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 piperidines 12 – 18 , piperidmose analogues of Neu5Ac ( 1 ) with a shortened side chain, were synthesized from N-acetyl-D -glucosamine via the azidoalkene 32 and tested as inhibitors of Vibrio cholerae sialidase. Deoxygenation at C(4) of the uronate 22 , obtained from the known D -GlcNAc derivative 20 , was effected by β-elimination (→ 23 ), exchange of the AcO at C(3) with a (t-Bu)Me2SiO group and hydrogenation (→ 26 ; Scheme 1). Chain extension of 26 by reaction with Me3SiCH2MgCl gave the D -ido-dihydroxysilane 28 , which was transformed into the unsaturated L -xylo-mesylate 29 and further into the L -lyxo-alcohol 30 , the mesylate 31 , and the L -xylo-azide 32 . The derivatives 29 – 31 prefer a sickle zig-zag and 32 mainly an extended zig-zag conformation (Fig. 2). The piperidinecarboxylate 15 was obtained from 32 by ozonolysis (→ 33 ), intramolecular reductive animation (→ 34 ), and deprotection, while reductive animation of 34 with glycolaldehyde (→ 35 ) and deprotection gave 16 (Scheme 2). An intramolecular azide-olefin cycloaddition of 32 yielded exclusively the fused dihydrotriazole 36 , while the lactone 39 did not cyclize (Scheme 3). Treatment of 36 with AcOH (→ 37 ) followed by hydrolysis (→ 38 ) and deprotection led to the amino acid 18 . To prepare the (hydroxymethyl)piperidinecarboxylates 12 and 17 , 32 was first dihydroxylated (Scheme 4). The L -gluco-diol 40 was obtained as the major product, in agreement with Kishi's rule. Silylation of 40 (→ 42 ), oxidation with periodinane (→ 44 ), and reductive animation gave the L -gluco-piperidine 45 . It was, on the one hand, deprotected to the amino acid 12 and, on the other hand, N-phenylated (→ 46 ) and deprotected to 17 . While 45 and 12 adopt a 2C5 conformation, the analogous N-Ph derivatives 46 and 17 adopt a 5C2 and a B3,6 conformation, respectively, on account of the allylic 1,3-strain. The conformational effects of this 1,3-strain are also evident in the carbamate 47 , obtained from 45 (Scheme 5), and in the C(2)-epimerized bicyclic ether 48 , which was formed upon treatment of 47 with (diethylamino)sulfur trifluoride (DAST). Fluorination of 40 with DAST (→ 49 ) followed by treatment with AcOH led to the D -ido-fluorohydrin 50 . Oxidation of 50 (→ 51 ) followed by a Staudinger reaction and reduction with NaBH3CN afforded the (fluoromethyl)piperidine 52 , while reductive amination of 51 with H2/Pd led to the methylpiperidine 55 , which was similarly obtained from the keto tosylate 54 and from the dihydrotriazole 36 . Deprotection of 52 and 55 gave the amino acids 13 and 14 , respectively. The aniline 17 does not inhibit V. cholerae sialidase; the piperidines 12 – 16 and 18 are weak inhibitors, evidencing the importance of an intact 1,2,3-trihydroxypropyl side chain.  相似文献   

2.
The 2,3-dihydro-1H-benz[f]indole-4,9-diones 3a–d , h were formed in a one-step reaction in 13–82% yield by an unprecedented [3 + 2] regioselective photoaddition of 2-amino-1,4-naphthoquinone ( 1 ) with various electronrich alkenes 2 (Scheme 1, Table). The [3 + 2] photoadducts derived from 1 with vinyl ethers and vinyl acetate gave 1H-benz[f]indole-4,9-diones 4e , f , i , in 33–72% yield, by spontaneous loss of the corresponding alcohol or AcOH from the resulting adducts; 4i has a kinamycin skeleton. The [3 + 2] photoaddition also took place on irradiation of the differently substituted amino-1,4-benzoquinones 6 , 7 , and 12 and excess alkenes 2 in benzene, giving 1H-indole-4,7-dione derivatives 13 and 14 (Scheme 3), 15a and 16 (Scheme 4), and 18 (Scheme 4), respectively. The initial products in these photoadditions were proved to be hydroquinones, the air oxidation of which yielded the heterocyclic quinones; 2,3-dihydro-2-methoxy-2-methyl-5-phenyl-1H-indole-1,4,7-triyl triacetate ( 19 ) was isolated after treatment of the crude photoaddition mixture obtained from 2-amino-5-phenyl-1,4-benzoquinone ( 7 ) and 2-methoxyprop-1-ene ( 2f ) with Ac2O and pyridine under N2. A pathway leading to the annelated hydroquinones involving ionic intermediates arising from an electron transfer in these photoadditions is proposed (Scheme 5).  相似文献   

3.
The pyrrolidine derivatives 3 , 4 , and 5 were prepared from the methyl ester 7 of Neu2en5Ac via lie pyrrolidine-borane adduct 33 . They inhibit Vibrio cholerae sialidase competitively with Ki = 4. 4 10?3 M, 5. 3 10?3 M, and 4. 0 10?2 M, respectively. Benzylation of 7 gave the fully O-benzylated 8 besides 9, 10 , and 11. Ozonolysis and reduction with NaBH4 of 8 and 9 gave the 1, 4-diols 12 and 15 , the hydroxy acetates 13 and 16 , and the furanoses 14 and 17 (Scheme 1), respectively. The diol 12 was selectively protected (→ 19 → 20 → 23 ) and transformed into the azide 27 by a Mitsunobu reaction. Selective base-catalysed deprotection of the diacetate 22 , obtained from 12 , was hampered by an easy acetyl-group migration. The mesylate 28 proved unstable. The azide 27 was transformed via 29 into the ketone 30 (Scheme 2). Hydrogenation of 30 gave the dihydropyrrole 31 and, hence, the pyrrole 32. The adduct 33 was obtained from 30 by a Staudinger reaction (→31) and reduction with LiBH4/HBF4. It was transformed into the pyrroudine 34 . The structure of 34 was established by X-ray analysis. Reductamination of the pyrrolidine-borane adduct with glyoxylic acid gave 40 and, hence, 3. N-Alkylation afforded 44 and, hence, the phosphonate 4. The acid 5 was obtained from 33 by acylation (→ 47 ) and deprotection (Scheme 4).  相似文献   

4.
Oxidation of 1,2-Thiazoles; A Convenient Approach to 1,2-Thiazol-3(2H)-one 1,1-Dioxides The 1,2-thiazoles obtained from 3-chloroalk-2-enals and ammonium thiocyanate ( 7 → 9 , Scheme 1) are easily transformed to 1,2-thiazol-3(2H)-one 1,1-dioxidcs 10 on treatment with H2O2 in AcOH at 80°. Hydrogenation of 10 in AcOH yields the corresponding saturated 1,2-thiazolidin-3-one 1,1-dioxides 16 (Scheme 3). Cycloalka[c]-1,2-thiazoles 18 are prepared from 2-[(thiocyanato)methyliden]cycloalkan-1-ones and ammonia (Scheme 4). Surprisingly, oxidation of 18a with H2O2 in AcOH yields the tricyclic oxaziridine 19.  相似文献   

5.
The synthesis of the phospha analogue 10 of DANA ( 2 ) is described. Bromo-hydroxylation of the known 11 (→ 12 and 13 ) followed by treatment of the major bromohydrin 13 with 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) gave the oxirane 14 (Scheme 1). Depending on the solvent, TiBr4 transformed 14 into 16 or into a 15 / 16 mixture. Reductive debromination of 16 (→ 17 ), followed by benzylation provided 18 . Oxidattve decarboxylation (Pb(OAc)4) of the acid, obtained by saponification of 18 , yielded the anomeric acetates 19 and 20 . While 19 was inert under the conditions of phosphonoylation, the more reactive imidate 22 , obtained together with 23 from 19 / 20 via 21 (Scheme 2), gave a mixture of the phosphonates 24 / 25 and the bicyclic acetal 26 . Debenzylation of 24 / 25 and acetylation led to the acetoxyphosphonates 27 / 28 . Since β-elimination of AcOH from 27 / 28 proved difficult, the bromide 34 was prepared from 27 / 28 by photobromination and subjected to reductive elimination with Zn/Cu (→ 35 ; Scheme 3). This two-step sequence was first investigated using the model compounds 30 and 31 . Transesterification of 35 , followed by deacetylation gave 10 , which is a strong inhibitor of the Vibrio Cholerae sialidase.  相似文献   

6.
The tetrazole 1 is a new β-glucosidase inhibitor (IC50=8·10?5 M , Emulsin), obtained (92%) by deprotection of 22 , the product of an intramolecular cycloaddition of the azidonitrile 20 . This azidonitrile was formed as an intermediate by treating the L -ido-bromide 14 or the L -ido-tosylate 19 with NaN3 at 110–120°. It was isolated in a separate experiment. The yield of 22 from 19 reached 70%; 21 was formed as by-product (10%). The bromide 14 (42%) and the iodide 15 (30–35%) were obtained from the nitrile 13 , together with the 2,5-anhydro-L -idononitrile 16, which was formed in ca. 35–45%. The tosylate 19 was obtained from 18 (97%). To obtain 18 , the nitrile 13 was oxidized according to Swern (→17, 92%) and then reduced (NaBH4, CeCl3), leading to 18 and 13 (92%, 18/13 93:7). Reduction of the tetrahydropyridotetrazole 22 with LiAlH4 afforded 83 % of the piperidine 23 , which was deprotected to (+)-1-deoxynojirimycin hydroacetate (2·AcOH, 86%) and further converted into the corresponding hydrochloride and into the free base 2 .  相似文献   

7.
The allylic and homoallylic alcohols 1 – 8 , prepared from (+)-camphor and (−)-fenchone, were ozonized in Et2O at −78° and treated with Et3N or LiAlH4 to give the chiral hydroxy carbonyl compounds 9 – 16 and the diols 17 – 24 , respectively (Scheme 1). In the case of the diols 19 and 24 , the formation of new chiral centers proceeded with high diastereoselectivity. These diols were prepared highly diastereoselectively also by LiAlH4 reduction of the hydroxy carbonyl compounds 11 and 16a , respectively (Scheme 2). The absolute configuration of the new chiral centers in 19 and 24 was determined by X-ray and NMR methods. The ozonization of compounds 2 , 3 , 7 , and 8 provided the relatively stable hydroxy-substituted 1,2,4-trioxolane derivatives (ozonides) 37 – 40 (Scheme 5) which were characterized by 1H- and 13C-NMR spectra, ESI-MS, and natural-abundance 17O-NMR spectra.  相似文献   

8.
The synthesis of 6-C-methyl-Neu2en5Ac ( 4 ), 6-C-(hydroxymethyl)-Neu2en5Ac ( 5 ), and 6-C-methyl-Neu5Ac ( 6 ) is described. The 4-methylumbellyferyl glycosides 8 and 9 were also prepared but proved unstable. Protection of the previously reported nitro ether 10 (→ 11 ) followed by a Kornblum reaction gave the branched-chain derivative 13 which was transformed into aldehyde 14 and hence via 16 into the-protected 6-C-hydroxymethylated 20 and into the 6-C-methyl-substituted 18 (Scheme 1). Debenzylidenation of 20 and 18 afforded the diols 21 and 19 , respectively. Selective oxydation of 19 followed by esterification (→ 22 ), acetylation (→ 23 ), and elimination led to the protected 6-C-methyl-Neu2en5Ac derivative 24 (Scheme 2). Bromomethoxylation yielded mainly 25 and some 26 , which were reductively debrominated to 27 and 28 , respectively. Attempted deprotection of 27 did not lead to the corresponding acid, but to the 2,7- and 2,8-anhydro compounds 29 and 30 which were characterised as their peracetylated esters 31 and 32 (Scheme 3). The structure of 32 was established by X-ray analysis. Oxydation of 19 and 21 , followed by deprotection, esterification, and acetylation gave 37 and 38 , respectively (Scheme 4). The branched-chain Neu2en5Ac derivatives 4 and 5 were obtained by β-elimination (→ 39 and 40 ) and deprotection. Omission of the esterification after oxydation of 33 and 34 gave the lactones 35 and 36 which were transformed into 37 and 38 , respectively. Bromoacetoxylation of 39 gave 41-43 which were reductively debrominated to 44 (from 41 and 42 ) and 45 (Scheme 5). Bromoacetoxylation of 40 yielded 46 which was debrominated to 47. Glycosidation of the glycosyl chlorides obtained from 44 and 47 led to the α -D-glycosides 48 and 49 and to the elimination products 39 and 40 , respectively (Scheme 6). Transesterification of 48 , followed by saponification gave the unstable glycoside 8 and hence 6-C-methyl-Neu5Ac ( 6 ). The unstable glycoside 9 was obtained by similar treatment of 49 but yielded 50 under acidic conditions. The branched-chain 4 and 5 were weak inhibitors of Vibrio cholera sialidase, and 8 and 9 were very poor substrates.  相似文献   

9.
Two cis-fused cyclopenteno-1,2,4-trioxanes, 1a and 1b , were subjected to Zn in AcOH or FeCl2 · 4H2O in MeCN. In the first case, the main course was deoxygenation to give cyclopentanone ( 18 ) and the 1,4-diphenyl- or 1,4-bis(4-fluorophenyl)cyclopent-3-ene-1,2-diol 10 (Scheme 5). In the second case, isomerization chiefly occurred resulting in the formation of a dimer 9 of the respective 3,5-diaryl-5-hydroxycyclopent-2-enyl 5-hydroxypentanoates 8 (Scheme 3).  相似文献   

10.
The synthesis of the trisaccharide α-L -Fucp-(1 → 2)-β-D -Galp-(1 → 3)-β-D -GalpNAc-1-OPr ( 2 ) is described. The N-acetylgalactosamine 6 was obtained from 4 by an intramolecular displacement of a (trifluoromethyl)sulfonyloxy by a pivaloyloxy group with its concomitant migration from position 3 to position 4 (Scheme 1). The galactosyl donor 9 was obtained from 7 via 8 by regioselective opening of the orthoester function with AcOH/pyridine followed by treatment with CCl3CN and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) (Scheme 2). Glycosylation of 6 with 9 in the presence of BF3 · OEt2 gave the disaccharide 10 . Selective deprotection of 10 at O? C(2′) followed by glycosylation with 12 and by standard deprotection afforded the title trisaccharide 2 (Scheme 3). Preliminary biological testing showed that 2 is able to inhibit the binding of the monoclonal antibody MBrl to the target tumor cells MCF7 in a dose-dependent manner.  相似文献   

11.
5-Ethyl-3-methyllumiflavinium salts 3 (Scheme 1 ; 5-EtFl+ox+, A- in Scheme 2) may arise in situ on adding an acid (HA) to solutions of the 4a-flavin adducts 5 in low polar solvents. The acidified solutions were kept under N2 at 25° in the dark to give spontaneous accumulations of the 5-ethyl-3-methyllumiflavin radical 6 (5-EtFl· and/or 5-EtFlH+·) and of some 3-methyllumiflavin 10 (Scheme 3) in dependence on the nature of the solvent and, on the nature and the concentration of the acid.The use of TFA;TCA;AcOH;α-ketoglutaric and salicyclic aci (Table 1) gave 6 and 10 in yields of 60–90% and 6–21%, respectively. The anaerobic production of 10 limits the formation of 6 to a theoretical yield of 66.7%. On suppressing the limiting pathway (eqn 3) the formation of 6 is increased which, however, will not always be revealed by an increased accumulation of 6. In a radical termination, 6 could react with another radical to give a 4a-flavin adduct. The use of TCA in MeCN gave a decrease of 10 coupled with the increased occurrence of Cl3CCOO· and Cl3C· radicals as appeared from the spontaneous generation of CO2 (eqns 3+7). 5-EtFl· was probably trapped by Cl3,C· to give 5-ethyl-3-methyl-4a-trichloromethyllumiflavin (eqn 8). In contrast, the use of HCOOH promised the achievements of quantitative accumulations of 6 which was indeed realized (Table 2 ; Figs 2 and 3).  相似文献   

12.
Addition of various amines to the 3,3‐bis(trifluoromethyl)acrylamides 10a and 10b gave the tripeptides 11a – 11f , mostly as mixtures of epimers (Scheme 3). The crystalline tripeptide 11f 2 was found to be the N‐terminal (2‐hydroxyethoxy)‐substituted (R,S,S)‐ester HOCH2CH2O‐D ‐Val(F6)‐MeLeu‐Ala‐OtBu by X‐ray crystallography. The C‐terminal‐protected tripeptide 11f 2 was condensed with the N‐terminus octapeptide 2b to the depsipeptide 12a which was thermally rearranged to the undecapeptide 13a (Scheme 4). The condensation of the epimeric tripeptide 11f 1 with the octapeptide 2b gave the undecapeptide 13b directly. The undecapeptides 13a and 13b were fully deprotected and cyclized to the [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐D ‐valine]]‐ and [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐L ‐valine]]cyclosporins 14a and 14b , respectively (Scheme 5). Rate differences observed for the thermal rearrangements of 12a to 13a and of 12b to 13b are discussed.  相似文献   

13.
The phosphonic acids 3 and 4 were prepared to compare their inhibitory activity on Vibrio cholerae sialidase with the one of the corresponding N-acetyl-2-deoxyneuraminic acids 5 and 6 . Thus, hydrogenation and benzylation of methyl N-acetyl-2,3-didehydro-2-deoxyneuraminate (1MeNeu2en5Ac; 7) gave a mixture of the fully O-benzylated benzyl and methyl esters 9 and 10 , the partially O-benzylated benzyl and methyl esters 11 and 12 , and the fully O-and N-benzylated benzyl and methyl esters 13 and 14 (Scheme 1). Transesterification of 9 to 10 and hydrolysis of 10 gave the acid 15 . Oxidative decarboxylation of 15 with Pb(OAc)4 gave a 1:9 mixture of the α-and β-D-glycero-D-galacto-acetates 16 and 17 . Phosphonoylation of 17 with P(OMe)3 and Me3SiOTf gave a 1.3:1 mixture of the phosphonates 18 and 19 , which were deprotected to give the (4-acetamido-2,4-dideoxy-D-glycero-α-and β-D-galacto-octopyranosyl)phosphonic acids 3 and 4 , respectively. The acid 6 was obtained by epimerization of the tert-butyl ester 23 with lithium N-cyclohexylisoproylamide and deprotection. The phosphonic acids 3 (Ki 5.5 10-5 M) and 4 (Ki 2.3.10?4 M ) are stronger inhibitors of Vibrio cholerae sialidase than the anomeric N-acetyl-2-deoxyneuraminic acids 5 (Ki 2.3 10?3 M ) and 6 . Both 3 and 4 inhibit the Vibrio cholerae sialidase, while only the carboxylic acid 5 , possessing an equatorial COOH group is an inhibitor.  相似文献   

14.
A new approach to ketose-derived nucteosides is described. It is based upon a chain elongation of 1-deoxy-1-nitroaldoses, followed by activation of the nitro group as a leaving group, and introduction of a pyrimidine or purine base. Thus, the nitroaldose 7 was prepared from 3 by pivaloylation (→ 4 ), synthesis of the anomeric nitrones 5/6 , and ozonolysis of 6 (Scheme 1). Partial hydrolysis of 4 yielded 8/9 , which were characterized as the acetates 10/11 and transformed into the nitrones 12/13 . Ozonolysis of 12/13 gave 14/15 , which were acetylated to 16/17 . Henry reaction of 7 lead to 19 and 20 , which were acetylated to 21 and 22 (Scheme 2). Michael addition of 7 to acrylonitrile and to methyl propynoate yielded the anomers 23/24 and 25/26 , respectively. Similar reactions of 16/17 were prevented by a facile β-elimination. Therefore, the nitrodiol 15 was transformed into the orthoesters 27 and then, by Henry reaction, partial hydrolysis, and acetylation, into 28 and 29 (Scheme 2). The structure of 19 was established by X-ray analysis. It was the major product of the kinetically controlled Henry reaction of 7 . Similarly, the β-D-configurated nitroaldoses 23 and 25 were the major products of the Michael addition. This indicates a preferred ‘endo’-attack on the nitronate anion derived from 7 . AMI calculations for this anion indicate a strong pyramidalization at C(1), in agreement with an ‘endo’-attack. Nucleosidation of 21 by 31 afforded 32 and 33 . Yields depended strongly upon the nature and the amount of the promoter and reached 77% for 33 , which was transformed into 34 , 35 , and the known ‘psicouridine’ ( 36 ; Scheme 3). To probe the mechanism, the trityl-protected 30 was nucleosidated yielding 37 , or 37 and 38 , depending upon the amount of FeCl3. Nucleosidation of the nitroacetate 28 was more difficult, required SnCl2 as a promoter, and yielded 39 and 40 . The β-D-anomer 40 was transformed into 36 . Nucleosidation of 23 (SnCl4) yielded the anomers 41 and 42 , which were transformed into 43 and 44 , and hence into 45 and 46 (Scheme 4). Similarly, nucleosidation of 25 yielded 47 and 48 , which were deprotected to 49 and 50 , respectively. The nucleoside 49 was saponified to 51 . Nucleosidation of 21 by 52 (SnCl2) afforded the adenine nucleosides 53 and 54 (Scheme 5). The adenine nucleoside 53 was deprotected (→ 55 → 56 ) to ‘psicofuranine’ (1), which was also obtained from 58 , formed along with 57 by nucleosidation of 28 . The structure and particularly the conformation of the nitroaldoses, nitroketoses, and nucleosides are examined.  相似文献   

15.
The title compounds 6 and 7 have been prepared from the known 2,3-di-O-benzyl-4,6-O-benzylidene-D -galactose ( 18 ) and N2-acetyl-tri-O-benzyl-D -glucosamine oxime ( 29 ) in eight and six steps, respectively. The azidonitrile leading to the benzylated galacto-tetrazole 16 was prepared from 14 and cyclized under the conditions of its formation (Scheme 1). The alcohol 13 was obtained by oxidation of 10 followed by reduction. Better yields and diastereoselectivities were realized, when the benzylidene-protected D -galacto-alcohol 20 was subjected to oxido-reduction, yielding the L -altro-alcohol 22 via the ketone 21 (Scheme 2). Treatment of the corresponding tosylate 24 with NaN3 yielded the tetrazole 25 , which was deprotected to 6 . The tetrabenzyl ether 16 (from 14 , or from 25 via 27 ) was reduced to 28 and deprotected to give the known deoxygalactostain 8 (Scheme 2). Oxidation of the hydroxynitrile 30 , derived from 29 , followed by reduction of 32 yielded mostly the L -ido-hydroxynitrile (Scheme 3), which was tosylated and treated with NaN3 to give the tetrazole 35a and its manno-isomer 36a , while Al(N3)3 yielded (E)- and (Z)- 38 (Scheme 4). The intermediate azide 39 was isolated besides 40 when NH4N3/DMF was used; thermolysis of 39 gave mostly 35a , which was deprotected to 7 , besides some elimination product 41 . Both 6 and 7 are stable in the pH range 1–10; at pH 12, 6 is unaffected but, 7 shows some epimerization to the manno-configurated isomer 43 . The tetrazole 6 is a competitive inhibitor of the β-galactosidases from E. coli (K1 = 1 μM , pH 6.8) and bovine liver (K1 = 0.8 μM , pH 7.0); the N-acetyl-β-D -glucosaminidase from bovine kidney is competitively inhibited by 7 (K1 ? 0.2 μM , pH 4.1).  相似文献   

16.
The addition of dienes, diazomethane, and carbenoids to the manno- and ribo-configurated thio-γ-O-lactones 1 and 2 was investigated. Thus, 1 (Scheme 1) reacted with 2,3-dimethylbutadiene (→ 4 , 73%), cyclopentadiene (→ 5a/b 1:1, 70%), cyclohexa- 1,3-diene (→ 9a/b 2:3, 92%), and the electron-rich butadiene 6 (→ 7a/b 3:1, 82%). Wheras 5a/b was separated by flash chromatography, 7a/b was desilylated leading to the thiapyranone 8 . Selective hydrolysis of one isopropylidene group of 9a/b and flash chromatography gave 10a and 10b . The structures of the adducts were elucidated by X-ray analysis ( 4 ), by NOE experiments ( 4 , 5a , 5b , 7a/b , 10a , and 10b ), and on the basis of a homoallylic coupling ( 7a/b ). The additions occurred selectively from the ‘exo’ -side of 1 . Only a weak preference for the ‘endo’-adducts was observed. Hydrogenation of 9a/b with Raney-Ni (EtOH, room temperature) gave the thiabicyclo [2.2.2]octane 11 . Under harsher conditions (dioxane, 110°), 9a/b was reduced to the cyclohexyl ß-D C-glycoside 12 which was deprotected to 13 . X-Ray analysis of 13 proved that the desulfuration occurred with inversion of the anomeric configuration. The regioselective addition of the dihydropyridine 14 to 1 (Scheme 2) and the methanolysis of the crude adduct 15 gave the lactams 16a (32%) and 16b (38%). Desilylation of 15 with Bu4NF · 3H2O, however, gave the unsaturated piperidinedione 17 (92%) which was deprotected to the tetrol 18 (65%). Similarly, 2 was transformed via 19 (62%) into the triol 20 (74%). The cycloaddition of 1 with CH2N2 (Scheme 3) gave a 35:65 mixture of the 2,5-dihydro- 1,3,4-triazole 21 and the crystalline 4,5-dihydro 1,2,3-triazole 22 . Treatment of 21 and 22 with base gave the hydroxytriazoles 23 and 24 , respectively. The structure of 24 was established by X-ray analysis. The triazole mixture 21/22 was separated by prep. HPLC at 5°. At room temperature, 21 already decomposed (half-life 21.6 h) leading in CDCI3 solution to a complex mixture (containing ca. 20–25% of the spirothiirane 27 and ca. 7–10% of its anomer) and in MeOH solution exclusively to the O,O,S-ortholactone 26 . Crystals of 22 proved be stable at 105°. Upon heating in petroleum ether at 100°, 22 was transformed into a ca. 1:1 mixture of 27 and the enol ether 28 . The reaction of 1 with ethyl diazoacetate (Scheme 4) in the presence of Rh2(OAc)4. 2H2O gave the unsaturated esters 29 (33%) and 30 (26%), whereas the analogous reaction with diethyl diazomalonate afforded the spirothiirane 31 (68%) and the enol ether 32 (29%). Complete transformation of 31 into 32 was achieved by the treatment with P(NEt2)3. Similary, 33 (69%) was prepared from 2 .  相似文献   

17.
Glycosylsulfenyl snf (Glycosylthio) sulfenyl Halides (Halogeno and Halogenothio 1-Thioglycosides, Resp.): Preparation and Reaction with Alkenes The disulfides 11–17 and 20 were prepared from 7, 9 , and 18 via the dithiocarbonates 8, 10 , and 19 , respectively (Scheme 2). The structure of 11 and of 13 was established by X-ray analysis. Chlorolysis (SO2Cl2) of 11 gave mostly the sulfenyl chloride 24 , characterized as the sulfenamide 26 , a small amount of 21 , characterized as the (glycosylthio)sulfenamide 23 , and the glycosyl chloride 27 (Scheme 3). Bromolysis of 11 followed by treatment of the crude with PhNH2 yielded only 28 . Chlorolysis of the diglycosyl disulfide 13 , however, gave mostly the (glycosylthio)sulfenyl chloride 21 and 27 , besides 24 . Bromolysis of 13 (→ 22 and traces of 25 ) followed by treatment with PhNH2 gave an even higher proportion of 23 . Similarly, 20 led to 29 and hence to 30 . In solution (CH2Cl2), the sulfenyl chloride 24 decomposes faster than the (thio)sulfenyl chloride 21 , and both interconvert. Addition of crude 24 to styrene (?78°) yielded the chloro-sulfide 31 and some 37 , both in low yields. The product of the addition of 24 to l-methylcyclohexene was transformed into the triol 32 . Silyl ethers of allylic alcohols reacted with 24 only at room temperature, yielding, after desilylation, isomer mixtures 33 and 34 , and pure 35 . Much higher yields were achieved for the addition of (thio)sulfenyl halides yielding halogeno-disulfides. Good diastereoselctivites were only obtained with 21 , its cyclohexylidene-protected analogue, and 22 , and this only in the addition to styrene (→ 36, 37, 38 ), to (E)-disubstituted alkenes (→ 46, 48, 49a/b, 50a/b, 53 ), and to trisubstituted alkenes (→ 47, 51, 52, 54, 55 ). Other monosubstituted alkenes (→ 41–45 ) and (Z)-hex-2-ene (→ 49c/d,50c/d ) reacted with low diastereoselectivities. Where structurally possible, a stereospecific trans-addition was observed; regioselectivity was observed in the addition to mono- and trisubstituted alkenes and to derivatives of allyl alcohols. The absolute configuration of the 2-chloro-disulfides was either established by X-ray analysis ( 47a ) or determined by transforming (LiAlH4) the chloro-disulfides into known thiiranes (Scheme 5). Thus, 37, 48 , and the mixture of 49a/b and 50a/b gave the thiiranes 56, 61 , and 64 , respectively, in good-to-acceptable yields (Scheme 5). Harsher conditions transformed 56 into the thiols 57 and 58 . Similarly, 61 gave 62 . The enantiomeric excesses of these thiols were determined by GC analysis of their esters obtained with (?)-camphanoyl chloride. Addition of 21 to {[(E)-hex-2-enyl]oxy}trimethylsilane, followed by LiAlH4 reduction and desilylation, gave the known 66 (63%, e.e. 74%). The diastereoselectivity of the addition of 21 to trans-disubstituted and trisubstituted alkenes is rationalized by assuming a preferred conformation of the (thio)sulfenyl chloride and destabilizing steric interactions with one of the alkene substituents, while the diastereoselectivity of the addition to styrene is explained by postulating a stabilizing interaction between the phenyl ring and the C(1)–S substituent (Fig.4).  相似文献   

18.
A linear and a convergent synthesis of uridine‐derived backbone‐base‐dedifferentiated (backbone including) oligonucleotide analogues were compared. The Sonogashira cross‐coupling of the alkyne 1 and the iodide 2 gave the dimer 4 that was C‐desilylated and again coupled with 2 to give the trimer 6 (Scheme 1). Repeating this linear sequence led to the pentamer 10 . Coupling yields were satisfactory up to formation of the trimer 6 , but decreased for the coupling to higher oligomers. Similarly, coupling of the alkynes 5, 7 , and 9 with the iodouridine 3 gave, in decreasing yields, the trimer 12 , tetramer 13 , and pentamer 14 , respectively. The dimeric iodouracil 20 was synthesized by coupling the alkyne 17 with the iodide 16 to the dimer 18 , followed by iodination at C(6/I) to 19 and O‐silylation (Scheme 2). The iodinated dimer 23 was prepared by iodinating and O‐silylating the known dimer 21 . Coupling of 20 and 23 with the dimer 5 , trimer 7 , and tetramer 9 gave the tetramers 8 and 13 , the pentamers 10 and 14 , and the hexamer 15 , respectively (Scheme 3). The oligomers up to the pentamer 14 were deprotected to provide the trimer 24 , tetramer 25 , and pentamer 26 (Scheme 4). There was no evidence for the heteropairing of the pentamer 26 and rA7 , nor for the pairing of rU5 and rA7, while a UV melting experiment showed the beginning of a sigmoid curve for the interaction of rU7 with rA7. Therefore, the pentamer 26 does not pair more strongly with rA7 than rU5.  相似文献   

19.
The 1,4‐cis‐diethynylated α‐D ‐mannopyranose analogue 11 has been prepared from 1,6 : 2,3‐dianhydro‐β‐D ‐allopyranose ( 6 ) by alkynylating epoxide and acetal opening (Scheme 2). Eglinton coupling of 11 gave the cyclodimer 18 (Scheme 3). Crystal‐structure analysis of the corresponding bis(methanesulfonate) 19 revealed substantially bent butadiyne moieties; one mannopyranosyl ring adopts the 4C1 and the other one a slightly distorted OS2 conformation (Fig. 1). Hydrogenation of 18 , followed by deprotection, gave the stable butane‐1,4‐diyl‐bridged cyclodimer 21 (Scheme 3). Crystal‐structure analysis shows the 4C1 conformation of the mannopyranosyl units (Fig. 2). The two butane fragments are characterised by a combination of gauche and antiperiplanar arrangements.  相似文献   

20.
Furanoid and pyranoid glyconothio-O-lactones were prepared by photolysis of S-phenacyl thioglycosides or by thermolysis of S-glycosyl thiosulfinates, which gave better results than the thionation of glyconolactones with Lawesson's reagent. Thermolysis of the thiosulfinates obtained from the dimannofuranosyl disulfide 7 or the manofuranosyl methly disulfide 8 (Scheme 2) gave low yields of the thio-O-lactone 2 . However, photolysis of the S-phenacyl thioglycoside 6 obtained by in situ alkylation of the thiolato anion derived from 5 led in 78–89% to 2 . Similarly, the dithiocarbonate 10 was transformed, via 11a , into the ribo-thio-O-lactone 12 (79%). Thermolysis of the peracetylated thiosulfinates 14 (Scheme 3) led to the intermediate thio-O-lactone 15 , which underwent facile β-elimination of AcOH (→ 16 , 75%) during chromatography. The perbenzylated S-glucopyranosyl dithiocarbonate 18 (Scheme 4) was transformed either into the S-phenacyl thioglucoside 19 or into a mixture of the anomeric methyl disulfides 21a/b . Whereas the photolysis of 19 led in moderate yield to 2-deoxy-thio-O-lactone 20 , oxidation of 21b and thermolysis of resulting thiosulfinates gave the thio-O-lactone 4 (79%), which was transformed into 20 (36%) upon photolysis. The pyranoid manno-thio-O-lactone 26 was prepared in the same way and in good yields from 22 via the dithiocarbonate 24b and the disulfide 25 . The ring conformations of the δ-thio-O-lactones, flattened 4C1 for 15 and 4 and B2,5 for 26 , are similar to the ones of the O-analogous oxo-glyconolactones. The reaction of 2 (Scheme 5) with MeLi and then with MeI gave the thioglycoside 27 (29%) and the dimeric thio-O-lactone 29 (47%). The analogous treatment of 2 with lithium dimethylcuprate (LiCuMe2) and MeI led to a 4:1 mixture (47%) of 31 and 27 . The structure of 2 was proven by an X-ray analysis, and the configuration at C(6) and C(5) of 29 was deduced from NOE experiments. Substitution of MeI by CD3I led to the CD3S analogues of 27 , 29 , and 31 , i.e. 28 , 30 , and 32 , respectively, evidencing carbophilic addition and ‘exo’-attack on 2 by MeLi and the enethiolato anion derived from 2 . The preferred ‘endo’-attack of LiCuMe2 is rationalized by postulating a single-electron transfer and a diastereoselective pyramidalization of the intermediate radical anion.  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