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 direct aldol reaction between aryl methyl ketones with aromatic aldehydes in the presence of tetrachlorosilane and a catalytic amount of a chiral bithiophene diphosphine oxide was studied; the product of double aldol addition was isolated as diacetate in good diastereoselectivity (up to 95:5) and enantioselectivities up to 91%. The reaction with heteroaromatic aldehydes was also investigated leading to the corresponding 1,3 diols, in some cases with excellent stereoselectivities. Chirality 25:643–647:, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

2.
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   

3.
The haloalkane 1,2-dibromo-3-chloropropane (DBCP), an environmental pollutant that was widely used as a soil fumigant, is a carcinogen and a mutagen and displays target-organ toxicity to the testes and the kidneys. Because little is known about effects of stereochemistry on the metabolism and toxicity of halogenated alkyl compounds and because DBCP, which has a chiral center at C-2, may show enantioselectivity in its metabolism and/or toxicities, the optically pure enantiomers of DBCP were tested in vivo in rats for organ toxicity as well as for bacterial mutagenicity. Organ toxicity studies showed that (S)-DBCP was slightly more renal toxic than (R)-DBCP but was not significantly more toxic than the racemate, and that no significant differences were observed in the extents of testicular necrosis and atrophy caused by either enantiomer or the racemate. In contrast, (R)-DBCP was more mutagenic than either (S)-DBCP or the racemate to Salmonella typhimurium (S. typhimurium) strains TA 100 and TA104. However, there was little or no enantioselectivity in glutathione S-transferase (GST)-catalyzed conjugation reactions of glutathione with DBCP based on the lack of selectivity in the rates of disappearance of the enantiomers of DBCP in the presence of glutathione (GSH) and GSTs as monitored by chiral gas chromatography (GC). © 1995 Wiley-Liss, Inc.  相似文献   

4.
The direct catalytic α‐amidoalkylation of dihydroquinolines with aldehydes bearing oxygen functionalities at different positions in a Mannich‐type reaction has been studied. β‐Alkoxy‐aldehyde 1d gave high enantioselectivity, albeit with an inherently poor diastereoselectivity, while the use of α‐alkoxy aldehydes 1c was detrimental also to enantioselectivity. Mannich‐type reactions have been studied for the first time using new chiral carbohydrate‐derived aldehydes 1a,b showing a reactivity markedly influenced by the presence of water. The chiral glycidic backbone showed a slight but significant influence on the overall stereochemical outcome only when present in α‐position of the aldehyde. The absolute stereochemistry of the products was studied by electronic circular dichroism (ECD) spectra and compared with theoretical calculations. ECD analysis easily provides the absolute configuration of 1,2‐dihydroquinoline derivatives such as quinoline‐1(2H)‐carboxylates.  相似文献   

5.
Chiral β-hydroxy α-amino acid structural motifs are interesting and common synthons present in multiple APIs and drug candidates. To access these chiral building blocks either multistep chemical syntheses are required or the application of threonine aldolases, which catalyze aldol reactions between an aldehyde and glycine. Bioinformatics tools have been utilized to identify the gene encoding threonine aldolase from Vanrija humicola and subsequent preparation of its recombinant version from E. coli fermentation. We planned to implement this enzyme as a key step to access the synthesis of our target API. Beyond this specific application, the aldolase was purified, characterized and the substrate scope of this enzyme further investigated. A number of enzymatic reactions were scaled-up and the products recovered to assess the diastereoselectivity and scalability of this asymmetric synthetic approach towards β-hydroxy α-amino acid chiral building blocks.  相似文献   

6.
Some 3-t-butyldimethylsilyloxy derivatives, synthesized from the cheap commercially available (1S,2S)-2-amino-1-phenyl-1,3-propanediol [(1S,2S)- 1 ], have been successfully employed as new chiral ligands in the asymmetric Reformatsky reaction on aldehydic substrates. The influence both of the substrate and of the ligand on the stereochemical pathway has been investigated by varying the structure of the carbonyl substrate and of the optically active aminodiols. © 1995 Wiley-Liss, Inc.  相似文献   

7.
Alkyne is a useful functionality incorporated in proteins for site-selective bioconjugation reactions. Although effective bioconjugation reactions such as copper(I)-catalyzed and/or copper-free 1,3-dipolar cycloadditions of alkynes and azides are the most common approaches, the development of new alkyne-based bioconjugation reactions is still an ongoing interest in chemical biology. In this work, a new approach has been developed for selective modification of alkyne-linked peptides and proteins through the formation of arylacetylenes by a cross-coupling reaction of 6-membered ring cyclometalated gold(III) (C^N) complexes (HC^N = 2-arylpyridines) with terminal alkynes. Screening of the reaction conditions with a series of cyclometalated gold(III) complexes with phenylacetylene gave an excellent yield (up to 82%) by conducting the reaction in slightly alkaline aqueous conditions. The reaction scope was expanded to various alkynes, including alkyne-linked peptides to achieve up to >99% conversion. Using fluorescent dansyl (1l) and BODIPY (1m)-linked gold(III) complexes, alkyne-linked lysozyme has been selectively modified.  相似文献   

8.
A series of chiral Ag(I) and Cu(II) complexes have been prepared from the reaction between AgX (X = NO3, PF6, OTf) or CuX2 (X = Cl, ClO4) and chiral biaryl-based N-ligands. The rigidity of the ligand plays an important role in the Ag(I) complex formation. For example, treatment of chiral N3-ligands 1-3 with half equiv of AgX (X = NO3, PF6, OTf) gives the chiral bis-ligated four-coordinated Ag(I) complexes, while ligand 4 affords the two-coordinated Ag(I) complexes. Reaction of AgX with 1 equiv of chiral N4-ligands 5, 7, 8 and 10 gives the chiral, binuclear double helicate Ag(I) complexes, while chiral mono-nuclear single helicate Ag(I) complexes are obtained with N4-ligands 6 and 9. Treatment of either N3-ligand 1 or N4-ligand 9 or 10 with 1 equiv of CuX2 (X = Cl, ClO4) gives the mono-ligated Cu(II) complexes. All the complexes have been characterized by various spectroscopic techniques, and elemental analyses. Seventeen of them have further been confirmed by X-ray diffraction analyses. The Cu(II) complexes do not show catalytic activity for allylation reaction, in contrast to Ag(I) complexes, but they do exhibit catalytic activity for Henry reaction (nitroaldol reaction) that Ag(I) complexes do not.  相似文献   

9.
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   

10.
Four novel chiral amino alcohols were synthesized from D‐(+)‐camphor and utilized as ligands in a Cu(I)‐catalyzed asymmetric Henry reaction. The reactions were carried out under mild conditions with excellent enantioselectivities and moderate yields without the exclusion of air or moisture. The highest enantioselectivity was observed up to 94% enantiomeric excess (ee) with ligand L1 in toluene at room temperature. Chirality 27:761–765, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

11.
This review provides a comprehensive treatment of literature data dealing with asymmetric synthesis of α-amino-β-hydroxy and α,β-diamino acids via homologation of chiral Ni(II) complexes of glycine Schiff bases using aldol and Mannich-type reactions. These reactions proceed with synthetically useful chemical yields and thermodynamically controlled stereoselectivity and allow direct introduction of two stereogenic centers in a single operation with predictable stereochemical outcome. Furthermore, new application of Ni(II) complexes of α-amino acids Schiff bases for deracemization of racemic α-amino acids and (S) to (R) interconversion providing additional synthetic opportunities for preparation of enantiomerically pure α-amino acids, is also reviewed. Origin of observed diastereo-/enantioselectivity in the aldol, Mannich-type and deracemization reactions, generality and limitations of these methodologies are critically discussed.  相似文献   

12.
An enzymatic reduction of 2-substituted cyclohexanones mediated by Saccharomyces cerevisiae was studied with respect to the stereochemical course and optical purity of the products. Reduction of ketones 1b-1f resulted in separable diastereoisomeric mixtures of cis- and trans-stereoisomers of 2-substituted cyclohexanols (2b-2f and 3b-3f) having the (S) absolute configuration at the chiral center bearing the hydroxyl functionality with high enantiomeric purity. Reduction of ketone 1a yielded mixture of cis-(1S, 2R)- and trans-(1R, 2R)-stereoisomers (2a and 3a) with lower enantiomeric purity. Changes in the nature of the C(2)-substituent affect the stereochemical course of the biotransformation. However, they significantly influenced the enantiomeric purity of the products. The diastereoselectivity of the process was studied as well; high diastereoselectivity was observed with the substrates 1a, 1e and 1f.  相似文献   

13.
Youn SW  Choi JY  Kim YH 《Chirality》2000,12(5-6):404-407
Chiral hydrazones 1 reacted with aryl- or alkyl-lithiums at -78 degrees C in a short reaction time, within 10 min, to afford arylated or alkylated chiral hydrazines 3 with extremely high diastereoselectivity (up to >99% de) and high chemical yields. The hydrazines are readily converted to chiral amino alcohols 4.  相似文献   

14.
The diastereoselective [2+2] photocycloaddition of cyclohexenonecarboxylates containing various chiral auxiliaries to ethylene is described. The effect of the auxiliary, reaction temperature, and solvent on diastereoselectivity was examined. The (?)‐8‐(p‐methoxyphenyl)menthyl group was found to be the most effective chiral auxiliary. The photoreaction of (?)‐8‐(p‐methoxyphenyl)menthyl cyclohexenonecarboxylate in methylcyclohexane at ?78°C gave the corresponding bicyclo[4.2.0]octanone derivative in 81% diastereomeric excess (d.e.). The extent of diastereoselectivity was found to be closely related to the most stable π‐stack conformation of the starting cyclohexenones. Chirality 15:504–509, 2003. © 2003 Wiley‐Liss, Inc.  相似文献   

15.
The first, target-oriented synthesis of pyranoquinolines as potential antibacterial agents by inverse electron demand (IED) Diels-Alder reaction using chiral salen-AlCl complex has been accomplished, the reactions proceed with moderate yields, and a very high degree of diastereoselectivity (>90%). The diastereoselectivity-enhanced pyranoquinolines exhibit potential bactericidal activity against seven strains of pathogenic gram-negative bacteria.  相似文献   

16.
P,C‐Stereogenic α‐amino phosphine oxides were prepared from the addition of (RP)‐menthyl phenyl phosphine oxide to chiral aldimines under neat condition at 80 °C in up to 91:9 drC and 99% yields. The diastereoselectivity was mainly induced by chiral phosphorus that showed matched or mismatched induction with (S)‐ or (R)‐aldimines, respectively. Chirality 28:132–135, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

17.
《Chirality》2017,29(5):213-220
Optically pure, diastereomeric aziridine amides built on the chiral skeletons of camphor, fenchone, and menthone have proven to be highly efficient ligands for enantioselective asymmetric direct aldol reaction in the presence of water and zinc triflate. Desired products were formed in moderate to high chemical yields (up to 95%) and with enantiomeric excess up to 99%. The influence of the stereogenic centers located at the aziridine subunit on the stereochemical course of the reaction is discussed.  相似文献   

18.
The ability of chiral β‐amino alcohols to catalyze the direct asymmetric aldol reaction was evaluated for the first time in aqueous micellar media. A family of cheap and easily accessible β‐amino alcohols, obtained in one step from naturally occurring amino acids, was shown to successfully catalyze the asymmetric aldol reaction between a series of ketones and aromatic aldehydes. These aldol reactions furnished the corresponding β‐hydroxy ketones with up to 93% isolated yield and 89% ee. (S)‐2‐phenylglycinol and Triton X‐100 proved to be the best organocatalyst and surfactant, respectively. Chirality 25:119–125, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

19.
Coordination of a chiral substrate to (meso‐salen)cobalt(II) nitrate and subsequent oxidation generates a Co(III) complex exhibiting a strong chiroptical readout that is attributed to spontaneous substrate‐to‐ligand chirality imprinting. The characteristic circular dichroism (CD) response of the (salen)cobalt complex can be used for enantiomeric analysis of a variety of chiral substrates based on a simple CD measurement at low concentration and without additional purification steps. This chirality sensing approach has potential for high‐throughput enantiomeric excess (ee) screening applications and minimizes solvent waste production. Chirality 26:379–384, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
Under neutral conditions, spontaneous mirror symmetry breaking has been occasionally reported for aldol reactions starting from achiral reagents and conditions. Chiral induction might be interpreted in terms of autocatalysis exerted by chiral mono‐aldol or bis‐aldol products as source of initial enantiomeric excesses, which may account for such experimental observations. We describe here a thorough Density Functional Theory (DFT) study on this complex and otherwise difficult problem, which provides some insights into this phenomenon. The picture adds further rationale to an in‐depth analysis by Moyano et al, who showed the isolation and characterization of bis‐aldol adducts and their participation in a complex network of reversible steps. However, the lack of enantiodiscrimination (ees vanish rapidly in solution) suggests, according to the present results, a weak association in complexes formed by the catalysts and substrates. The latter would also be consistent with almost flat transition states having similar heights for competitive catalyst‐bound transition structures (actually, we were unable to locate them at the level explored). Overall, neither autocatalysis as once conjectured nor mutual inhibition of enantiomers appears to be operating mechanisms. Asymmetric amplification in early stages harnessing unavoidable enantiomeric imbalances in reaction mixtures of chiral products represents a plausible interpretation.  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