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 25 毫秒
1.
The effects of reaction solvent on inclusion complexation of a chiral dipeptide (3S,6S)‐ 1 derived from (S)‐proline toward racemic BINOL was investigated, discovering that the reaction solvent played a crucial role in determining the inclusion complexation behavior of dipeptide (3S,6S)‐ 1 toward rac‐BINOL. (3S,6S)‐ 1 did not show any chiroselective or achiroselective complexation toward rac‐BINOL in polar protic solvents such as methanol and ethanol, polar aprotic solvents including trichloromethane and THF, while in polar aprotic solvent ethyl acetate and apolar aprotic solvents benzene, (3S,6S)‐ 1 displayed achiroselective complexation toward rac‐BINOL. However, the resulting heterocomplex HC‐ 2 from benzene and HC‐ 3 from ethyl acetate have a different composition. Single crystal X‐ray diffraction analysis demonstrates that the two heterocomplexes are formed via different H‐bond interaction patterns, in which the reaction solvent has a dramatic effect. Furthermore, this work provides a relatively green method for quantitative enantiomeric enrichment of nonracemic BINOL, in which unacceptable and toxic benzene was replaced by ethyl acetate.  相似文献   

2.
Mikami K  Ueki M  Matsumoto Y  Terada M 《Chirality》2001,13(9):541-544
Chiral tetranuclear Ti cluster, a cubic structure constituted of four Ti atoms and OHs, and six (R)-binaphthols (BINOL) bridged two Ti atoms as ligands, is shown to be a novel chiral Lewis acid catalyst for the [2+3] cycloaddition reaction with nitrones. The chiral Ti clusters with 7,7'-substituted (R)-BINOL ligands was synthesized to give enhanced enantiomeric excesses up to 78% ee.  相似文献   

3.
A series of (S)‐BINOL ligands substituted at the 3 position with some five‐membered nitrogen‐containing aromatic heterocycles were effectively prepared and their catalytic abilities were evaluated in the asymmetric addition of diethylzinc to benzaldehyde in the presence of titanium tetraisopropoxide. Under the optimized reaction conditions, titanium complex of (S)‐3‐(1H‐benzimidazol‐1‐yl)‐1,1′‐bi‐2‐naphthol was found to be the most efficient catalyst for asymmetric ethylation of various aldehydes to generate the corresponding secondary alcohols in up to 99% yield and 91% ee. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
Four new chiral pincer-complexes were prepared based on coupling of BINOL and TADDOL moieties with iodoresorcinol followed by oxidative addition of palladium(0). The X-ray analysis of complex 5a revealed that the BINOL rings form a well-defined chiral pocket around the palladium atom. This chiral environment can be further modified by γ-substitution of the BINOL rings. Preliminary studies for electrophilic allylation of sulfonimine 2 with allylstannane revealed that the presented chiral complexes are promising asymmetric catalysts for preparation of chiral homoallyl amines. The best result was achieved employing catalytic amounts of γ-Me BINOL complex 6 affording homoallyl amine 4 with 59% ee and 74% isolated yield.  相似文献   

5.
The reaction of MoO2(mes)2 with S-H2BINOL (mes = 2,4,6-Me3C6H2; H2BINOL = 1,1′-bi-2-naphthol) and (CF3)2MeCOH in THF yielded the novel dioxo-molybdenum(VI) complexes MoO2(S-BINOL)(THF)2 and MoO2[OCMe(CF3)2]2(THF), respectively. Similar tungsten derivatives WO2(S-BINOL)(THF) and WO2[OCMe(CF3)2]2(THF) have been prepared by the reaction of WO2Cl2(DME) with the corresponding lithium salts of BINOL and 1,1,1,3,3,3-hexafluoro-2-methyl-2-propanol, respectively. Catalytic experiments have shown that MoO2(S-BINOL)(THF)2 is an active catalyst in the sulfoxidation of methyl phenyl sulfide and in the epoxidation of cis-cyclooctene with tert-butylhydroperoxide, under mild conditions. The BINOL complex was, however, not found to be enantioselective. In addition, the catalytic activity of the molybdenum species MoO2(S-BINOL)(THF)2 and MoO2[OCMe(CF3)2]2(THF) in the hydrosilylation of carbonyl groups has been explored.  相似文献   

6.
Chen YX  Yang LW  Li YM  Zhou ZY  Lam KH  Chan AS  Kwong HL 《Chirality》2000,12(5-6):510-513
A new chiral ligand 6, 6'-dihydroxy-5, 5'-biquinoline (BIQOL, 2) was prepared via Cu2+ mediated coupling. The resolution was carried out by separating the corresponding ditrifluomethanesulfonate on chiral column. When applied to the enantioselective addition of diethylzinc to aromatic aldehydes, this ligand induced the reaction with enantioselectivity equivalent to that induced by BINOL. The effects of solvent and reaction temperature on enantioselectivity were also studied.  相似文献   

7.
《Chirality》2017,29(7):340-347
Acrylamide (AM) was copolymerized with ethylene glycol dimethacrylate (EGDMA) in the presence of (R )‐1,1′‐binaphthalene‐2‐naphthol (BINOL) as the template molecules on the surface of silica gel by a free radical polymerization to produce a chiral stationary phase based on the surface molecularly imprinted polymer (SMIP‐CSP). The SMIP‐CSP showed a much better separation factor (α = 4.28) than the CSP based on the molecularly imprinted polymer (MIP‐CSP) without coating on the silica gel (α = 1.96) during the chiral separation of BINOL enantiomers by high‐performance liquid chromatography. The influence of the pretreatment temperature and the content of the template molecule ((R )‐BINOL) of the SMIP‐CSP, and the mobile phase composition on the separation of the racemic BINOL were systematically investigated.  相似文献   

8.
(S)‐6‐Br‐BINOL‐derived phosphoramidite, a s imple monodentate ligand with a stereogenic center at the phosphorus atom, was synthesized for the first time. This stereoselector generated a high level of enantioselectivity (80–95% ee) in the rhodium‐catalyzed hydrogenation of α‐dehydrocarboxylic acid esters and was also successfully employed in the asymmetric palladium‐catalyzed allylic substitution of (E)‐1,3‐diphenylallyl acetate. The optical yield also showed significant dependence with reaction type: up to 70% ee for allylic amination, up to 75% ee for allylic sulfonylation, and up to 90% ee for allylic alkylation. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

9.
A simple chiral analysis of amino acid esters by fluorine‐19 nuclear magnetic resonance (19F NMR) through the modified James–Bull method is described. Thus, amino acid ester acid salt was treated with 5‐fluoro‐2‐formylphenylboronic acid and (S)‐BINOL in the presence of triethylamine (TEA) and MS4A for 10 minutes. The reaction mixture was analysed by 19F NMR directly to afford good quantifications.  相似文献   

10.
Using chiral BINOL‐derived phosphoric acids (PA's) to activate substrates for enhanced reactivity is now regarded as a powerful strategy to control enantioselectivity in asymmetric synthesis. Generally, most substituents at the 3,3′‐positions of BINOL PA's are aryl derivatives. These derivatives are pivotal in attaining high selectivity. PA's with alkyl substituents in these positions have rarely been reported. Herein, we introduced alkyl‐based substituents at the 3,3′‐position of PA's. These new potential catalysts, if applied in reactions, may allow altered noncovalent interactions (as opposed to the typical aryl substituents in these positions) with substrates used in chiral PA‐catalyzed chemistry in the future.  相似文献   

11.
《Inorganica chimica acta》2004,357(13):3854-3864
This paper reports the synthesis of a polymerizable dppe-derivative with eight pendent styrenyl units linked to the dppe core through four tyrosine residues. This diphosphine ligand was utilized in the synthesis of a series of P2PdX2 complexes (X2=(R)-BINOL, (S)-BINOL, Cl2 and π-1,3-Ph2-allyl+). The compounds were used as comonomers for the synthesis of porous organic polymers (poly EDMA). Molecular imprinting effects on X2-ligand removal were investigated. Overall, the presence of a cavity of significant size was found to be beneficial to the rate of the allylic alkylation reaction, however, chiral BINOL shaped cavities did not influence the enantio-selectivity of the reaction.  相似文献   

12.
A convenient enantioselective approach for the pinacol coupling of 1‐acetylazulene involving easily accessible (R)‐ or (S)‐BINOLs as chiral additive is reported. This supposes the preformation of the chiral titanium–BINOL complex in 1:2 ratio and subsequent reduction with zinc when, 2,3‐di(azulen‐1‐yl)butane‐2,3‐diol can be isolated in around 60% enantiomeric excess. The absolute configuration of the isolated enantiomers was assigned by comparison of the experimental and Boltzmann‐weighted calculated VCD and ECD spectra and assigned as (+)‐(2S;3S)‐di(azulen‐1‐yl)butane‐2,3‐diol. Chirality 27:826–834, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
In order to define an enantioselective nuclear magnetic resonance (NMR) method for the antiasthmatic drug montelukast, a series of nine easily available products were evaluated as NMR chiral solvating agents (CSAs): D‐dibenzoyltartaric acid, D‐ditoluoyltartaric acid, (+)‐camphorsulfonic acid, (S)‐BINOL, (S)‐3,3’‐diphenyl‐2,2’‐binaphthyl‐1,1’‐diol, (R)‐3,3'′‐di‐9‐anthracenyl‐1,1'′‐bi‐2‐naphthol, (R)‐3,3'′‐di‐9‐phenanthrenyl‐1,1'′‐bi‐2‐naphthol, Pirkle's alcohol, and (?)‐cinchonidine. It was proved that most of the studied agents constitute diastereomeric complexes with both drug enantiomers in CD2Cl2 or CDCl3 solutions, thus permitting the direct 1H NMR detection of the unwanted S‐enantiomer, even at levels of 0.75%. (?)‐Cinchonidine was found to be the more convenient CSA in terms of NMR enantiodiscrimination power and ease of experimental requirements. The final method was validated and applied to the fast monitoring of the optical purity of montelukast “in‐process” samples, circumventing the need for tedious and slower analytical procedures like enantioselective chromatography or capillary electrophoresis. In addition, a method for the enantiopurity control of the commercial drug (montelukast sodium salt) was also established using (S)‐BINOL as NMR CSA. Chirality 25: 780–786, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

14.
A practical synthetic method for 2,2'‐disubstituted fluorinated binaphthyl derivatives was achieved using magnesium bis(2,2,6,6‐tetramethylpiperamide) [Mg(TMP)2], prepared from LiTMP (2 equiv) and MgBr2 (1 equiv), which allows for access to a variety of fluorinated binaphthyl compounds. The utility of the fluorinated binaphthyl backbone was evaluated in F10BINOL derived chiral mono‐phosphoric acid (R)‐ 19 as the chiral Brønsted acid catalyst. The catalyst (R)‐ 19 performs exceptionally well in the catalytic enantioselective imino‐ene reaction, demonstrating the potential of a fluorinated binaphthyl framework. Chirality 27:464–475, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

15.
A novel functionalized steroidal BINOL scaffold is produced via a remarkable sequence of dehydrogenation, chlorination and atroposelective phenolic coupling steps. Its formation is straightforwardly obtained in a one-step synthesis under solid state conditions and in the absence of chiral additives.  相似文献   

16.
Hui Wang 《Chirality》2010,22(9):827-837
The enantiomeric atropoisomers of 1,1′‐binaphthyl‐2,2′‐diol (BINOL) have become one of the most widely used chiral ligands and auxiliaries for asymmetric synthesis. This review provides an overview of enantioselective synthesis of optical active BINOLs by straightforward asymmetric oxidative coupling of identical 2‐naphthol and its substituted derivatives. Chirality 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

17.
Zirconium-mediated asymmetric baeyer-villiger oxidation   总被引:2,自引:0,他引:2  
Bolm C  Beckmann O 《Chirality》2000,12(5-6):523-525
Combinations of axially chiral C2-symmetric diols were used as ligands in zirconium-mediated Baeyer-Villiger reactions. The in situ preformed Zr-diol species proved effective in the asymmetric oxidation of bicyclic and monosubstituted cyclobutanones when a hydroperoxide was employed as oxidant. Asymmetric induction could be preserved upon replacement of one out of two enantiopure BINOL ligands by conformationally flexible 2,2'-biphenol.  相似文献   

18.
Yu H  Yin C  Jia C  Jin Y  Ke Y  Liang X 《Chirality》2012,24(5):391-399
Two "click" binaphthyl chiral stationary phases were synthesized and evaluated by liquid chromatography. Their structures incorporate S-(-)-1,1'-binaphthyl moiety as the chiral selector and 1,2,3-triazole ring as the spacer. These chiral stationary phases (CSPs) allowed the efficient resolution for a wide range of racemic BINOL derivatives, particularly for nonpolar diether derivatives and 3-phenyl indolin-2-one analogs. The chromatographic data showed that the π-π interaction was crucial for enantiorecognition of these CSPs. Loss of enantioselectivity observed on CSP3, which are lacking the triazole ring linkage, indicated that the triazole ring linkage took part in the enantioseparation process, although it was remote from the chiral selector of the CSP. The substitution of the phenyl group at 6 and 6' positions can significantly improve the separation ability of the CSP. The chiral recognition mechanism was also investigated by tracking the elution orders and studying the thermodynamic parameters.  相似文献   

19.
Zhou YF  Han ZJ  Qiu L  Liang JY  Ren FB  Wang R 《Chirality》2009,21(4):473-479
Various new chiral hydroxysulfonamide ligands (3a-3n, 4a-4d) were prepared. Compounds 3a, 3g, 3i, 3k-3n, 4a-4d could accelerate the reaction and reduce reaction time, and 3a, 3g, 3i, 3k-3n catalyzed the reaction without titanium. The results obtained were promising in terms of yields and enantiomeric excesses (3k up to 85% ee, 4a up to 83% ee).  相似文献   

20.
Qinghan Li  Han‐Mou Gau 《Chirality》2011,23(10):929-939
Three alkyltitanium reagents of RTi(O‐i‐Pr)3 (R = Cy ( 1a ), i‐Bu ( 1b ), and n‐Bu ( 1c )) were prepared in good yields. The high‐resolution mass spectroscopy showed that 1b and 1 c in the gas phase are monomeric species. However, the solid state of 1a revealed a dimeric structure. Asymmetric additions of 1a , 1b , 1c to aldehydes catalyzed by a titanium catalyst of (R)‐H8‐BINOL were studied at room temperature. The reactions produced desired secondary alcohols in good yields with good to excellent enantioselectivities of up to 94% ee. Reactivity and enantioselectivity differences, in terms of steric bulkiness of the R nucleophiles, are herein described. The addition reactions of secondary c‐hexyl to aldehydes were slower than the reactions of primary i‐butyl or n‐butyl nucleophiles. For the primary alkyls, lower enantioselectivities were obtained for products from addition reactions of the linear n‐butyl as compared with the enantioselectivities of products from the addition reactions of the branched i‐butyl group. The same stereochemistry of RTi(O‐i‐Pr)3 addition reactions as the addition reactions of organozinc, organoaluminum, Grignard, or organolithium reagents directly supports the argument of that titanium‐catalyzed addition reactions of aldehydes involve an addition of an organotitanium nucleophile. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  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