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.
The addition of phosphite anions and of tris(trimethylsilyl) phosphite (P(OSiMe3)3) to N-glycosyl-C-arylnitrones was examined. While these nitrones proved inert towards the phosphite anions, they reacted with P(OSiMe3)3 under catalysis by Lewis acids. Thus, P(OSiMe3)3 reacted with the crystalline (Z)-N-glycosylnitrones 2 and 8 to give the optically active N-hydroxy-α-aminophosphonic acids 4 and 10 , respectively, and hence the α-aminophosphonic acids 5 and 11 in yields up to 92% and with an enantiomeric excess (e.e.) up to 97% (Scheme 1). The absolute configuration of the phosphonates depend upon the nature and – in one case – upon the quantity of the catalyst (Figure). Upon catalysis by HCIO4 or Zn(OTF)2, p(OSiMe3)3 added to 2 to give, in both cases, the (+)-(R)-phenylphosphaglycine 5 (optical purity 79–84 and 90–93%, resp.). The optical purity (o.p.) was hardly influenced by the amount of these catalysts (0.02-;1 equiv.). However, catalysis by ZnCl2 gave, with trace quantities of the catalyst, (–)-(S)- 5 (o.p. 79%), while an equimolar amount of ZnCl2 yielded (+)-(R)- 5 (o.p. 82%). The HClO4-catalyzed addition of P(OSiMe3)3 to the nitrone 14 (Scheme 2) led to (+)-(R)-N-hydroxyphosphavaline 15 (78%) and hence to (–)-(R)-phosphavaline 16 (71% from 14 e.e. 95%). Under conditions leading from the nitrones 2 , 8 , 14 , and 20 (Schemes 1 and 2) predominantly to (R)-α-aminophosphonic acids, the addition of P(OSiMe3)3 to nitrone 18 , possessing a benzyloxy substituent as an additional potential ligand for the catalyst, gave (S)-phosphaserine 19 . The addition of P(OSiMe3)3 to the nitrone 20 , catalyzed by Zn(OTf)2, led to (+)-(R)-N-hydroxyphosphamehionine 21 (71%, e.e. 77%) and hence to (–)-(R)-phosphamethionine 22 (77% from 20 , e.e. 79%). Catalysis by trace quantities of ZnCl2 gave (+)-(S)- 22 (85%, e.e. 61%). The enantiomerically pure aminophosphonic acids 5 , 11 , and 16 were obtained by recrystalliztion. The e.e. of the N-hydroxyaminosphosphonic acids 10 , 15 , and 21 and the aminophosphonic acids 5 , 11 , 16 , and 22 were determined by the HPLC analysis of the dimethyl N-naphthoyl-α-aminophosphonats 7 , 13 , 17 , and 23 , on a chiral stationary phase.  相似文献   

2.
Addition of dialkyl phosphites to the nitrone 6 , formed in situ from the oxime 5 and formaldehyde gave the hydroxylamines 7 (86%) and 8 (88%), which reacted with p-benzoquinone in the presence of ethylene via the C-dialkoxyphosphonoylnitrones 9 and 10 to yield the cycloaddition products 11 – 14 (80–85%) with a diastereoselectivity of about 50%. The cycloaddition products were transformed into the monoisopropylidene derivatives 15 – 18 and the diacetates 19 – 22 . Comparison of the NMR spectra and the specific rotations of the compounds 19 – 22 with those of the corresponding α-ammo-acid derivatives 23 – 26 of known configuration indicated preferential formation of the L -isomers. The cycloaddition products were transformed in good yield into the L -α-aminophosphonic acids 29 , 30 , 36 , and 39 .  相似文献   

3.
The reaction between the chiral allenic acid (+)-(S)- 1 and the carbodiimides 2a – d and the keten-imine 6 gives, under mild conditions, the tricyclic compounds 3 – 5 , 7 , and 8 . Low diastereoselectivity and a partial loss of optical activity are observed. A stepwise mechanistic pathway via a biradical intermediate is postulated.  相似文献   

4.
The N-benzyl- and N-alkyl-substituted 1,2-thiazetidin-3-one 1,1-dioxides 1b – d reacted readily with NH3 and primary amines via ring opening. The reaction with NH3 proceeded at −78°→room temperature yielding ring-opened adducts via nucleophilic attack of NH3 at the sulfonyl group, whereas the reactions with amines at room temperature yielded products via attack at the carbonyl group. The N-unsubstituted analogue 1a , when reacted with benzylamine in refluxing EtOH, also gave a product of ring opening via nucleophilic attack at the carbonyl group of 1a . The transamidation-like reactions of the 2-(aminoalkyl)-1,2-thiazetidin-3-one 1,1-dioxides 19a – d proceeded via six-, seven-, and eight-membered intermediates, giving the ring-enlarged eight-, nine-, and ten-membered products 21 – 24 (Schemes 8 and 9), respectively, in 42 – 87% yields. The products resulted from the nucleophilic attack of the amino group of the side chain at the carbonyl C-atom. The structure of the eight-membered product 24 with an asymmetrically situated methyl substituent was established by X-ray crystallography.  相似文献   

5.
Abstract

The addition of the lithium carbanion of (R)-(+)-3,4-dimethoxybenzyl 2-methoxy-1-naphthyl sulfoxide 3 to cyclic nitrone 4, under kinetically controlled conditions gave isoquinoline sulfoxide derivatives 5 and 6 in high diastereoselectivities, Under equilibrium controlled conditions poor diastereoselectivity results. The chiral (R)-(+)-3,4-dimethoxybenzyl 2-methoxy-I -naphthyl sulfoxide 3 was easily prepared by the reaction of 3.4-dimethoxyben-zylmagnesium chloride 2 with (-)-(S)menthyl 2-methoxy-naphthalenesulfinate 1 in dry benzene. This methodology allows for the synthesis of the isoquinoline alkaloid (R)-(-)-norlaudanosine 8 in three efficient synthetic steps.  相似文献   

6.
Carbocycles from monosaccharides. III. Concerning the diastereoselective formation of cyclopentane derivatives. Transformations in the galactose series. The diastereoselectivity of the intramolecular nitrone-olefine cycloaddition of 1 , 3 and 4 (Scheme 1), yielding only 2 , 5 and 6 but none of the isomers 8 , 9 and 10 is explained by assuming a kinetic control and postulating that the relative activation energies of the two relevant transition states in the cyclization of e.g. 1 can be estimated from the conformers A and B , the latter being destabilized by a synperiplanar arrangement of the nitrone function and the 2-alkoxy-group (Scheme 2). It is further postulated, that this destabilization is responsible for the formation of (2,3)-trans configurated products. Since 2 , 5 and 6 are presumably thermodynamically more stable than 8 , 9 and 10 , a case was investigated, where the cycloaddition can either give thermodynamically less stable (2,3)-trans-product such as 12 or a thermodynamically more stable (2,3)-cis-product such as 13. 12 and 13 could both be formed from the aldehyde 25 via the nitrone 11 (Schemes 3 and 5). Treatment of the galactoside 16 first with Zn in aqueous butanol (forming among other products 25 and its 2-debenzyl-oxy-derivative) and then with N-Methyldroxylamine yielded the isoxazolidines 12 (72%), 13 (2%) and 27 (7%) (Schemes 4 and 6). Similarily, the anomeric silylated galactosides 17 and 23 gave 29 (78% from 17 , 77% from 23 ) and 27 (5% and 3%). Upon desilylation, 29 gave 32 , which was converted into 12 . The structure of the isoxazolidines was unambiguously deduced from their NMR. spectra and those of their derivatives 33 and 34 . Compound 32 was further transformed into its deoxyderivative 36 . The high diastereoselectivity of the cycloaddition restricts the number of diastereomeric, pentasubstituted cyclopentanes available by this method. However, cyclization of the 2-Hydroxy-aldehyde 37 (Scheme 8) gave the kinetically less favoured isomer 40 in a higher proportion, showing the differential influence of hydrogen-bonds on the relevant activation energies. Thermolysis of 32 gave 40 (79%) and 41 (11%). The structure of 41 was deduced from its NMR. spectra and those of its derivatives 42 and 43 . Thermolysis of 29 gave, after desilylation, 41 (42%), 40 (22%) and 32 (13%) and thermolysis of 6 lead to a 25 : 75 equilibrium with 44 (combined yield 90%). These transformations illustrate means leading to additional isomers and are in agreement with the proposed explanation of the diastereoselectivity in question.  相似文献   

7.
Intramolecular N-alkylation of 2,3-O-isopropylidene-5-O-methanesulfonyl-6-O-t-butyldimethylsilyl-d-mannofuranose-oxime 7 afforded a five-membered cyclic nitrone 9, which on NO bond reductive cleavage followed by deprotection of –OTBS and acetonide functionalities gave 1,4-dideoxy-1,4-imino-l-allitol (DIA) 3. Addition of allylmagnesium chloride to nitrone 9 afforded α-allylated product 10a in high diastereoselectivity providing an easy entry to N-hydroxy-C1-α-allyl-substituted pyrrolidine iminosugar 4a after removal of protecting group, while NO bond reductive cleavage in 10a afforded C1-α-allyl-pyrrolidine iminosugar 4b.  相似文献   

8.
The oxidation of (‐)‐tabersonine ( 1 ) with dimethyldioxirane (DMD) in neutral and acidic medium gave 16‐hydroxytabersonine‐N‐oxide ( 3 ) and the didehydrovincamine isomers 4 and 5 , respectively. (+)‐14,15‐Didehydro‐quebrachamine ( 7 ) furnished the hydroxyindolenine 9 , and the pentacyclic derivative 11 . (+)‐Quebrachamine ( 8 ) and DMD in neutral medium gave (7S,20S)‐(+)‐rhazidigenine ( 12 ) which was converted to (2R,7S,20S)‐(+)‐rhazidine ( 13b ) with hydrochloric acid.  相似文献   

9.
《Tetrahedron: Asymmetry》2001,12(12):1683-1688
The enantioselective Michael addition of a chiral imine of 4-protected 2-methylcyclohexane-1,4-dione to phenyl crotonate led after cyclization to the corresponding bicyclic lactam. Reductive cleavage of the chiral moiety followed by saponification gave the corresponding keto-acid, which was cyclized to afford a lactone. Belleau–Fujimoto reaction of the lactone then led to the title building block (diastereoselectivity 96:4, e.e. >98%) in 11% overall yield from the starting dione. (8R,8aS)-(+)-8,8a-Dimethyl-3,4,6,7,8,8a-hexahydronaphthalen-2(1H)-one was obtained after reduction of the carbonyl group, acetylation, and reductive cleavage–deprotection (52% overall yield), representing a formal synthesis of (+)-valencenol.  相似文献   

10.
《合成通讯》2013,43(22):3413-3417
ABSTRACT

Conditions for efficient resolution of (±)-1-(9-anthryl)ethylamine ((±)-1) by fractional crystallization of its salts with (S)-(+)-mandelic acid (2) are reported. When crystallization was performed by fast addition of chloroform solution of an equivalent of (±)-1 to the hot chloroform solution of (+)-2, crystals of mandelate of (+)-1-(9-anthryl)ethylamine ((R,S)-3) are collected in 56% yield. (R)-(+)-1 (98.6% e.e.) is isolated by extraction from bicarbonate solution of mandelate salt. Ulterior collection of four crops afforded (R,S)-3 with 71.5% cumulative yield and >98% e.e. of (+)-1 in a any single crop. With only 0.5 equivalents of (+)-2 crystallization afforded (R,S)-3 in 47.4% yield and (+)-1 with 98.1% e.e.  相似文献   

11.
The antibiotic myxovirescine M2 was synthesized from seven building blocks ( 1 – 7 , Scheme 1), with the following chiral starting materials being employed: (S)-malic acid, (+)-D -ribonolactone, (S)-2-(hydroxymethyl)butanoate, and (2R,4S)-5-hydroxy-2,4-dimethylpenLanoate. Three new nucleophilic reagents, 8 – 10 , for C-C bond formation have been used. The key steps of the synthesis are: a Suzuki coupling between an alkyl borane and a vinyl bromide ( 4 + 12e → 13 ), a Julia olefinalion ( 14 + 17 → 18 ), and a Yamaguchi macrolactonizalion to form the 28-membered lactone ( 18 → 19 ), This extremely convergenl synthetic approach will allow the preparation of a number of the 31 known myxovirescine molecules.  相似文献   

12.
L -Aspartic acid by successive N-tosylation, anhydride formation, and reduction was converted into (3S)-3-(tosylamino)butano-4-lactone ( 4 ). Electrophilic methylation of 4 , subsequent iodo-esterification and nucleophilic methylation, followed by saponification and deprotection, gave (2S, 3R)-3-amino-2-methylpentanoic acid ( 2 ) with an ee of > 99% in seven steps and in an overall yield of 34%.  相似文献   

13.
The replacement of 2,3-O-cyclohexylidene-d-glyceraldehyde with (2R,5R,6R)-5,6-dimethoxy-5,6-dimethyl-1,4-dioxane-2-carbaldehyde (Ley’s aldehyde) has led to significant improvements in the isolation of both diastereoisomers of the respective 2,3-O-BDA 1,2,3-trihydroxypropylphosphonates. The triethylamine-catalysed addition of dialkyl phosphites and lithium diethyl phosphonate gave the products in moderate (ca. 1:2) diastereoselectivity while the application of diethyl trimethylsilyl phosphite afforded a 1:9 mixture of diethyl (R)- and (S)-hydroxy-[(2R,5R,6R)-5,6-dimethoxy-5,6-dimethyl-1,4-dioxan-2-yl]methylphosphonates.  相似文献   

14.
Addition of dialkyl phosphites to (S)-N-tert-butanesulfinyl imine derived from fluoral afforded, under mild conditions, the corresponding N-tert-butanesulfinyl α-aminophosphonates in moderate to high yields and diastereoselectivity. The major diastereomers of N-tert-butanesulfinyl α-aminophosphonates of (SS,S) configuration were isolated, and after partial or complete deprotection, converted into enantiomerically pure phosphonotrifluoroalanine and its dialkyl esters.  相似文献   

15.
The synthesis and catalytic properties of a new type of enantioselective phase-transfer catalysts, incorporating both the quinuclidinemethanol fragment of Cinchona alkaloids and a 1,1′-binaphthalene moiety, are described. Catalyst (+)-(aS,3R,4S,8R,9S)- 4 with the quinuclidine fragment attached to C(7′) in the major groove of the 1,1′-binaphthalene residue was predicted by computer modeling to be an efficient enantioselective catalyst for the unsymmetric alkylation of 6,7-dichloro-5-methoxy-2-phenylindanone ( 1 ; Scheme 1, Fig. 1). Its synthesis involved the selective oxidative cross-coupling of two differently substituted naphthalen-2-ols to afford the asymmetrically substituted 1,1′-binaphthalene derivative (±)- 17 in high yield (Scheme 3). Chromatographic optical resolution via formation of diastereoisomeric camphorsulfonyl esters and functional-group manipulation gave access to the 7-bromo-1,1′-binaphthalene derivative (−)-(aS)- 11 (Scheme 4). Nucleophilic addition of lithiated (−)-(aS)- 11 to the quinuclidine Weinreb amide (+)-(3R,4S,8R)- 8 afforded the two ketones (aS,3R,4S,8R)- 27 and (aS,3R,4S,8S)- 28 as an inseparable mixture of diastereoisomers (Scheme 6). Stereoselective reduction of this mixture with DIBAL-H (diisobutylaluminum hydride; preferred formation of the C(8)−C(9) erythro-pair of diastereoisomers with 18% de) or with NaBH4 (preferred formation of the threo-pair of diastereoisomers with 50% de) afforded the four separable diastereoisomers (+)-(aS,3R,4S,8S,9S)- 29 , (+)-(aS,3R,4S,8R,9R)- 30 , (−)-(aS,3R,4S,8S,9R)- 31 , and (+)-(aS,3R,4S,8R,9S)- 32 (Scheme 6). A detailed conformational analysis, combining 1H-NMR spectroscopy and molecular-mechanics computations, revealed that the four diastereoisomers displayed distinctly different conformational preferences (Figs. 2 and 3). These novel Cinchona-alkaloid analogs were quaternized to give (+)-(aS,3R,4S,8R,9S)- 4 , (+)-(aS,3R,4S,8S,9S)- 5 , (+)-(aS,3R,4S,8R,9R)- 6 , and (−)-(aS,3R,4S,8S,9R)- 7 (Scheme 7) which were tested as phase-transfer agents in the asymmetric allylation of phenylindanone 1 . Without any optimization work, (+)-(aS,3R,4S,8R,9S)- 4 was found to catalyze the allylation of 1 yielding the predicted enantiomer (+)-(S)- 3b in 32% ee. The three diastereoisomeric catalysts (+)- 5 , (+)- 6 , and (−)- 7 gave access to lower enantioselectivities (6 to 22% ee's), which could be rationalized by computer modeling (Fig. 4).  相似文献   

16.
The 1,2,3,4-tetrahydro-N-pivaloyl-isoquinoline-3-carboxylic acids 1d , 2d , and 3d , derived from (R)- or (S)-phenylalanine, (S)-dopa, and (S)-α-methyldopa, respectively, are doubly deprotonated with (tert-butyl)lithium in THF and alkylated at the 1-position (products 5 – 10 ). The major diastereoisomers formed are the result of electrophilic attack from the face opposite to the carboxylate group (rel. topicity ul-1,3). Even the addition to benzaldehyd (→ 7,8 ) is highly stereoselective (one of four diastereoisomers is formed exclusively (300-MHz 1H-NMR analysis)), if MgBr2 etherate is added prior to the electrophile. Some of the obtained amino-acid derivatives are decarboxylated by anodic oxidation in MeOH (→ 11 , 12 , 17 ) and NaBH3CN reduction, and converted to the known 1-methyl- and 1-benzyltetrahydroisoquinolines ( 15 , 16 ) of > 95% ee as well as to the phthalide isoquinoline alkaloid (+)-corlumine of ≥80% ee. The synthetic approach described here is compared with other methods of synthesizing enantiomerically pure 1-substituted tetrahydroisoquinolines (and thus an important group of alkaloids, Scheme 1).  相似文献   

17.
1,2,5-Thiadiazolidin-3-one 1,1-dioxide derivatives (±)- 1a – d and (±)- 2 were designed by molecular modeling as MHC (major histocompatibility complex) class-II inhibitors. They were prepared from the unsymmetrically N,N′-disubstituted acyclic sulfamides (±)- 4a – d (Scheme 1) and (±)- 11 (Scheme 2). These N-alkyl-N′-arylsulfamide precursors were synthesized by nucleophilic substitution of either a sulfamoyl-chloride or a N-sulfamoyloxazolidinone. Extension of base-induced cyclization methods from aliphatic to aromatic sulfamides gave access to the desired target molecules. The N-alkyl-1,2,5-thiadiazolidin-3-one 1,1-dioxide derivatives (±)- 3a – c were also prepared by the oxazolidinone route (Scheme 4) for coupling to a tetrapeptide fragment. The X-ray crystal structure of 1,2,5-thiadiazolidin-3-one 1,1-dioxide (±)- 21a was solved, and the directionality of the H-bond donor (N−H) and acceptor (SO2) groups of the cyclic scaffold determined (Figs. 1 and 2). The pKa value of the N−H group in (±)- 21a was determined by 1H-NMR titration as 11.9 (Fig. 3). Compounds (±)- 1a – d were shown to inhibit competition peptide binding to HLA-DR4 molecules in the single-digit millimolar concentration range.  相似文献   

18.
About the Stereospecific α-Alkylation of β-Hydroxyesters It was found, that dianions derived from β-hydroxyesters with lithium diisopropylamide (LDA) at ?50 to ?20° were alkylated stereospecifically (Scheme 1). The stereospecificity was 95–98%, the threo-compound (threo -2, -3 and -4) being the main product. This was proved for threo -2 and -3 by preparing the β-lactones 7 and 8 , respectively, which were pyrolyzed to trans-1, 4-hexadiene (9) and trans-1-phenyl-2-butene (10) , respectively (Scheme 2). Moreover, the acid threo -6 from threo -3 was converted by dimethylformamide-dimethylacetal to cis-1-phenyl-2-butene (11) (s. footnote 6). The alkylation of α-monosubstituted β-hydroxyesters also turned out to be stereospecific. Reduction of 16 and 18 with actively fermenting yeast furnished (+) -17 and (+) -2. respectively (Scheme 4), which were each mixtures of the (2R, 3S)- and the (2S, 3S)-isomers. Alkylation of (+) -17 with allyl bromide yielded after chromatography (2S, 3S) -19 and of (+) -2 with methyl iodide (2R, 3S) -19 , the oxidation of which finally gave (S)-(?) -20 and (R)-(+) -20 , respectively.  相似文献   

19.
Enantioselective Reactions on Porphine Type Nickel Complexes The thermodynamically controlled addition of alcohols to (+)-(1R)-[1-methyl-8H-HDP]nickelperchlorate ( 1 ; e.e. 92%) yields exclusively the corresponding cis-1,11-disubstituted porphinoids. Chemical transformation of functional groups in the alkoxy side-chain of the chiral addition product followed by acid catalyzed elimination yields the derived alcohols and 1 . By this procedure, the following enantioselective transformations were studied: methylation of meso-2,3-butandiol ( 5 ) to (+)-(2R,3S)-3-methoxy-2-butanol ( 8a ; e.e. 87%), diimide reduction of 2-ethylallyl alcohol ( 9 ) to (+)-(2R)-2-methyl-1-butanol ( 12a ; e.e. 15%), and hydride reduction of 4-hydroxy-2-butanone ( 13 ) to (+)-(3S)-1,3-butandiol ( 16a ; e.e. 20%). Addition of 2,2-dimethyl-1,3-propandiol ( 17 ) to 4 , followed by esterification of the free hydroxy group with trifluoromethanesulfonic anhydride and solvolysis of the sulfonate 19 yielded a bridged complex with unrearranged alkyl chain for which structure 20 is proposed.  相似文献   

20.
The regio-and diastereoselectivity of reactions of selected α-(1-hydroxyalkyl)acrylate derivatives with sodium methanethiolate have been investigated. The hydroxy compounds typically undergo conjugate addition with up to 66% d.e., while the acetoxy and bromo analogues favour SN' and SN reactions, respectively.  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