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.
A new and convenient stereocontrolled synthesis of the optically pure (S)-α-methyl,α-amino acids 6(ad) that exploits the chiral synthon 1,4-N,N-[(S)-1-phenylethyl]-piperazine-2,5-dione (1) is described. The (S)-1-phenylethyl group, bonded to each of the N-atoms of the 2,5-diketopiperazine, acts as a chiral inductor in the first alkylation, while the steric hindrance appears to be the determining factor of stereocontrol in third and forth alkylation.  相似文献   

2.
Summary.  N-[(Z)-N-Benzoyl- or N-boc-(2-fluorophenyl)dehydroalanyl]-(R)- or (S)-phenyl-alanines 1,2,5 and 6 were hydrogenated in the presence of chiral and achiral rhodium complexes. The optical induction is compared to the results obtained using the corresponding esters as substrates. Received November 11, 2001 Accepted January 15, 2002  相似文献   

3.
A neodymium-(S)-PDTA (PDTA = N,N,N′,N′-tetrakis[(hydroxycarbonyl)methyl]-1,2-diaminopropane) complex was found exceptionally useful in the quantitative determination of enantiomer ratios of water-soluble natural amino acids by 13C-NMR. The method is demonstrated on mixtures of l- and d-enantiomers of various amino acids. The interactions of the chiral shift reagent with the amino acid molecules were rationalized by molecular orbital calculations.  相似文献   

4.
Salicylidenimine palladium(II) complexes trans‐Pd(O,N)2 adopt step and bowl arrangements. A stereochemical analysis subdivides 52 compounds into 41 step and 11 bowl types. Step complexes with chiral N‐substituents and all the bowl complexes induce chiral distortions in the square planar system, resulting in Δ/Λ configuration of the Pd(O,N)2 unit. In complexes 1 , 2 , 3 , 4 , 5 , 6 with enantiomerically pure N‐substituents ligand chirality entails a specific square chirality and only one diastereomer assembles in the lattice. Dimeric Pd(O,N)2 complexes with bridging N‐substituents in trans‐arrangement are inherently chiral. For dimers 7 , 8 , 9 , 10 , 11 different chirality patterns for the Pd(O,N)2 square are observed. The crystals contain racemates of enantiomers. In complex 12 two independent molecules form a tight pair. The (RC) configuration of the ligand induces the same Δ chirality in the Pd(O,N)2 units of both molecules with varying square chirality due to the different crystallographic location of the independent molecules. In complexes 13 and 14 atrop isomerism induces specific configurations in the Pd(O,N)2 bowl systems. The square chirality is largest for complex 15 [(Diop)Rh(PPh3)Cl)], a catalyst for enantioselective hydrogenation. In the lattice of 15 two diastereomers with the same (RC,RC) configuration in the ligand Diop but opposite Δ and Λ square configurations co‐crystallize, a rare phenomenon in stereochemistry. Chirality 25:663–667, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

5.
Abstract

The chiral synthesis of (1S,3S,4S)-1-(3,4-dihydroxycyclopent-1-yl)-1H?thymine (carbocyclic 5′-nor thymidine, 4) has been achieved in 5 steps from (+)-(lR,4S)-4-hydroxy-2-cyclopenten-1-yl acetate (5) and N3?benzoylthymine. Compound 4 is viewed as a monomeric building block for poly-T-like oligomers.  相似文献   

6.
Over last decade, the use of Ni(II) complexes, derived from of glycine Schiff bases with chiral tridentate ligands, has emerge as a leading methodology for preparation of structurally diverse Tailor-Made Amino Acids, the key structural units in modern medicinal chemistry, and drug design. Here, we report asymmetric synthesis of derivatives of (S)-α-(octyl)glycine ((S)-2-aminodecanoic acid) and its N-Fmoc derivative via alkylation of chiral nucleophilic glycine equivalent with n-octyl bromide. Under the optimized conditions, the alkylation proceeds with excellent yield (98.1%) and diastereoselectivity (98.8% de). The observed stereochemical outcome and convenient reaction conditions bode well for application of this method for large-scale asymmetric synthesis of (S)-2-aminodecanoic acid and its derivatives.  相似文献   

7.
The stereoisomers of α,α′-bis[3-(N,N-diethylcarbamoyl)-piperidino]-p-xylene ( 1 ) were synthesized. Rac ethyl nipecotate was resolved by diastereomeric (-)-D - and (+)-L-tartrate salt formation. The enantiomeric esters were hydrolyzed to the corresponding nipecotic acids, which were then converted into t-BOC derivatives. Treatment of the latter with diethylamine/isobutyl chloroformate and removal of the t-BOC protecting group afforded (R)- and (S)-N,N-diethylnipecotamides. Condensation of the latter with α,α′-dibromo-p-xylene gave (R,R)- and (S,S)- 1 . The meso-diastereomer was obtained by stereospecific synthesis in addition to our earlier procedure involving fractional crystallization of the diastereomeric mixture obtained by synthesis. The latter was resolved earlier into 1A , 1B , and 1C using chiral high-performance liquid chromatography (HPLC). Based on the stereospecific synthesis now achieved, 1A and 1B are assigned the configurations, (R,R) and (S,S) respectively, and 1C is assigned the meso configuration. The (R,S) structure of the latter is also confirmed by X-ray crystallography. © 1995 Wiley-Liss, Inc.  相似文献   

8.
The enantioselectivity potential of five strains of Saccharomyces cerevisiae was studied for the reduction of ethyl N-{2-{4-[(2-oxocyclohexyl)methyl]phenoxy}ethyl} carbamate (1), an insect juvenile hormone bioanalog. The products of the reaction, the cis and trans isomers of ethyl N-{2-{4-[(2-hydroxycyclohexyl)methyl]phenoxy}ethyl} carbamate (2 and 3), were obtained in 45–49% (w/w) chemical yields and with 79 to > 99% enantiomeric purity values. The absolute configurations of the major products were assigned as ethyl (1S,2S)-N-{2-{4-[(2-hydroxycyclohexyl)methyl]phenoxy}ethyl} carbamate (2) and ethyl (1S,2R)-N-{2-{4-[(2-hydroxycyclohexyl)methyl]phenoxy}ethyl} carbamate (3). The products 2 and 3 belong to the series of the chiral insect juvenile hormone analogs.  相似文献   

9.
Enantiopure 3((R)‐ and 3((S)‐1‐phenylethyl)‐4‐oxazoline‐2‐ones were evaluated as chiral building blocks for the divergent construction of heterocycles with stereogenic quaternary centers. The N‐(R)‐ or N‐(S)‐1‐phenylethyl group of these compounds proved to be an efficient chiral auxiliary for the asymmetric induction of the 4‐ and 5‐positions of the 4‐oxazolin‐2‐one ring through thermal and MW‐promoted nucleophilic conjugated addition to Michael acceptors and alkyl halides. The resulting adducts were transformed via a cascade process into fused six‐membered carbo‐ and heterocycles. The structure of the reaction products depended on the electrophiles and reaction conditions used. Alternative isomeric 4‐methylene‐2‐oxazolidinones served as chiral precursors for a versatile and divergent approach to highly substituted cyclic carbamates. DFT quantum calculations showed that the formation of bicyclic pyranyl compounds was generated by a diastereoselective concerted hetero‐Diels‐Alder cycloaddition.  相似文献   

10.
A series of axially chiral ethers synthesized from biscarboline N,N′‐dioxides, (S)‐ 1a to (S)‐ 1n , was investigated in enantioselectivity addition reactions of allyltrichlorosilane with a series of substituted aldehydes, including bulky substituted aldehydes. High enantioselectivities (up to 96%ee) were achieved using the catalyst (S)‐ 1k at 1 mol % loading.  相似文献   

11.
New efficient catalysts based on electrophilic N‐fluoro quaternary ammonium salts are reported for catalytic allylation of (E)‐N,1‐diphenylmethanimine. The chiral version of the catalyst based on cinchonidine (F‐CD‐BF4) shows high catalytic activity with approximately 94% ee and TOF (>800 h?1). The F‐CD‐BF4 is prepared from cinchonidine and Selectfluor by one‐step transfer fluorination.  相似文献   

12.
An efficient approach to the asymmetric syntheses of α-methylglutamic acid and α-methylornithine is described. Two chiral reagents, (2′S)-N-(2′-methoxymethylpyrrolidine)-2-isocyanopropionamide 4 and (2′S)-N-(2′-hydroxymethylpyrrolidine)-2-isocyanopropionamide 5, were employed for the asymmetric induction. α-Methylglutamic acid 7 was synthesized by the asymmetric Michael-addition of methyl acrylate to 4 and 5 as the key step. The optical yield of 7 was 10~45% (R-form). α-Methylornithine 12 was also synthesized by the reaction of 4 with acrylonitrile as the key step. The optical yield of 12 was 31.7% (R-form).  相似文献   

13.
Two new chiral monochloro-s-triazines (MCT) were synthesized [viz N-(4-chloro-6-piperidinyl-[1,3,5]-triazine-2-yl)-l-leucine amide and N-(4-chloro-6-piperidinyl-[1,3,5]-triazine-2-yl)-l-leucine) (CDR 1 and 2, respectively)] by the nucleophilic displacement of chlorine atoms in s-triazine moiety. One of the Cl atoms was replaced with piperidine, and the second Cl atom in the 6-piperidinyl derivative was replaced with amino acid amide (viz l-Leu–NH2) and amino acid (l-Leu). These reagents were characterized and used as CDRs for chiral separation of protein and non-protein amino acids, and were separated on a reversed-phase C18 column. The reaction conditions were optimized for the synthesis of diastereomers using one MCT reagent. The separation method was validated for limit of detection, linearity, accuracy, precision, and recovery.  相似文献   

14.
The lipase-catalyzed kinetic resolution of trans- and cis-2-azidocycloalkanols and the preparation of enantiomerically pure trans- and cis-2-aminocycloalkanols are described.

Four kinds of lipases were screened for the acetylation of trans- and cis-2-azidocycloalkanols. Among them, Pseudomonas sp. lipases (lipase PS and lipase AK, Amamo Pharmaceutical Co.) showed the highest enantioselectivity. These products were converted to the corresponding 2-aminocycloalkanols to determine their enantiomeric excess (ee) and absolute configurations by HPLC and CD analyses, using (S)-TBMB carboxylic acid [(S)-2-tert-butyl-2-methyl-1,3-benzodioxole-4-carboxylic acid] as the chiral conversion reagent. The results of the CD analysis proved N,O-bis-(S)-TBMB carboxylated cis-2-aminocycloalkanols to adopt a predominantly N-equatorial conformation.

The partially resolved trans- and cis-2-aminocycloalkanols, except for trans-2-aminocyclopentanol, were recrystallized from ethyl acetate to give enantiomerically pure forms.  相似文献   

15.
Our interest on engineering non-ribosomal synthetase responsible for SW-163 biosynthesis prompted us to determine the relative and absolute configuration of antitumor cyclic depsipeptide SW-163s. We first isolated and identified SW-163 homologs D, F and G as known compounds UK-63598, UK-65662 and UK-63052, respectively. Both enantiomers of the unusual constitutive amino acid, N-methylnorcoromic acid, were synthesized in chiral forms starting from (R)- and (S)-1,2-propanediol. The hydrolyzate of SW-163D, a major constituent of this family, was converted with Marfey’s reagent, 1-fluoro-2,4-dinitrophenyl-5-L-alanine-amide (L-FDAA), and the resulting mixture of amino acid derivatives was subjected to an LC/MS analysis. Compared with authentic samples, the analytical data unambiguously show that SW-163D consisted of L-Ala, D-Ser and (1S, 2S)-N-methylnorcoronamic acid. The remaining stereochemistry of the N-methylcysteine moieties was determined from NOE data.  相似文献   

16.
Axial chirality in N,N-dimethylaminopyridines as well as N,N-dipropylaminopyridines bearing an internal carboxy group were evaluated based on their racemization barriers and circular dichroism spectra. The half-life of racemization of N,N-dipropylaminopyridine derivative 2 was estimated to be 19.7 days at 20°C. Its enantiomers isolated as optically active forms showed positive-negative and negative-positive Cotton effects for (+)- 2 and (−)- 2 , respectively, from 310 to 210 nm. Furthermore, (−)- 2 was applied as a chiral nucleophilic catalyst and exhibited asymmetric induction in acylative kinetic resolution of 1-(1-naphthyl)ethane-1-ol.  相似文献   

17.
In this study, an optically active diamine, N,N′-(pyromellitoyl)-bis{N-[4(4-aminophenoxy)phenyl]-2-(4-methyl)pentanamide} (1) containing amino acid l-leucine was prepared in three steps. The step-growth polymerization of this chiral diamine with several diisocyanates in room temperature ionic liquid (IL), 1,3-dipropylimidazolium bromide as an environmentally friendly solvent and in a volatile organic solvent, is investigated. The polymerization yields and inherent viscosities of the resulting poly(amide-ether-imide-urea)s are compared in both solvents. The results show that the IL to be the superior polymerization media. All of the obtained polymers exhibited good solubility in some polar aprotic organic solvents such as N,N-dimethyacetamide, N,N-dimethyformamide, dimethyl sulfoxide while thermal stability was not disturbed based on thermogravimetric analysis and differential scanning calorimetry experiments. X-ray diffraction analysis of polymers shows that they are amorphous. The observation of optical rotation confirms the optical activity of prepared polymers.  相似文献   

18.
The N-(n-butylamide) of (S)-2-(phenylcarbamoyloxy)propionic acid, easily prepared starting from the inexpensive L -ethyl lactate, can be used as convenient chiral solvating agent (CSA) to determine the enantiomeric composition of N-(3,5-dinitrobenzoyl)amino acid methyl esters.  相似文献   

19.
Polyhydroxyalkanoates (PHA) are a family of polyesters consisting of over 150 chiral hydroxyalkanoic acids (HA). This paper reviews the physiological functions of (R)-3-hydroxybutyric acid (3HB) and (R)-4-hydroxybutyric acid and summarizes the technologies developed to produce various HA [3HB, (R)-3-hydroxyoctanoic acid, (R)-3-hydroxydecanoic acid, etc.] and the applications of chiral HA. Their outlooks and perspectives are discussed.  相似文献   

20.
The achiral hydrocarbon tetraphenylethylene crystallizes in enantiomorphous forms (chiral space group: P21) to afford right- and left-handed hemihedral crystals, which can be recognized by solid-state circular dichroism spectroscopic analysis. Chiral organic crystals of tetraphenylethylene mediated enantioselective addition of diisopropylzinc to pyrimidine-5-carbaldehyde to give, in conjunction with asymmetric autocatalysis with amplification of chirality, almost enantiomerically pure (S)- and (R)-5-pyrimidyl alkanols whose absolute configurations were controlled efficiently by the crystalline chirality of the tetraphenylethylene substrate. Tetrakis(p-chlorophenyl)ethylene and tetrakis(p-bromophenyl)ethylene also show chirality in the crystalline state, which can also act as a chiral substrate and induce enantioselectivity of diisopropylzinc addition to pyrimidine-5-carbaldehyde in asymmetric autocatalysis to give enantiomerically enriched 5-pyrimidyl alkanols with the absolute configuration correlated with that of the chiral crystals. Highly enantioselective synthesis has been achieved using chiral crystals composed of achiral hydrocarbons, tetraphenylethylenes, as chiral inducers. This chemical system enables significant amplification of the amount of chirality using spontaneously formed chiral crystals of achiral organic compounds as the seed for the chirality of asymmetric autocatalysis.  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