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 48 毫秒
1.
It was shown that racemic (±)‐ 2 [1′‐benzyl‐3‐(3‐fluoropropyl)‐3H‐spiro[[2]benzofuran‐1,4′‐piperidine], WMS‐1813 ] represents a promising positron emission tomography (PET) tracer for the investigation of centrally located σ1 receptors. To study the pharmacological activity of the enantiomers of 2 , a preparative HPLC separation of (R)‐2 and (S)‐2 was performed. The absolute configuration of the enantiomers was determined by CD‐spectroscopy together with theoretical calculations of the CD‐spectrum of a model compound. In receptor binding studies with the radioligand [3H]‐(+)‐pentazocine, (S)‐2 was thrice more potent than its (R)‐configured enantiomer (R)‐2 . The metabolic degradation of the more potent (S)‐enantiomer was considerably slower than the metabolism of (R)‐2 . The structures of the main metabolites of both enantiomers were elucidated by determination of the exact mass using an Orbitrap‐LC‐MS system. These experiments showed a stereoselective biotransformation of the enantiomers of 2 . Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

2.
The enantiomeric separation ability of the newly prepared chiral stationary phases containing acridino‐18‐crown‐6 ether selectors was studied by high‐performance liquid chromatography (HPLC). The chiral stationary phases separated the enantiomers of selected protonated primary aralkylamines efficiently. The best results were found for the separation of the mixtures of enantiomers of NO2‐PEA. Chirality 26:651–654, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

3.
Two racemic tetrathiafulvalene‐[2.2]paracyclophane electron donors EDT‐TTF‐[2.2]paracyclophane 1 and (COOMe)2‐TTF‐[2.2]paracyclophane 2 have been synthesized via the phosphite mediated cross coupling strategy. Chiral HPLC allowed the optical resolution of the (RP) and (SP) enantiomers for both compounds. Solid‐state structures of (RP)‐ 1 and (rac)‐ 2 have been determined by single crystal X‐ray analysis. Intermolecular π‐π and S???S interactions are disclosed in the packing. Single crystal X‐ray analysis of (RP)‐ 1 combined with experimental and theoretical circular dichroism spectra allowed the assignment of the absolute configuration of the enantiomers of 1 and 2 .  相似文献   

4.
《Chirality》2017,29(10):603-609
d ‐ and l ‐Tryptophan (Trp) and d ‐ and l ‐kynurenine (KYN) were derivatized with a chiral reagent, (S )‐4‐(3‐isothiocyanatopyrrolidin‐1‐yl)‐7‐(N,N‐dimethylaminosulfonyl)‐2,1,3‐benzoxadiazole (DBD‐PyNCS), and were separated enantiomerically by high‐performance liquid chromatography (HPLC) equipped with a triazole‐bonded column (Cosmosil HILIC) using tandem mass spectrometric (MS/MS) detection. Effects of column temperature, salt (HCO2NH4) concentration, and pH of the mobile phase in the enantiomeric separation, followed by MS detection of (S )‐DBD‐PyNCS‐d ,l ‐Trp and ‐d ,l ‐KYN, were investigated. The mobile phase consisting of CH3CN/10 mM ammonium formate in H2O (pH 5.0) (90/10) with a column temperature of 50–60 °C gave satisfactory resolution (R s) and mass‐spectrometric detection. The enantiomeric separation of d ,l ‐Trp and d ,l ‐KYN produced R s values of 2.22 and 2.13, and separation factors (α) of 1.08 and 1.08, for the Trp and KYN enantiomers, respectively. The proposed LC–MS/MS method provided excellent detection sensitivity of both enantiomers of Trp and KYN (5.1–19 nM).  相似文献   

5.
Chiral high‐performance liquid chromatography (HPLC) separation of trans‐bis[2‐(2‐pyridyl)aminophenolato] dichlorocyclotriphosphazene 1 was achieved and the absolute configuration of (+)-1 was assigned to be S,S by single‐crystal X‐ray structural analysis. The optically pure 1,2‐diphenyl‐1,2‐ethanediolate derivatives (+)‐ 2a and (?)‐ 2b were synthesized by the reactions of (+)-1 and (-)-1 with (R,R)‐hydrobenzoin, respectively, in refluxing toluene in the presence of an excess amount of triethylamine and a catalytic amount of 4‐(dimethylamino)pyridine. The racemization of the enantiomers of 1 and the epimerization of diastereomers of 2 were not observed in refluxing toluene neither under acidic nor basic conditions. The stereochemistry of (+)-1 was confirmed by the crystal structure of (+)‐ 2a and bis[(4‐methyl‐2‐pyridyl)oxy]cyclotriphosphazene (+)-3 derived from (+)-1 . Chirality 28:556–561, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

6.
《Chirality》2017,29(1):48-56
For the enantiopure synthesis of novel chiral GABA uptake inhibitors, nipecotic acid ( 1 ) is an important key precursor. To characterize accurately the pharmacological activity of these interesting target compounds, the determination of the correct enantiomeric purity of nipecotic acid as the starting material is indispensable. In this report, a sensitive high‐performance liquid chromatography (HPLC) based method for the separation and quantitation of both enantiomers of nipecotic acid as 1‐(7‐nitrobenzo[c ][1,2,5]oxadiazol‐4‐yl) derivatives ( 5 ) on a Chiralpak ID‐3 column (Daicel, Illkirch, France) was established. UV/Vis‐detection at 490 nm was chosen to ensure a selective determination of even highly enantioenriched samples. Reliability was demonstrated by validation of specificity, linearity, lower limit of quantification (LLOQ), accuracy, and precision. By spiking highly enantiopure samples with small amounts of racemic rac ‐ 5 , it was proven that the established HPLC method is able to detect even slight changes in enantiomeric excess (ee) values. Thus, accurate determination of ee values up to 99.87% ee for (R )‐ 5 and 99.86% ee for (S )‐ 5 over a linear concentration range of 11500 μM for (R )‐ 5 and of 11455 μM for (S )‐ 5 could be demonstrated.  相似文献   

7.
A sensitive and specific method for the simultaneous determination of the enantiomeric purity of 2,6‐dimethylphenoxyacetyl derivatives as trans or cis racemic and enantiomeric forms with 2‐ or 4‐aminocyclohexanol moiety ( 1 , 2 , 3 , 6 ) and their amine analogs ( 8 , 9 ) was developed. The compounds studied are known for their anticonvulsant activity and the most interesting pharmacological results were those for (±)‐trans‐2‐(2,6‐dimethylphenoxy)‐N‐(2‐hydroxycyclohexyl)acetamide ( 1 ) as well as (±)‐trans‐2‐[(2,6‐dimethylphenoxy)ethyl]aminocyclohexanol ( 8 ). The analytical method for determining the enantiomeric purity of the compounds studied is based on direct separation of the analytes using a chiral stationary phase (Chiralpak AS column). The mass spectrometric analysis was done on a coupled liquid chromatograph–mass spectrometer system with an electrospray ionization source (LC/ESI‐MS). For the compounds 1 , 8 , and 9 , the method allows an excellent separation of enantiomers, with a resolution higher than 3.2, and a tailing factor of less than 1.67 with a final enantiomer purity better than 97.5%. Chirality 26:144–149, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

8.
Alpha‐cypermethrin (α‐CP), [(RS)‐a‐cyano‐3‐phenoxy benzyl (1RS)‐cis‐3‐(2, 2‐dichlorovinyl)‐2, 2‐dimethylcyclopropanecarboxylate], comprises a diastereoisomer pair of cypermethrin, which are (+)‐(1R‐cis‐αS)–CP (insecticidal) and (?)‐(1S‐cis‐αR)–CP (inactive). In this experiment, the stereoselective degradation of α‐CP was investigated in rat liver microsomes by high‐performance liquid chromatography (HPLC) with a cellulose‐tris‐ (3, 5‐dimethylphenylcarbamate)‐based chiral stationary phase. The results revealed that the degradation of (?)‐(1S‐cis‐αR)‐CP was much faster than (+)‐(1R‐cis‐αS)‐CP both in enantiomer monomers and rac‐α‐CP. As for the enzyme kinetic parameters, there were some variances between rac‐α‐CP and the enantiomer monomers. In rac‐α‐CP, the Vmax and CLint of (+)‐(1R‐cis‐αS)–CP (5105.22 ± 326.26 nM/min/mg protein and 189.64 mL/min/mg protein) were about one‐half of those of (?)‐(1S‐cis‐αR)–CP (9308.57 ± 772.24 nM/min/mg protein and 352.19 mL/min/mg protein), while the Km of the two α‐CP enantiomers were similar. However, in the enantiomer monomers of α‐CP, the Vmax and Km of (+)‐(1R‐cis‐αS) ‐CP were 2‐fold and 5‐fold of (?)‐(1S‐cis‐αR)‐CP, respectively, which showed a significant difference with rac‐α‐CP. The CLint of (+)‐(1R‐cis‐αS)–CP (140.97 mL/min/mg protein) was still about one‐half of (?)‐(1S‐cis‐αR)–CP (325.72 mL/min/mg protein) in enantiomer monomers. The interaction of enantiomers of α‐CP in rat liver microsomes was researched and the results showed that there were different interactions between the IC50 of (?)‐ to (+)‐(1R‐cis‐αS)‐CP and (+)‐ to (?)‐(1S‐cis‐αR)‐CP(IC50(?)/(+) / IC50(+)/(?) = 0.61). Chirality 28:58–64, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

9.
Enantiomers of a 2‐benzofuran‐1(3H)‐one derivative [(–)‐ 1 and (+)‐ 1 ] and four known analogs ( 2 , 3 , 4 , 5 ) were isolated and identified from the culture extract of Eurotium rubrum MA–150, a fungus obtained from the mangrove‐derived rizospheric soil. Their structures were established by detailed interpretation of nuclear magnetic resonance (NMR) data and the structure of (±)‐ 1 was confirmed by single‐crystal X‐ray diffraction analysis. The absolute configuration of the enantiomers (–)‐ 1 and (+)‐ 1 was determined by means of online high‐performance liquid chromatography – electronic circular dichroism (HPLC‐ECD) measurements and time‐dependent Density Functional Theory – electronic circular dichroism (TDDFT‐ECD) calculations. Compounds (±)‐ 1 as well as 2 and 3 exhibited potent DPPH radical scavenging activities with IC50 values of 1.23, 2.26, and 3.99 μg/mL, respectively. Chirality 28:581–584, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

10.
《Chirality》2017,29(1):26-32
The purpose of this study was to compare intestinal permeability between enantiomers of 2‐(2‐hydroxypropanamido) benzoic acid ((R )‐/ (S )HPABA), a marine‐derived antiinflammatory drug, using an in situ single‐pass intestinal perfusion (SPIP) model in rats. Concentrations, isolated regions of small intestine, and p ‐glycoprotein (P‐gp) inhibitor were performed to investigate their influences on the intestinal absorption of (R )‐/ (S )HPABA. In addition, a molecular docking method was performed to illustrate our prediction. The absorption rate coefficients (K a ) and permeability values (P eff ) of (R )‐/ (S )HPABA were calculated. The permeability of (S )‐HPABA was significantly (P <  0.01) higher than that of (R )‐HPABA in jejunum, and ileum permeability of (R )‐/ (S )HPABA appeared best in ileum; the investigated concentrations ranged from 20 to 80 μg/mL, K a and P eff values of (R )‐/ (S )HPABA increased linearly; in the presence of P‐gp inhibitor (verapamil), P eff values of two enantiomers were increased significantly; and the effect of P‐gp on absorption of (R )‐HPABA is stronger than that of (S )‐HPABA in ileum segment. Based on these results, carrier‐mediated transport or passive transport combined with carrier‐mediated transport seems to be the mechanism for intestinal absorption of (R )‐/ (S )HPABA, and (R )‐/ (S )HPABA may be recognized as the P‐gp substrate. In addition, the intestinal permeability of (S )‐HPABA is higher than that of (R )‐HPABA.  相似文献   

11.
Oxcarbazepine is a second‐generation antiepileptic drug indicated as monotherapy or adjunctive therapy in the treatment of partial seizures or generalized tonic–clonic seizures in adults and children. It undergoes rapid presystemic reduction with formation of the active metabolite 10‐hydroxycarbazepine (MHD), which has a chiral center at position 10, with the enantiomers (S)‐(+)‐ and R‐(?)‐MHD showing similar antiepileptic effects. This study presents the development and validation of a method of sequential analysis of oxcarbazepine and MHD enantiomers in plasma using liquid chromatography with tandem mass spectrometry (LC‐MS/MS). Aliquots of 100 μL of plasma were extracted with a mixture of methyl tert‐butyl ether: dichloromethane (2:1). The separation of oxcarbazepine and the MHD enantiomers was obtained on a chiral phase Chiralcel OD‐H column, using a mixture of hexane:ethanol:isopropanol (80:15:5, v/v/v) as mobile phase at a flow rate of 1.3 mL/min with a split ratio of 1:5, and quantification was performed by LC‐MS/MS. The limit of quantification was 12.5 ng oxcarbazepine and 31.25 ng of each MHD enantiomer/mL of plasma. The method was applied in the study of kinetic disposition of oxcarbazepine and the MHD enantiomers in the steady state after oral administration of 300 mg/12 h oxcarbazepine in a healthy volunteer. The maximum plasma concentration of oxcarbazepine was 1.2 µg/mL at 0.75 h. The kinetic disposition of MHD is enantioselective, with a higher proportion of the S‐(+)‐MHD enantiomer compared to R‐(?)‐MHD and an AUC0‐12 S‐(+)/R‐(?) ratio of 5.44. Chirality 25:897–903, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
Compounds based on the pyrroloquinoxaline system can interact with serotonin 5‐HT3, cannabinoid CB1, and μ‐opioid receptors. Herein, a chiral pool synthesis of diastereomerically and enantiomerically pure bromolactam (S,R,R,R)‐ 14A is presented. Introduction of the cyclohexenyl ring at the N‐atom of (S)‐proline derivatives 8 or methyl (S)‐pyroglutamate ( 12 ) led to the N‐cyclohexenyl substituted pyrrolidine derivatives 4 and 13 , respectively. All attempts to cyclize the (S)‐proline derivatives 4 with a basic pyrrolidine N‐atom via [3 + 2] cycloaddition, aziridination, or bromolactamization failed. Fast aromatization occurred during treatment of cyclohexenamines under halolactamization conditions. In contrast, reaction of a 1:1 mixture of diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB with LiOtBu and NBS provided the tricyclic bromolactam (S,R,R,R)‐ 14A with high diastereoselectivity from (S,R)‐ 13bA , but did not transform the diastereomer (S,S)‐ 13bB . The different behavior of the diastereomeric pyroglutamates (S,R)‐ 13bA and (S,S)‐ 13bB is explained by different energetically favored conformations. Chirality 26:793–800, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
The transport and uptake of individual propranolol (PPL) enantiomers were studied in human intestinal Caco‐2 cell monolayers, and a reversed‐phase HPLC‐UV assay was used for quantitative analysis. S‐PPL and R‐PPL across Caco‐2 cell monolayers was determined in the concentrations range of 10–500 μM in both apical (AP) to basolateral (BL) and BL to AP directions. S‐PPL exhibited greater permeability than R‐PPL in the AP to BL direction, whereas in the BL to AP direction S‐enantiomer transported less than R‐enantiomer. Uptake of R‐PPL was significantly higher than that of S‐PPL either from AP side or from BL side. The statistically significant differences in uptake were observed at the concentrations range from 10 to 50 μM. Furthermore, the apparent Michaelis constant (Km) and maximal velocity (Vmax) also showed significant difference between the two enantiomers. Moreover, the AP to BL transport of PPL enantiomer was markedly decreased by lowering the pH of the apical side but it did not affect the stereoselectivity of PPL across Caco‐2 cell monolayers. The transport and uptake of PPL in the BL to AP direction was not influenced by several protein inhibitors. The results suggest that PPL enantiomers showed stereoselective transport and uptake across the Caco‐2 cell monolayers. A special transport mechanism capable of directing the PPL enantiomers might be present in the Caco‐2 monolayers. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

14.
Applications of the on‐line HPLC‐exciton CD analysis using (S)‐2‐tert‐butyl‐2‐methyl‐1,3‐benzodioxole‐4‐carboxylic acid [(S)‐TBMBC‐OH] that can simultaneously determine the enantiomeric compositions and the absolute configuration of cyclohexane‐1,2‐diols and diamines as well as acyclic vicinal diols and amino alcohols were studied. Di‐O‐ or di‐N,O‐(S)‐TBMBC derivatives of acyclic terminal vicinal diols, 2‐hydroxy‐1‐amines, and nonterminal vicinal diols gave symmetrical exciton CD spectra between enantiomers, indicating their absolute configurations. However, Di‐N,O‐(S)‐TBMBC derivatives of 2‐amino‐1‐ols did not always give symmetrical exciton CD spectra between enantiomers, but their 2‐phthalimido‐1‐O‐(S)‐TBMBC derivatives gave symmetrical exciton CD spectra, indicating their absolute configurations. All these (S)‐TBMBC derivatives were separated by normal‐phase HPLC and unequivocally determined by the on‐line HPLC‐exciton CD analysis without recourse to reference samples. Chirality 11:149–159, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

15.
A novel method for chiral separation of flurbiprofen enantiomers was developed using aqueous two‐phase extraction (ATPE) coupled with biphasic recognition chiral extraction (BRCE). An aqueous two‐phase system (ATPS) was used as an extracting solvent which was composed of ethanol (35.0% w/w) and ammonium sulfate (18.0% w/w). The chiral selectors in ATPS for BRCE consideration were L‐dioctyl tartrate and L‐tryptophan, which were screened from amino acids, β‐cyclodextrin derivatives, and L‐tartrate esters. Factors such as the amounts of L‐dioctyl tartrate and L‐tryptophan, pH, flurbiprofen concentration, and the operation temperature were investigated in terms of chiral separation of flurbiprofen enantiomers. The optimum conditions were as follows: L‐dioctyl tartrate, 80 mg; L‐tryptophan, 40 mg; pH, 4.0; flurbiprofen concentration, 0.10 mmol/L; and temperature, 25 °C. The maximum separation factor α for flurbiprofen enantiomers could reach 2.34. The mechanism of chiral separation of flurbiprofen enantiomers is discussed and studied. The results showed that synergistic extraction has been established by L‐dioctyl tartrate and L‐tryptophan, which enantioselectively recognized R‐ and S‐enantiomers in top and bottom phases, respectively. Compared to conventional liquid–liquid extraction, ATPE coupled with BRCE possessed higher separation efficiency and enantioselectivity without the use of any other organic solvents. The proposed method is a potential and powerful alternative to conventional extraction for separation of various enantiomers. Chirality 27:650–657, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

16.
Chiral discrimination observed in high‐performance liquid chromatography (HPLC) with the novel chiral stationary phase (CSP‐18C6I) derived from (+)‐(R)‐18‐crown‐6 tetracarboxylic acid [(+)‐18C6H4] was investigated by X‐ray crystallographic analysis of the complex composed of the R‐enantiomer of 1‐(1‐naphthyl)ethylamine (1‐NEA) and (+)‐18C6H4. Mixtures of 1‐NEA (the R‐ or S‐enantiomer) and (+)‐18C6H4 were dissolved in methanol‐water (1:1) solution and allowed to stand for crystallization. The R‐enantiomer crystallized with (+)‐18C6H4 as a co‐crystal, although the S‐enantiomer did not. This result was in good agreement with the enantiomer elution order of 1‐NEA in CSP‐18C6I. The apparent binding constants (Ka) of the enantiomers to the (+)‐18C6H4 obtained from 1H‐NMR experiments also supported the above‐mentioned result. The X‐ray crystal structure of the 1:1 complex of the R‐enantiomer and (+)‐18C6H4 indicated the four sets of hydrogen bond association between the naphthylethylammonium cation and oxygen of polyether ring or carbonyl group of (+)‐18C6H4. Chirality 11:173–178, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

17.
Resolution of rac‐3,3,3‐trifluorolactic acid by diastereomeric salt formation was reinvestigated. The use of (S)‐1‐phenylethylamine gives coprecipitation of two diastereomeric phases, 1 (S)‐[NH3CH(CH3)Ph](S)‐[CF3CH(OH)COO] and 2 (S)‐[NH3CH(CH3)Ph](R)‐[CF3CH(OH)COO]·H2O. Pure phase 1 may be obtained using molecular sieves as desiccants. Resolution by (S,S)‐2‐amino‐1‐phenylpropan‐1,3‐diol gives monoclinic (S,S)‐[NH3CH(CH2OH)CHOHPh] (R)‐[CF3CH(OH)‐COO] 3 with minor (S)‐3,3,3‐trifluorolactate contamination, which is precluded in the recrystallized orthorhombic form 4 . A new resolution using inexpensive phenylglycinol gives pure phase 5 (S)‐[NH3CH(CH2OH)Ph] (S)‐[CF3CH(OH)COO] in 76% yield, 94% ee in a single step, in preference to its (S)‐(R) diastereomer 6 . Overall efficient resolution for both enantiomers of the trifluorolactic acid (each ca. 70% yield, 99% ee) may be achieved by various two‐step “tandem” crystallizations, involving direct addition of either water or a second base to the filtrate from the initial reaction.  相似文献   

18.
L ‐Cysteine was condensed with glyoxylic acid monohydrate in acetic acid at 30°C to give (4R)‐2,4‐thiazolidinedicarboxylic acid [(4R)‐TDA] as a mixture of two diastereoisomers, (2R,4R)‐ and (2S,4R)‐TDA. An attempt was made to separate (2S,4R)‐TDA from the diastereoisomeric salts of (4R)‐TDA with 1‐propylamine, 2‐methyl‐2‐propylamine, benzylamine, and (R)‐ and (S)‐1‐phenylethylamines [(R)‐ and (S)‐PEA]. The salts of (2S,4R)‐TDA were preferentially crystallized as less soluble diastereoisomeric salts. When the salt with (R)‐PEA was employed, the separation was successfully achieved to afford optically pure (2S,4R)‐TDA in a yield of 41%, based on the starting amount of the diastereoisomeric mixture of (4R)‐TDA. Chirality 11:326–329, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

19.
Economic and enantioselective synchronous fluorescence spectroscopy and high‐performance thin‐layer chromatography methods have been developed and validated as per ICH guidelines for the separation of zopiclone enantiomers using L‐(+)‐tartaric acid as a chiral selector, followed by determination of the chiral‐switching eszopiclone. Synchronous fluorescence spectroscopy was successfully applied for chiral recognition of R & S enantiomers of zopiclone at  = 110 nm based on creating of diastereomeric complexes with 0.06M tartaric acid in an aqueous medium containing 0.2M disodium hydrogen orthophosphate. Synchronous fluorescence intensities of eszopiclone were recorded at 296 nm in concentration range 0.2‐ to 4‐μg/mL eszopiclone. High‐performance thin‐layer chromatography method depends on resolution of zopiclone enantiomers on achiral HPTLC silica‐gel plates using acetonitrile:methanol:water (8:2:0.25, v/v/v) containing L‐(+)‐tartaric acid as a chiral mobile‐phase additive followed by densitometric measurements at 304 nm in concentration range of 1 to 10 μg/band of eszopiclone. The effect of chiral‐selector concentration, pH, and temperature on the resolution have been studied and optimized for the proposed methods. The cited procedures were successfully applied to determine eszopiclone in commercial tablets of pure and racemic forms. Enantiomeric excess was evaluated using optical purity test and integrated peak area to describe the enantiomeric ratio. Thermodynamics of chromatographic separation, enthalpy, and entropy were evaluated using the Van't Hoff equation. The proposed methods were found to be selective for identification and determination of the eutomer in drug substances and products.  相似文献   

20.
《Chirality》2017,29(6):257-272
A few new l ‐threitol‐based lariat ethers incorporating a monoaza‐15‐crown‐5 unit were synthesized starting from diethyl l ‐tartrate. These macrocycles were used as phase transfer catalysts in asymmetric Michael addition reactions under mild conditions to afford the adducts in a few cases in good to excellent enantioselectivities. The addition of 2‐nitropropane to trans ‐chalcone, and the reaction of diethyl acetamidomalonate with β‐nitrostyrene resulted in the chiral Michael adducts in good enantioselectivities (90% and 95%, respectively). The substituents of chalcone had a significant impact on the yield and enantioselectivity in the reaction of diethyl acetoxymalonate. The highest enantiomeric excess (ee ) values (99% ee ) were measured in the case of 4‐chloro‐ and 4‐methoxychalcone. The phase transfer catalyzed cyclopropanation reaction of chalcone and benzylidene‐malononitriles using diethyl bromomalonate as the nucleophile (MIRC reaction) was also developed. The corresponding chiral cyclopropane diesters were obtained in moderate to good (up to 99%) enantioselectivities in the presence of the threitol‐based crown ethers.  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