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.
Kenji Mori 《Tetrahedron letters》2007,48(32):5609-5611
Absolute configuration of gomadalactones A (1), B (2) and C (3), the pheromone components of the white-spotted longicorn beetle (Anoplophora malasiaca) was assigned as (1S,4R,5S)-1, (1R,4R,5R)-2 and (1S,4R,5S,8S)-3 by comparing their published CD spectra with those of (1R,5R)-(+)-4,4,8-trimethyl-3-oxabicyclo[3.3.0]oct-7-ene-2,6-dione (4) and (1S,5R,8S)-(+)-4,4,8-trimethyl-3-oxabicyclo[3.3.0]octane-2,6-dione (5) prepared from (R)-(−)-carvone (6).  相似文献   

2.
The coupling reaction of 1-tributylstannylthianthrene (5) and 2-tributylstannylthianthrene (7) in the presence of copper catalysts at rt afforded the thianthrene dimer 1,1′-bithianthrene (3), 2,2′-bithianthrene (8), and 1,2′-dithianthrene (9) in high yields. Also we obtained thianthrene oxide dimer (R,R) (S,S)-1-(10-S-monoxythianthrene-1-yl)thianthrene-10-S-monoxide (12) and (R,S) (S,R)-1-(10-S-monoxythianthrene-1-yl)thianthrene-10-S-monoxide (13) from 1-tributylstannyl-10-S-monoxythianthrene (10) under the same reaction condition. The final structural conformation of 3, 8, 9, and 12 was performed by X-ray crystallographic analysis. Further, the solvent effects in the coupling reactions were also examined.  相似文献   

3.
To synthesize (3′R,5′S)-3′-hydroxycotinine [(+)-1], the main metabolite of nicotine (2), cycloaddition of C-(3-pyridyl)nitrones 3a, 3c, and 15 with (2R)- and (2S)-N-(acryloyl)bornane-10,2-sultam [(2R)- and (2S)-8] was examined. Among them, l-gulose-derived nitrone 15 underwent stereoselective cycloaddition with (2S)-8 to afford cycloadduct 16, which was elaborated to (+)-1.  相似文献   

4.
The reaction of trans-cyclooctene with S8O yielded a novel bicyclic 1,2,3-trithiolane and trans-9,10,11-trithiabicyclo[6.3.0]undecane (7). Oxidation of the trithiolane with dimethyldioxirane yielded three monoxides, which are assigned to two isomeric 9-oxides, rel-(1R,8R,9S)-9-oxide (15) and rel-(1R,8R,9R)-9-oxide (16), and 10-oxide (17). Further oxidation of rel-(1R,8R,9S)-9-oxide (15) provided rel-(1R,8R,9S,11S)-9,11-dioxide (18) and rel-(1R,8R,9R,11S)-9,11-dioxide (19), while that of rel-(1R,8R,9R)-9-oxide (16) gave rel-(1R,8R,9R,11S)-9,11-dioxide (19) and rel-(1R,8R,9R,11R)-9,11-dioxide (20). The structures of 18 and 19 were determined by X-ray crystallography. The structures of other oxides were elucidated by the spectroscopic data and results of further chemical transformations. Two isomers, 15 and 16, isomerized to one another. A 9,11-dioxide 20 isomerized to 19, which is in equilibrium with 18, where 18 is thermodynamically the most stable isomer.  相似文献   

5.
All four stereoisomers of 4,8-dimethyldecanal (1) were synthesized from the enantiomers of 2-methyl-1-butanol and citronellal. Enantioselective GC analysis enabled separation of (4R,8R)-1 and (4R,8S)-1 from a mixture of (4S,8R)-1 and (4S,8S)-1, when octakis-(2,3-di-O-methoxymethyl-6-O-tert-butyldimethylsilyl)-γ-cyclodextrin was employed as a chiral stationary phase. Complete separation of the four stereoisomers of 1 on reversed-phase HPLC at −54 °C was achieved after oxidation of 1 to the corresponding carboxylic acid 12 followed by its derivatization with (1R,2R)-2-(2,3-anthracenedicarboximido)cyclohexanol, and the natural 1 was found to be a mixture of all the four stereoisomers.  相似文献   

6.
The synthesis of (4R,8R)- and (4S,8R)-4,8-dimethyldecanal 1 and 1a has been achieved connecting the chiral building block (R)-2-methyl-1-bromobutane 4 with (R)- and (S)-citronellol derivatives. The key intermediate 4 was obtained optically pure in five steps from methyl (S)-3-hydroxy-2-methylpropionate 2.  相似文献   

7.
β2-(3,4-Dihydroxybenzyl)-β-alanine [β2-Homo-Dopa, 1] is a novel β-amino acid homologue of Dopa, the most successful therapeutic agent in the treatment of Parkinson's disease. Enantioenriched (R)-1 and (S)-1 were obtained via the diastereoselective alkylation of enantiopure pyrimidinone (R)- and (S)-3, chiral derivatives of β-alanine, with veratryl iodide. The major diastereomeric products (2S,5R)-4 and (2R,5S)-4 were hydrolyzed with 57% HBr, and the desired β-amino acids were purified by silica gel chromatography. Alternatively, enantioenriched (R)- and (S)-1 were prepared by means of the highly diastereoselective alkylation (3,4-dimethoxybenzyl iodide) of open-chain β-aminopropionic acid derivatives (R,R,S)-8 and (S,S,R)-8 containing the chiral auxiliary α-phenylethylamine. Finally, nearly enantiopure (R)- and (S)-1 were obtained by resolution of racemic N-benzyloxycarbonyl-2-(3,4-dibenzyloxybenzyl)-3-aminopropionic acid, rac-12, with (R)- or (S)-α-phenylethylamine, followed by catalytic hydrogenolysis.  相似文献   

8.
Mikio Fujii  Hiroyuki Akita 《Tetrahedron》2008,64(22):5147-5149
The concise synthesis of (8aR)-(−)-albaconol (1) from (8aR)-albicanol (2) obtained from the lipase-assisted asymmetric acetylation of rac-2, was achieved in 45% overall yield (eight steps). By comparison of the sign of specific rotation of between synthetic (8aR)-(−)-albaconol (1) and natural (+)-albaconol (1), the absolute structure of natural (+)-1 was determined to be 1R,2R,4aS,8aS configuration.  相似文献   

9.
Six samples [(3R,14R,26R)-, (3R,14S,26R)-, (3S,14R,26S)-, and (3S,14S,26S)-1, a mixture of (3R,14R,26S)- and (3S,14R,26R)-1, and a mixture of (3R,14S,26S)- and (3S,14S,26R)-1] of miyakosyne A [1, (4E,24E)-14-methyloctacosa-4,24-diene-1,27-diyne-3,26-diol] were synthesized starting from the enantiomers of citronellal (2), employing olefin cross metathesis and R-selective asymmetric acetylation of a stereoisomeric mixture of acetylenic alcohols with vinyl acetate and lipase PS as key reactions. Separation of the eight stereoisomer of 1 by reversed phase HPLC at −56 °C was achieved after their esterification with (1R,2R)-2-(anthracene-2,3-dicarboximido)cyclohexanecarboxylic acid (16), and the natural miyakosyne A was found to be a mixture of 95.7% of (3R,14R,26R)-1 and 4.3% of (3R,14S,26R)-1. This is different from the (3R,14S,26R)-configuration of 1 as tentatively assigned by X-ray analysis.  相似文献   

10.
Tanja Grkovic 《Tetrahedron》2009,65(32):6335-207
A survey of the secondary metabolite chemistry profiles of New Zealand sponges of the genus Latrunculia has yielded new members of the discorhabdin A- and B-type families. The structure elucidation of (+)-(6R,8S)-1-thiomethyldiscorhabdin G/I (5) and both enantiomers of 16a,17a-dehydrodiscorhabdin W (6) are reported. Absolute configurations were assigned by comparison with a dataset of recently reported electronic circular dichroism spectra of discorhabdin alkaloids. A stereochemical correction of the recently reported natural product (+)-3-dihydrodiscorhabdin A from (3S,5R,6S,8S)-(7) to the C3-epimeric (+)-(3R,5R,6S,8S)-(8) and assignment of absolute configuration is also presented. Semi-synthesis of (+)-(3S,5R,6S,8S)-(7) from (+)-discorhabdin A provided further evidence for this structure revision. Cytotoxicity data is reported for 5-8.  相似文献   

11.
A convergent synthesis of (4R,15R,16R,21S)-rollicosin (1) and (4R,15S,16S,21S)-rollicosin (2) was accomplished. Hydroxy lactone 6a and/or 6b were synthesized from 4-pentyn-1-ol, and α,β-unsaturated lactone 7 was synthesized from γ-lactone 8 and 5-hexen-1-ol. Inhibitory activity of these compounds was examined with bovine heart mitochondrial complex I.  相似文献   

12.
The relative configuration of tulearin A (1) is determined by X-ray diffraction analysis of a cyclic carbonate derivative 2 and the absolute configuration (2R,3R,5S,8S,9S,15R,17S) from the 9-MTPA-esters 1R and 1S is determined using the modified Mosher’s method. A mechanism for the unexpected formation of carbonate 2 is suggested. Two N-phenyltriazolinedione derivatives 3 and 4 are also prepared. Two additional tulearins, B and C (5 and 6) are isolated in very small amounts and their structures are elucidated by spectroscopic means.  相似文献   

13.
Machiko Ono  Yuki Shida 《Tetrahedron》2007,63(41):10140-10148
(±)-(4,5-anti)-4-Benzyloxy-5-hydroxy-(2E)-hexenoic acid 6 was subjected to δ-lactonization in the presence of 2,4,6-trichlorobenzoyl chloride and pyridine to give the α,β-unsaturated-δ-lactone congener (±)-7 (87% yield) accompanied by trans-cis isomerization. This δ-lactonization procedure was applied to the chiral synthesis of (+)-(4S,5R)-7 or (−)-(4R,5S)-7 from the chiral starting material (+)-(4S,5R)-6 or (−)-(4R,5S)-6. Deprotection of the benzyl group in (+)-(4S,5R)-7 or (−)-(4R,5S)-7 by the AlCl3/m-xylene system gave the natural osmundalactone (+)-(4S,5R)-5 or (−)-(4R,5S)-5 in good yield, respectively. Condensation of (−)-(4R,5S)-5 and tetraacetyl-β-d-glucosyltrichloroimidate 22 in the presence of BF3·Et2O afforded the condensation product (−)-8 (97% yield), which was identical to tetra-O-acetylosmundalin (−)-8 derived from natural osmundalin 9.  相似文献   

14.
Yuji Takashima 《Tetrahedron》2010,66(1):197-2519
A general approach to the (S)- and (R)-isoflavans was invented, and efficiency of the method was demonstrated by the synthesis of (S)-equol ((S)-3), (R)-sativan ((R)-4), and (R)-vestitol ((R)-5). The key step is the allylic substitution of (S)-6a (Ar1=2,4-(MeO)2C6H3) and (R)-6b (Ar1=2,4-(BnO)2C6H3) with copper reagents derived from CuBr·Me2S and Ar2-MgBr (7a, Ar2=4-MeOC6H4; 7b, 2,4-(MeO)2C6H3; 7c, 2-MOMO-4-MeOC6H3), furnishing anti SN2′ products (R)-8a and (S)-8b,c with 93-97% chirality transfer in 60-75% yields. The olefinic part of the products was oxidatively cleaved and the Me and Bn groups on the Ar1 moieties was then removed. Finally, phenol bromide 9a and phenol alcohols 9b,c underwent cyclization with K2CO3 and the Mitsunobu reagent to afford (S)-3 and (R)-4 and -5, respectively.  相似文献   

15.
A novel and simple process for the preparation of enantiomerically pure (SS)-benzenesulfinamide (SS)-3a, (SS)-p-toluenesulfinamide (SS)-3b, (SS)-p-chloro-benzenesulfinamide (SS)-3c and (SS)-p-fluorobenzenesulfinamide (SS)-3d has been developed. The treatment of arylsulfinyl chlorides with (R)-N-benzyl-1-phenylethanamine in the presence of excess triethylamine gave diastereomeric mixtures of N-benzyl-N-(1-phenylethyl)-arylsulfinamides 1, which underwent spontaneous crystallization to furnish diastereomerically pure (R,SS)-N-benzyl-N-(1-phenylethyl)-arylsulfinamides (R,SS)-1a-1d in 28%, 29%, 27% and 31% yields, respectively. The diastereomerically pure compounds (R,SS)-1 were then converted into four enantiopure (RS)-methyl arylsulfinates (RS)-2, and finally into four enantiopure (SS)-arylsulfinamides (SS)-3 in good yields.  相似文献   

16.
(1R,2S,3S,5R,7aR)-1,2-Dihydroxy-3-hydroxymethyl-5-methylpyrrolizidine[(−)-3-epihyacinthacine A5, 1a] and (1S,2R,3R,5S 7aS)-1,2-dihydroxy-3-hydroxymethylpyrrolizidine[(+)-3-epihyacinthacine A5, 1b] have been synthesized either by Wittig's or Horner-Wadsworth-Emmond's (HWE's) methodology using aldehydes 4 and 9, both prepared from (2S,3S,4R,5R)-3,4-dibenzyloxy-2′-O-tert-butyldiphenylsilyl-2,5-bis(hydroxymethyl)pyrrolidine (2, partially protected DADP), and the appropriate ylides, followed by cyclization through an internal reductive amination process of the resulting α,β-unsaturated ketones 5 and 10, respectively, and total deprotection.  相似文献   

17.
Chiral tetrahydropentalenes (3aR,6aR)-1 have been prepared and used as ligands in the Rh-catalyzed 1,4-addition of 1-alkenylboronic acids to cyclic enones 5. It has been discovered that the stereochemistry of the reaction was controlled by the steric properties of the aryl groups in 1 rather than their electronic nature. In the vinylation with (E)-2-phenylethenylboronic acid 5, ligands (3aR,6aR)-1 provided enantioselectivity up to 87% ee and gave high yields of ethenylketones 6 in the presence of 1 (6.6 mol %). The configuration of all ketone products obtained with (3aR,6aR)-1 is (S). Rh-catalyzed reaction of cyclopentenone 4a and (Z)-propenylboronic acid 7 in the presence of ligands (3aR,6aR)-1 yielded at 50 °C an inseparable mixture of (Z)- and (E)-ketones 8 with (Z)-8 as the major product and both in only moderate enantiomeric excess.  相似文献   

18.
Epoxidations of trans-β-methylstyrene, trans-stilbene and trans-methyl p-methoxycinnamate using chiral dioxiranes derived from both enantiopure diastereomers of α-fluoro cyclohexanones, (2S, 5R)-3a-6a and (2R, 5R)-3e-6e are studied and compared. From ab initio calculations at the HF/6-31G level of conformational inter-conversion for (2S, 5R)-D5a and (2R, 5R)-D5e dioxiranes it was found that, due to the α-fluorine atom, conformer K1 is more stable in the case of (2S, 5R)-D5a while conformer K2 is more stable in the case of (2R, 5R)-D5e. However, in both cases, the more stable conformers, K1 and K2, undergo rapid inter-conversion. Therefore, based on slow epoxidation reactions and rapid ring inversion of six-membered ring dioxiranes the Curtin-Hammett principle holds. Conformation K2 with axial fluorine having been found to be more reactive, the inversion of configuration observed for the epoxides obtained with ketones 3e-6e (compared with ketones 3a-6a) could be rationalized from competitive reactions of K2 and K1 conformations leading to simultaneous production of both (−) and (+) epoxides in the case of ketones 3e-6e.  相似文献   

19.
Masayoshi Tsubuki 《Tetrahedron》2005,61(5):1095-1100
Wittig rearrangement of 17(20)-ethylidene-16-furfuryloxy steroids 5-8 was examined. Reaction of 17E(20)-ethylidene-16α-furfuryloxy steroid 5 with t-BuLi in THF afforded (20S,22S)- and (20S,22R)-22-hydroxy steroids 9, 10 and 17Z(20)-ethylidene-16α-(2-furyl)hydroxymethyl steroid 11 in 61, 28, and 9% yields, respectively. Base treatment of 17E(20)-ethylidene-16β-furfuryloxy steroid 7 gave (20R,22R)-22-hydroxy steroid 13 and 17Z(20)-ethylidene-16β-(2-furyl) hydroxymethyl steroid 14 in 60 and 17% yields. In contrast, 17Z(20)-ethylidene-16-furfuryloxy steroids 6, 8 led to the corresponding 2,3-rearranged products in low yields (25% for (20R,22S)-22-hydroxy steroid 12; 31% for (20S,22R)-22-hydroxy steroid 10). Both (20S,22S)- and (20S,22R)-22-hydroxy steroids 9, 10 were converted by catalytic hydrogenation into known compounds 16, 17, key intermediates for the synthesis of biologically active steroids.  相似文献   

20.
Kazuhiko Sakaguchi 《Tetrahedron》2003,59(34):6647-6658
Cationic rearrangement of several α-hydroxysilanes is described. Treatment of both (1R,1′R,2′S)-α-hydroxycyclopropylsilane syn-9 and (1S,1′R,2′S)-anti-9 under aqueous H2SO4 underwent rearrangement via a common α-silyl cation intermediate A to give a mixture of the ring-opened (R)-vinylsilane 13, the tandem [1,2]-CC bond migration product (1R,2S,1′R)-14, and its 1′S isomer 15. On the other hand, the acidic treatment of (R,E)-α-hydroxyalkenylsilane 8 or (R,Z)-8 was each accompanied with partial racemization to give an enantiomeric isomer of allylic alcohol 23 via a preferential syn-facial SN2′ reaction, respectively. Both α-hydroxyalkynylsilane 6 and α-hydroxyalkylsilane 12 were inert to the acidic conditions; however, treatment of (R)-α-mesyloxyalkynylsilane 26 under aqueous H2SO4 gave a mixture of the optically active rearranged allene 27, α,β-unsaturated ketone 28, and (S)-α-hydroxyalkynylsilane 6 with partial racemization. Comparisons of the reactivities of these α-hydroxysilanes under acidic conditions are also disclosed.  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