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 hypothesis which explains the diastereoselectivity of the 1,3-dipolar cycloaddition of the N-glycosylnitrones 1 – 3 leading to the 5,5-disubstituted isoxazolidines 4 – 6 on the basis of a kinetic anomeric effect predicts that nucleophiles should add to N-glycosylnitrones with a high degree of diastereoselectivity. To test this prediction, the nucleophilic addition of lithium and potassium dialkylphosphites to the crystalline (Z)-nitrone 11 , prepared from oxime 9 and (benzyloxy)acetaldehyde has been examined. The addition of lithium phosphites gave the N-glycosyl-N-hydroxyaminophosphonates 12 – 16 (d. e. 78–92%) in high yields (Scheme 4). The addition of potassium phosphites showed a much lower diastereoselectivity. Glycoside cleavage, hydrogenolysis, and dealkylation of 12 – 16 gave (+)-(S)-phosphoserine (+)- 19 (34–45% from 9 ). Its absolute configuration was confirmed by an X-ray analysis of the N-(3,3,3-trifluoro-2-methoxy-2-phenylpropionyl) derivative 24 . Similarly, the crystalline nitrone 25 gave the N-glycosyl-N-hydroxyaminophosphonate 26 , which was transformed into (+)-(S)-phosphovaline (+)- 31 (42% from 9 ). The diastereoselectivity of the nucleophilic addition and the enantiomeric purity of (+)- 31 were determined by the analysis of the derivative 30 (d.e. 92%) and 32 (d.e. 93%), respectively. The addition of lithium diethyl phosphite to the nitrone 33 , prepared in situ, gave the N-glycosyl-N-hydroxyaminophosphonate 34 , (41%; d.e. 91%), which was transformed in (+)-(S)-phosphoalanine (+)- 37 (21% from 9 ).  相似文献   

2.
L -Aspartic acid by tosylation, anhydride formation, and reduction with NaBH4 was converted into (3S)-3-(tosylamino)butan-4-olide ( 8 ; Scheme 1). Tretment of 8 with ethanolic trimethylsilyl iodide gave the N-protected deoxy-iodo-β-homoserine ethyl ester 9 . The latter, on successive nucleophilic displacement with lithium dialkyl-cuprates ( → 10a–e ), alkaline hydrolysis ( → 11a–e ), and reductive removal of the tosyl group, produced the corresponding 4-substituted (3R)-3-aminobutanoic acids 12a–e (ee > 99%). Electrophilic hydroxylation of 8 ( → 19 ; Scheme 3), subsequent iodo-esterification ( → 21 ; Scheme 4), and nucleophilic alkylation and phenylation afforded, after saponification and deprotection, a series of 4-substituted (2S, 3R)-3-amino-2-hydroxybutanoic acids 24 including the N-terminal acids 24e ( = 3 ) and 24f ( = 4 ) of bestatin and microginin (de > 95%), respectively.  相似文献   

3.
Lithium or sodium dialkyl phosphites and diamido phosphites 3 undergo addition to (+)-(S)-benzylidene-p-toluenesulfinamide 2 affording N-sulfinyl-α-aminophosphonates 4 in a diastereoisomeric ratio from 63:37 to 94:6. The major diastereoisomers formed in addition of lithium dimethyl phosphite 3a and lithium bis-diethylamido phosphite 3e to (+)-(S)-2 were separated and converted into enantiopure (+)-(R)- and (−)-(S)-α-aminobenzyl phosphonic acids 5, respectively.  相似文献   

4.
Authentic α-santolinenone ( = (+)-(4R)-1(7)-p-menthen-2-one; (+)- 1 ) is made available for the the first time in 30% overall yield from (+)-(4R)-p-menthene ((+)- 2 ) via the diastereoisomeric allylic alcohols (+)- 4a /(+)- 4b , which are oxidized to (+)- 1 with Ag2CO3/Celite. Yields are good, except for the last stage; indeed, only alcohol (+)- 4a , with equatorial OH-group, undergoes oxidation, and (+)- 1 is partly substracted via a hetero Diels-Alder dimerization giving a mixture of the diastereoisomeric dihydropyrans (+)- 5a /(+)- 5b . When Cr(VI) reagents ae used, (+)- 4a /(+)- 4b mainly give phellandral ( 6 ) and carvotanacetone ( 7 ), NnO2 reacts too sluggishly with (+)- 4a /(+)- 4b . A camphor pyrolyzate, previously thought to be 1 must be a different compound, probably 7 .  相似文献   

5.
Depsipeptides and cyclodepsipeptides are analogues of the corresponding peptides in which one or more amide groups are replaced by ester functions. Reports of crystal structures of linear depsipeptides are rare. The crystal structures and conformational analyses of four depsipeptides with an alternating sequence of an α,α‐disubstituted α‐amino acid and an α‐hydroxy acid are reported. The molecules in the linear hexadepsipeptide amide in (S)‐Pms‐Acp‐(S)‐Pms‐Acp‐(S)‐Pms‐Acp‐NMe2 acetonitrile solvate, C47H58N4O9·C2H3N, ( 3b ), as well as in the related linear tetradepsipeptide amide (S)‐Pms‐Aib‐(S)‐Pms‐Aib‐NMe2, C28H37N3O6, ( 5a ), the diastereoisomeric mixture (S,R)‐Pms‐Acp‐(R,S)‐Pms‐Acp‐NMe2/(R,S)‐Pms‐Acp‐(R,S)‐Pms‐Acp‐NMe2 (1:1), C32H41N3O6, ( 5b ), and (R,S)‐Mns‐Acp‐(S,R)‐Mns‐Acp‐NMe2, C30H37N3O6, ( 5c ) (Pms is phenyllactic acid, Acp is 1‐aminocyclopentanecarboxylic acid and Mns is mandelic acid), generally adopt a β‐turn conformation in the solid state, which is stabilized by intramolecular N—H…O hydrogen bonds. Whereas β‐turns of type I (or I′) are formed in the cases of ( 3b ), ( 5a ) and ( 5b ), which contain phenyllactic acid, the torsion angles for ( 5c ), which incorporates mandelic acid, indicate a β‐turn in between type I and type III. Intermolecular N—H…O and O—H…O hydrogen bonds link the molecules of ( 3a ) and ( 5b ) into extended chains, and those of ( 5a ) and ( 5c ) into two‐dimensional networks.  相似文献   

6.
Crystalline N-hydroxyamino-acid derivatives 4 , readily available from non-chiral acyl chlorides 2 and sultams 1 , were treated with aldehydes in the presence of NaBH3CN to give N-alkylhydroxylamines 5 . N,O-Hydrogenolysis of 5 and saponification of 6 furnished (S)-N-alkylamino acids 7 in high optical purity. Similarly, (R)-N-alkylamino acids 12 were obtained from the antipodal acylsultams 8 .  相似文献   

7.
Pure (E)- or (Z)-enoylsultams 2 were Oxidized with OsO4/N-Methylmorpholine N-oxide in a stereospecific and highly π-face-selective manner. Acetalization of the resulting 1,2-diols furnished, after purification, the stable, crystalline acetals 6 in >99% d.e. and in 63–74% overall yield from 2 . Reductive or hydrolytic cleavage of 6 gave enantiomerically pure alcohols 8 or carboxylic acids 9 with recovery of the sultan auxiliary 1 .  相似文献   

8.
Preparation of Enantiomerically Pure Derivatives of 3-Amino- and 3-Mercaptobutanoic Acid by SN2 Ring Opening of the β-Lactone and a 1,3-Dioxanone Derived from 3-Hydroxybutanoic Acid From (S)-4-methyloxetan-2-one ( 1 ), the β-butyrolactone readily available from the biopolymer ( R )-polyhydroxybutyrate (PHB) and various C, N, O and S nucleophiles, the following compounds are prepared:(s)-2-hydroxy-4-octanone ( 3 ), (R)-3-aminobutanoic acid ( 7 ) and its N-benzyl derivative 5 , (R)-3-azidobutanoic acid ( 6 ) (R)-3-mercaptobutanoic acid ( 10 ), (R)-3-(phenylthio)butanoic acid ( 8 ) and its sulfoxide 9 . The (6R)-2,6-dimethyl-2-ethoxy-1,3-dioxan-4-one ( 4 ) from (R)-3-hydroxybutanoic acid undergoes SN2 ring opening with benzylamine to give the N-benzyl derivative (ent- 5 ) of (S)-3-aminobutanoic acid in 30?40% yield.  相似文献   

9.
A new route to completely protected α-methylated α-amino acids starting from alanine is described (see Scheme). These derivatives, which are obtained via base-catalyzed opening of the oxazolidinones (2S,4R)- and (2R,4S)- 2 , can be directly employed in peptide synthesis. The synthesis of both enantiomers of Z-protected α-methylaspartic acid β-(tert-butyl)ester (O4-(tert-butyl) hydrogen 2-methylaspartates (R) or (S)- 4a ), α-methyl-glutamic acid γ-(tert-butyl) ester (O5-(tert-butyl) hydrogen 2-methylglutamate (R)- or (S)- 4b ), and of Nε-bis-Boc-protected α-methyllysine (N6,N6-bis[(tert-butyloxy)carbonyl]-2-methyllysine (R)- or (S)- 4c ) is described in full detail.  相似文献   

10.
The branched tripodal chloro‐methyl‐siloxanes of the general formula tBuSi[{OSiMe2}yOSiMe3–xClx]3 [x = 0–3; y = 0–2] were synthesized, starting with tert‐Butyl‐trisilanol ( 1 ). The treatment of 1 with the chloro‐methyl‐silanes (Me3–xSiClx+1) (x = 0–3) in the presence of triethylamine leads to the compounds tBuSi(OSiMe2Cl)3 ( 2 ), tBuSi(OSiMeCl2)3 ( 3 ) and tBuSi(OSiCl3)3 ( 4 ). The siloxanes 2 – 4 are colourless oily liquids, which can be purified by distillation. Their yields decrease with the number of chloro substituents. In the reaction of compound 2 with three equivalents of water the silantriol tBuSi(OSiMe2OH)3 ( 5 ) is generated which is used to create the branched tripodal chloro‐methyl‐siloxanes tBuSi(OSiMe2OSiMe3)3 ( 6 ), tBuSi(OSiMe2OSiMe2Cl)3 ( 7 ), tBuSi(OSiMe2OSiMeCl2)3 ( 9 ) and tBuSi(OSiMe2OSiCl3)3 ( 10 ). Compound ( 7 ) is only a side product with a yield of 25 %., The cyclic tBuSi[{(OSiMe2)2Cl}(OSiMe2)3O] ( 8 ) can be isolated and characterised. The transformation of the compound tBuSi(OSiMe2OSiMe2Cl)3 ( 7 ) into the trisilanol tBuSi(OSiMe2OSiMe2OH)3 ( 11 ) allows to prepare the tripodale siloxane tBuSi(OSiMe2OSiMe2OSiMe3)3 ( 12 ) in good yields., The reaction of tBuSi(OSiMe2Cl)3 ( 2 ) with tert‐butyl trisilanol 1 leads to the formation of bicyclic tBuSi(OSiMe2O)3SitBu ( 13 ). An X‐ray structure determination on 13 reveals a [3.3.3]‐bicycle with a C3 axis, which crystallizes in the cubic crystal system in the space group Pa . The reported compounds 2 – 13 were characterised by NMR‐ and IR spectroscopy ( 5 , 11 ) and show correct elemental analyses. The 29Si‐NMR‐data of the compounds show interesting trends with respect to the Si–O chain length and the chloro substistuents.  相似文献   

11.
Treatment of methyl 2-(1-hydroxyalkyl)prop-2-enoates 1 with conc. HBr solution afforded methyl (Z)-2-(bromomethyl)alk-2-enoates 2 , which were transformed regioselectively into N-substituted methyl (E)-2- (aminomethyl)alk-2-enoates 3 (SN2 reaction) and into N-substituted methyl 2-(1-aminoalkyl)prop-2-enoates 4 (SN2′ reaction). Regiocontrol of nucleophilic attack by amine was accomplished simply by choice of solvent, the SN2 reaction occurring in MeCN and the SN2′ reaction in petroleum ether. Hydrolysis and lactamization afforded β-lactams 7 and 8 , containing an exocyciic alkylidene and methylidene group at C(3), respectively.  相似文献   

12.
Naturally occurring (?)-(R,R)-α-necrodol ((?)- 1 ) and its C(4)-epimer (?)- 2 are obtained in 84 and 44% yields, respectively, by lithium ethylenediamide (LEDA) treatment of the corresponding β-necrodols (?)- 3 and (?)- 4 (Scheme 1, Table), both readily available from (?)-campholenyl acetate ((?)- i ) by an efficient stereoselective synthesis. The thermodynamically preferred (?)-(R)-γ-necrodol ((?)- 5 ) becomes the major product (≥ 80% yield) after either prolonged treatment with LEDA or exposure of α- and β-necrodols to BF3·Et2O. In an alternative route, (+)- 5 is prepared starting from (+)-campholenal ((+)- ii ) via Pd-catalysed decarbonylation to (?)-(S)-1,4,5,5-tetramethylcyclopent-l-ene ((?)- 6 ) and subsequent application of an acid-catalysed CH2O-addition/rearrangement sequence (Scheme 2).  相似文献   

13.
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   

14.
Mitsunobu displacement of (−)-(1S,4R,5S,6S)-4,5,6-tris{[(tert-butyl)dimethylsilyl]oxy}cyclohex-2-en-1-ol ((−)- 12 ; a (−)-conduritol-F derivative) with 4-ethyl-7-hydroxy-2H-1-benzopyran-2-one ( 16 ) provided a 5a-carba-β-D -pyranoside (+)- 17 that was converted into (+)-4-ethyl-7-[(1′R,4′R,5′S,6′R)-4′,5′,6′-trihydroxycyclohex-2′-en-1′-yloxy]-2H-1-benzopyran-2-one ((+)- 5 ) and (+)-4-ethyl-7-[(1′R,2′R,3′S,4′R)-2′,3′,4′-trihydroxycyclohexyloxy]-2H-1-benzopyran-2-one ((+)- 6 ). The 5a-carba-β-D -xyloside (+)- 6 was an orally active antithrombotic agent in the rat (venous Wessler's test), but less active than racemic carba-β-xylosides (±)- 5 and (±)- 6 . The 5a-carba-β-L -xyloside (−)- 6 was derived from the enantiomer (+)- 12 and found to be at least 4 times as active as (+)- 6 . (+)-4-Cyanophenyl 5-thio-β-L -xylopyranoside ((+)- 3 ) was synthesized from L -xylose and found to maintain ca. 50% of the antithrombotic activity of its D -enantiomer. Compounds (±)- 5 , (±)- 6 , and (−)- 6 are in vitro substrates for galactosyltransferase 1.  相似文献   

15.
The homoserine and homomethionine sila analogs, (R)‐HOSi(Me2)CH2CH(NH2)CO2H ((R)‐ 21 ) and (R)‐MeSCH2Si(Me2)CH2CH(NH2)CO2H ((R)‐ 24 ), respectively, were each synthesized in nine steps and in 30 and 23% overall yield, respectively, from commercially available ClSiMe2CH2Cl. The key step of both syntheses was the asymmetric α‐bromination of an Evans amide to introduce the stereogenic center of the amino acids with defined absolute configuration. While the preparation of the homomethionine analog (R)‐ 24 followed the expected pathway, the sila analog of homoserine, (R)‐ 21 , was unexpectedly formed during the catalytic hydrogenation of an N3CH2‐substituted silane derivative.  相似文献   

16.
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   

17.
Starting from inexpensive (E)‐β‐farnesene ( 1 ), an eight‐step enantioselective synthesis of the olfactively precious Ambrox® ((?)‐ 2a ) has been performed. The crucial step is the catalytic asymmetric isomerization of (2E,6E)‐N,N‐diethylfarnesylamine ( 3 ) to the corresponding enamine (?)‐(R,E)‐ 4a , applying Takasago's well‐known industrial methodology. The resulting dihydrofarnesal ((+)‐(R)‐ 5 ) (90% yield, 96% ee), obtained after in situ hydrolysis (AcOH, H2O), was then cyclized under catalytic SnCl4 conditions, via its corresponding unreported enol acetate (?)‐(R)‐ 4b , to afford trans‐decalenic aldehyde (+)‐ 6a . Subsequent transformations furnished bicyclic ketone (?)‐ 8a and unsaturated nitrile (+)‐ 11 , both reported as intermediates to access to (?)‐ 2a .  相似文献   

18.
Chiral Building Blocks for Syntheses by Kolbe Electrolysis of Enantiomerically Pure β-Hydroxybutyric-Acid Derivatives. (R)- and (S)-Methyl-, and (R)-Trifluoromethyl-γ-butyrolactones, and -δ-valerolactones The coupling of chiral, non-racemic R* groups by Kolbe electrolysis of carboxylic acids R*COOH is used to prepare compounds with a 1.4- and 1.5-distance of the functional groups. The suitably protected β-hydroxycarboxylic acids (R)- or (S)-3-hydroxybutyric acid, (R)-4,4,4-trifluoro-3-hydroxybutyric acid (as acetates; see 1 – 6 ), and (S)-malic acid (as (2S,5S)-2-(tert-butyl)-5-oxo-1,3-dioxolan-4-acetic acid; see 7 ) are decarboxylatively dimerized or ‘codimerized’ with 2-methylpropanoic acid, with 4-(formylamino)butyric acid, and with monomethyl malonate and succinate. The products formed are derivatives of (R,R)-1,1,1,6,6,6-hexafluoro-2,5-hexanediol (see 8 ), of (R)-5,5,5-trifluoro-4-hydroxypentanoic acid (see 9,10 ), of (R)- and (S)-5-hydroxyhexanoic acid (see 11 ) and its trifluoro analogue (see 12, 13 ), of (S)-2-hydroxy- and (S,S)-2,5-dihydroxyadipic acid (see 23, 20 ), of (S)-2-hydroxy-4-methylpentanoic acid (‘OH-leucine’, see 21 ), and of (S)-2-hydroxy-6-aminohexanoic acid (‘OH-lysine’, see 22 ). Some of these products are further converted to CH3- or CF3-substituted γ- and δ-lactones of (R)- or (S)-configuration ( 14 , 16 – 19 ), or to an enantiomerically pure derivative of (R)-1-hydroxy-2-oxocyclopentane-1-carboxylic acid (see 24 ). Possible uses of these new chiral building blocks for the synthesis of natural products and their CF3 analogues (brefeldin, sulcatol, zearalenone) are discussed. The olfactory properties of (R)- and (S)-δ-caprolactone ( 18 ) are compared with those of (R)-6,6,6-trifluoro-δ-caprolactone ( 19 ).  相似文献   

19.
Selective Amide Cleavage in Peptides Containing α,α-Disubstituted α-Amino Acids A new synthesis of dipeptides with terminal α,α-disubstituted α-amino acids, using 2,2-disubtituted 3-amino-2H-azirines 1 as amino-acid equivalents, is demonstrated. The reaction of 1 with N-protected amino acids leads to the corresponding dipeptide amides in excellent yield. It is shown that the previously described selective hydrolysis (HCl, toluene, 80°, or HCl, MeCN/H2O, 80°) of the terminal amide group results in an extensive epimerization of the second last amino acid. An acid-catalyzed enolization in the intermediate oxazole-5(4H)-ones is responsible for this loss of configurational integrity. In the present paper, a selective hydrolysis of the terminal amide group under very mild conditions is described: In 3N HCl (THF/H2O 1:1), the dipeptide N,N-dimethylamides or N-methytlanilides are hydrolized at 25–35° to the optically pure dipeptides in very good yield.  相似文献   

20.
Biotransformation of (±)‐threo‐7,8‐dihydroxy(7,8‐2H2)tetradecanoic acids (threo‐(7,8‐2H2)‐ 3 ) in Saccharomyces cerevisiae afforded 5,6‐dihydroxy(5,6‐2H2)dodecanoic acids (threo‐(5,6‐2H2)‐ 4 ), which were converted to (5S,6S)‐6‐hydroxy(5,6‐2H2)dodecano‐5‐lactone ((5S,6S)‐(5,6‐2H2)‐ 7 ) with 80% e.e. and (5S,6S)‐5‐hydroxy(5,6‐2H2)dodecano‐6‐lactone ((5S,6S)‐5,6‐2H2)‐ 8 ). Further β‐oxidation of threo‐(5,6‐2H2)‐ 4 yielded 3,4‐dihydroxy(3,4‐2H2)decanoic acids (threo‐(3,4‐2H2)‐ 5 ), which were converted to (3R,4R)‐3‐hydroxy(3,4‐2H2)decano‐4‐lactone ((3R,4R)‐ 9 ) with 44% e.e. and converted to 2H‐labeled decano‐4‐lactones ((4R)‐(3‐2H1)‐ and (4R)‐(2,3‐2H2)‐ 6 ) with 96% e.e. These results were confirmed by experiments in which (±)‐threo‐3,4‐dihydroxy(3,4‐2H2)decanoic acids (threo‐(3,4‐2H2)‐ 5 ) were incubated with yeast. From incubations of methyl (5S,6S)‐ and (5R,6R)‐5,6‐dihydroxy(5,6‐2H2)dodecanoates ((5S,6S)‐ and (5R,6R)‐(5,6‐2H2)‐ 4a ), the (5S,6S)‐enantiomer was identified as the precursor of (4R)‐(3‐2H1)‐ and (2,3‐2H2)‐ 6 ). Therefore, (4R)‐ 6 is synthesized from (3S,4S)‐ 5 by an oxidation/keto acid reduction pathway involving hydrogen transfer from C(4) to C(2). In an analogous experiment, methyl (9S,10S)‐9,10‐dihydroxyoctadecanoate ((9S,10S)‐ 10a ) was metabolized to (3S,4S)‐3,4‐dihydroxydodecanoic acid ((3S,4S)‐ 15 ) and converted to (4R)‐dodecano‐4‐lactone ((4R)‐ 18 ).  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