COSY, TOCSY, NOESY, ROESY, HSQC, and HMBC), and by FAB‐MS. The new compounds were characterized as (3β,16α)‐28‐{[Oβ‐D ‐glucopyranosyl‐(1→3)‐Oβ‐D ‐xylopyranosyl‐(1→4)‐O‐6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl]oxy}‐16‐hydroxy‐28‐oxoolean ‐12‐en‐3‐yl β‐D ‐glucopyranosiduronic acid ( 1 ), (3β,16α)‐16‐hydroxy‐28‐oxo‐28‐{{Oβ‐D ‐xylopyranosyl‐(1→3)‐O‐[β‐D ‐xylopyranosyl‐(1→4)]‐O‐6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl}oxy}olean‐12‐en‐3‐yl β‐D ‐glucopyranosiduronic acid ( 2 ), and (3β,9β,10α,11α,24R)‐ and (3β,9β,10α,11α,24S)‐25‐(β‐D ‐glucopyranosyloxy)‐9‐methyl‐19‐norlanost‐5‐en‐3‐yl 2‐Oβ‐D ‐glucopyranosyl‐β‐D ‐glucopyranoside ( 3 and 4 , resp.).  相似文献   

11.
Starting from methyl 2,3‐O‐isopropylidene‐α‐D ‐mannofuranoside ( 5 ), methyl 6‐O‐benzyl‐2,3‐O‐isopropylidene‐α‐D ‐lyxo‐hexofuranosid‐5‐ulose ( 12 ) was prepared in three steps. The addition reaction of dimethyl phosphonate to 12 , followed by deoxygenation of 5‐OH group, provided the 5‐deoxy‐5‐dimethoxyphosphinyl‐α‐D ‐mannofuranoside derivative 15a and the β‐L ‐gulofuranoside isomer 15b . Reduction of 15a and 15b with sodium dihydrobis(2‐methoxyethoxy)aluminate, followed by the action of HCl and then H2O2, afforded the D ‐mannopyranose ( 17 ) and L ‐gulopyranose analog 21 , each having a phosphinyl group in the hemiacetal ring. These were converted to the corresponding 1,2,3,4,6‐penta‐O‐acetyl‐5‐methoxyphosphinyl derivatives 19 and 23 , respectively, structures and conformations (4C1 or 1C4, resp.) of which were established by 1H‐NMR spectroscopy.  相似文献   

12.
Phytochemical analyses were carried out on the rhizomes of Clintonia udensis (Liliaceae) with particular attention paid to the steroidal glycoside constituents, resulting in the isolation of three new polyhydroxylated spirostanol glycosides, named clintonioside A ( 1 ), B ( 2 ), and C ( 3 ). On the basis of their spectroscopic data, including 2D‐NMR spectroscopy, in combination with acetylation and hydrolytic cleavage, the structures of 1 – 3 were determined to be (1β,3β,23S,24S,25R)‐1,23,24‐trihydroxyspirost‐5‐en‐3‐yl Oβ‐D ‐glucopyranosyl‐(1→4)‐O‐[α‐L ‐rhamnopyranosyl‐(1→2)]‐β‐D ‐glucopyranoside ( 1 ), (1β,3β,23S,24S)‐3,21,23,24‐tetrahydroxyspirosta‐5,25(27)‐dien‐1‐yl Oα‐L ‐rhamnopyranosyl‐(1→2)‐O‐[β‐D ‐xylopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 2 ), and (1β,3β,23S,24S)‐21‐(acetyloxy)‐24‐[(6‐deoxy‐β‐D ‐gulopyranosyl)oxy]‐3,23‐dihydroxyspirosta‐5,25(27)‐dien‐1‐yl Oα‐L ‐rhamnopyranosyl‐(1→2)‐O‐[β‐D ‐xylopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 3 ).  相似文献   

13.
The title disaccharide, 2‐{2‐{2‐[(2‐ethoxy‐3,4‐dioxocyclobut‐1‐en‐1‐yl)amino]ethoxy}ethoxy}ethyl 2‐O‐(3,6‐dideoxy‐α‐L ‐xylo‐hexopyranosyl)‐β‐d‐ galactopyranoside cyclic 4,6‐(potassium phosphate) ( 2 ), was synthesized from the two isomeric linker‐equipped galactose acceptors 9 and 10 , obtained by phosphorylation of 2‐[2‐(2‐azidoethoxy)ethoxy]ethyl 3‐O‐benzyl‐β‐d‐ galactopyranoside ( 8 ), which were glycosylated with ethyl 2,4‐di‐O‐benzyl‐3,6‐dideoxy‐1‐thio‐β‐l‐ xylo‐hexopyranoside ( 12 ; Scheme). Mainly the fully protected α‐(1 → 2)‐linked products 13 α and 14 α were formed. Catalytic hydrogenolysis/hydrogenation effected global deprotection, thereby removing the chirality at the P‐atom, and simultaneously converted the azido group in the linker to an amino group (→ 15 ). Final treatment with diethyl squarate (= 3,4‐diethoxycyclobut‐3‐ene‐1,2‐dione) gave target compound 2 , amenable for conjugation to proteins.  相似文献   

14.
The 1,3,4,6‐tetra‐O‐acetyl‐2‐azido‐2‐deoxy‐β‐D ‐mannopyranose ( 4 ) or the mixture of 1,3,6‐tri‐O‐acetyl‐2‐azido‐2‐deoxy‐4‐O‐(2,3,4,6‐tetra‐O‐acetyl‐β‐D ‐galactopyranosyl)‐β‐D ‐mannopyranose ( 10 ) and the corresponding α‐D ‐glucopyranose‐type glycosyl donor 9 / 10 reacted at room temperature with protected nucleosides 12 – 15 in CH2Cl2 solution in the presence of BF3?OEt2 as promoter to give 5′‐O‐(2‐azido‐2‐deoxy‐α‐D ‐glycosyl)nucleosides in reasonable yields (Schemes 2 and 3). Only the 5′‐O‐(α‐D ‐mannopyranosyl)nucleosides were obtained. Compounds 21, 28, 30 , and 31 showed growth inhibition of HeLa cells and hepatoma Bel‐7402 cells at a concentration of 10 μM in vitro.  相似文献   

15.
Iodination of N2‐isobutyryl‐5‐aza‐7‐deazaguanine ( 7 ) with N‐iodosuccinimide (NIS) gave 7‐iodo‐N2‐isobutyryl‐5‐aza‐7‐deazaguanine ( 8 ) in a regioselective reaction (Scheme 1). Nucleobase‐anion glycosylation of 8 with 2‐deoxy‐3,5‐di‐O‐toluoyl‐α‐D ‐ or α‐L ‐erythro‐pentofuranosyl chloride furnished anomeric mixtures of D ‐ and L ‐nucleosides. The anomeric D ‐nucleosides were separated by crystallization to give the α‐D ‐anomer and β‐D ‐anomer with excellent optical purity. Deprotection gave the 7‐iodo‐5‐aza‐7‐deazaguanine 2′‐deoxyribonucleosides 3 (β‐D ; ≥99% de) and 4 (α‐D ; ≥99% de). The reaction sequence performed with the D ‐series was also applied to L ‐nucleosides to furnish compounds 5 (β‐L ; ≥99% de) and 6 (α‐L ; ≥95% de).  相似文献   

16.
Four new flavonoid glycosides, 3‐O‐[α‐L ‐arabinopyranosyl‐(1→2)‐β‐D ‐galactopyranosyl]‐7‐Oβ‐D ‐glucopyranosylkaempferol ( 1 ), 3‐O‐(α‐L ‐arabinopyranosyl‐(1→2)‐{4‐O‐[(E)‐caffeoyl]‐β‐D ‐galactopyranosyl})‐7‐Oβ‐D ‐glucopyranosylquercetin ( 2 ), 3‐O‐{2‐O‐[(E)‐caffeoyl]‐α‐L ‐arabinopyranosyl‐(1→2)‐β‐D ‐galactopyranosyl}‐7‐Oβ‐D ‐glucopyranosylkaemperfol ( 3 ), and 3‐O‐{2‐O‐[(E)‐caffeoyl]‐α‐L ‐arabinopyranosyl‐(1→2)‐β‐D ‐galactopyranosyl}kaempferol ( 4 ), together with two known compounds were isolated from the aerial parts of Ranunculus chinensis Bge . The structures of the new glycosides were determined on the basis of spectroscopic analysis, including 1D‐ and 2D‐NMR, and ESI‐MS techniques, and chemical methods.  相似文献   

17.
Two new lanostane‐type nonsulfated pentasaccharide triterpene glycosides, 17‐dehydroxyholothurinoside A ( 1 ) and griseaside A ( 2 ), were isolated from the sea cucumber Holothuria grisea. Their structures were elucidated by spectroscopic methods, including 2D‐NMR and MS experiments, as well as chemical evidence. Compounds 1 and 2 possess the same pentasaccharide moieties but differ slightly in their side chains of the holostane‐type triterpene aglycone. The structures of the two new glycosides were established as (3β,12α)‐22,25‐epoxy‐3‐{(Oβ‐D ‐glucopyranosyl‐(1→4)‐O‐[O‐3‐O‐methyl‐β‐D ‐glucopyranosyl‐(1→3)‐Oβ‐D ‐glucopyranosyl‐(1→4)‐6‐deoxy‐β‐D ‐glucopyranosyl‐(1→2)]‐β‐D ‐xylopyranosyl)oxy}‐12,20‐dihydroxylanost‐9(11)‐en‐18‐oic acid γ‐lactone ( 1 ) and (3β,12α)‐3‐{(Oβ‐D ‐glucopyranosyl‐(1→4)‐O‐[O‐3‐O‐methyl‐β‐D ‐glucopyranosyl‐(1→3)‐Oβ‐D ‐glucopyranosyl‐(1→4)‐6‐deoxy‐β‐D ‐glucopyranosyl‐(1→2)]‐β‐D ‐xylopyranosyl)oxy}‐12,20,22‐trihydroxylanost‐9(11)‐en‐18‐oic acid γ‐lactone ( 2 ). The 17‐dehydroxyholothurinoside A ( 1 ) and griseaside A ( 2 ) exhibited significant cytotoxicity against HL‐60, BEL‐7402, Molt‐4, and A‐549 cancer cell lines.  相似文献   

18.
Three new dammarane‐type triterpenoid saponins, 1 – 3 , were isolated and identified as (20S)‐20‐O‐[β‐D ‐xylopyranosyl‐(1→6)‐β‐D ‐glucopyranosyl‐(1→6)‐β‐D ‐glucopyranosyl]dammar‐24‐ene‐3β,6α,12β, 20‐tetrol ( 1 ), (20S)‐6‐O‐[(E)‐but‐2‐enoyl‐(1→6)‐β‐D ‐glucopyranosyl]dammar‐24‐ene‐3β,6α,12β,20‐tetrol ( 2 ), and (20S)‐6‐O‐[β‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐xylopyranosyl]dammar‐24‐ene‐3β,6α,12β,20‐tetrol ( 3 ) from the roots of Panax notoginseng (Burkill ) F.H.Chen (Araliaceae). Their structures were elucidated on the basis of spectroscopic analyses, including 1D‐ and 2D‐NMR techniques and HR‐ESI‐MS, as well as by acidic hydrolysis.  相似文献   

19.
Seventeen flavonoids, five of which are flavone C‐diosides, 1 – 5 , were isolated from the BuOH‐ and AcOEt‐soluble fractions of the leaf extract of Machilus konishii. Among 1 – 5 , apigenin 6‐Cβ‐D ‐xylopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 2 ), apigenin 8‐Cα‐L ‐arabinopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 4 ), and apigenin 8‐Cβ‐D ‐xylopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 5 ) are new. Both 4 and 5 are present as rotamer pairs. The structures of the new compounds were elucidated on the basis of NMR‐spectroscopic analyses and MS data. In addition, the 1H‐ and 13C‐NMR data of apigenin 6‐Cα‐L ‐arabinopyranosyl‐2″‐Oβ‐D ‐glucopyranoside ( 3 ) were assigned for the first time. The isolated compounds were assayed against α‐glucosidase (type IV from Bacillus stearothermophilus). Kaempferol 3‐O‐(2‐β‐D ‐apiofuranosyl)‐α‐L ‐rhamnopyranoside ( 12 ) was found to possess the best inhibitory activity with an IC50 value of 29.3 μM .  相似文献   

20.
The four new acylated triterpene saponins 1 – 4 , isolated as two pairs of isomers and named libericosides A1/A2 and B1/B2, one pair of isomers 5 / 6 , the (Z)‐isomer libericoside C2 ( 5 ) being new, one new sucrose ester, atroximoside ( 7 ), and eight known compounds were isolated from the roots of Atroxima liberica by repeated MPLC and VLC on normal and reversed‐phase silica gel. Their structures were elucidated on the basis of extensive 1D‐ and 2D‐NMR studies (1H‐ and 13C‐NMR, DEPT, COSY, TOCSY, NOESY, HSQC, and HMBC) and mass spectrometry as 3‐Oβ‐D ‐glucopyranosylpresenegenin 28‐{Oα‐L ‐arabinopyranosyl‐(1→3)‐Oβ‐D ‐xylopyranosyl‐(1→4)‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐4‐O‐[(E)‐3,4‐dimethoxycinnamoyl]‐β‐D ‐fucopyranosyl} ester ( 1 ) and its (Z)‐isomer 2 , 3‐Oβ‐D ‐glucopyranosylpresenegenin 28‐{Oα‐L ‐arabinopyranosyl‐(1→4)‐Oβ‐D ‐xylopyranosyl‐(1→4)‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐O‐[O‐β‐D ‐xylopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐(1→3)]‐4‐O‐[(E)‐3,4‐dimethoxycinnamoyl]‐β‐D ‐fucopyranosyl} ester ( 3 ) and its (Z)‐isomer 4 , 3‐Oβ‐D ‐glucopyranosylpresenegenin 28‐{Oβ‐D ‐xylopyranosyl‐(1→4)‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐O‐[6‐O‐acetyl‐β‐D ‐glucopyranosyl‐(1→3)]‐4‐O‐[(Z)‐3,4‐dimethoxycinnamoyl]‐β‐D ‐fucopyranosyl} ester ( 5 ), and 3‐O‐[(Z)‐feruloyl]‐β‐D ‐fructofuranosyl α‐D ‐glucopyranoside ( 7 ). Compounds 1 – 6 and the known saponins 8 / 9 were evaluated against the human colon cancer cells HCT 116 and HT‐29 and showed moderate to weak cytotoxicity.  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