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 synthesis of enantiomerically pure (+)‐ and (−)‐γ‐ionone 3 is reported. The first step in the synthesis is the diastereoisomeric enrichment of 4‐nitrobenzoate derivatives of racemic γ‐ionol 12 . The enantioselective lipase‐mediated kinetic acetylation of γ‐ionol 13b afforded the acetate 14 and the alcohol 15 , which are suitable precursors of the desired products (−)‐ and (+)‐ 3 , respectively. The olfactory evaluation of the γ‐ionone isomers shows a great difference between the two enantiomers both in fragrance response and in detection threshold. The selective reduction of (−)‐ 3 and (+)‐ 3 to the γ‐dihydroionones (−)‐(R)‐ 16 and (+)‐(S)‐ 17 , respectively, allowed us to assign unambiguously the absolute configuration of the γ‐ionones.  相似文献   

2.
The cross‐aldolization of (−)‐(1S,4R,5R,6R)‐6‐endo‐chloro‐5‐exo‐(phenylseleno)‐7‐oxabicyclo[2.2.1]heptan‐2‐one ((−)‐ 25 ) and of (+)‐(3aR,4aR,7aR,7bS)‐ ((+)‐ 26 ) and (−)‐(3aS,4aS,7aS,7bR)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]furo[2,3‐d]isoxazole‐3‐carbaldehyde ((−)‐ 26 ) was studied for the lithium enolate of (−)‐ 25 and for its trimethylsilyl ether (−)‐ 31 under Mukaiyama's conditions (Scheme 2). Protocols were found for highly diastereoselective condensation giving the four possible aldols (+)‐ 27 (`anti'), (+)‐ 28 (`syn'), 29 (`anti'), and (−)‐ 30 (`syn') resulting from the exclusive exo‐face reaction of the bicyclic lithium enolate of (−)‐ 25 and bicyclic silyl ether (−)‐ 31 . Steric factors can explain the selectivities observed. Aldols (+)‐ 27 , (+)‐ 28 , 29 , and (−)‐ 30 were converted stereoselectively to (+)‐1,4‐anhydro‐3‐{(S)‐[(tert‐butyl)dimethylsilyloxy][(3aR,4aR,7aR,7bS)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]‐furo[2,3‐d]isoxazol‐3‐yl]methyl}‐3‐deoxy‐2,6‐di‐O‐(methoxymethyl)‐α‐D ‐galactopyranose ((+)‐ 62 ), its epimer at the exocyclic position (+)‐ 70 , (−)‐1,4‐anhydro‐3‐{(S)‐[(tert‐butyl)dimethylsilyloxy][(3aS,4aS,7aS,7bR)‐3a,4a,7a,7b‐tetrahydro‐6,6‐dimethyl[1,3]dioxolo[4,5]furo[2,3‐d]isoxazol‐3‐yl]methyl}‐3‐deoxy‐2,6‐di‐O‐(methoxymethyl)‐α‐D ‐galactopyranose ((−)‐ 77 ), and its epimer at the exocyclic position (+)‐ 84 , respectively (Schemes 3 and 5). Compounds (+)‐ 62 , (−)‐ 77 , and (+)‐ 84 were transformed to (1R,2R,3S,7R,8S,9S,9aS)‐1,3,4,6,7,8,9,9a‐octahydro‐8‐[(1R,2R)‐1,2,3‐trihydroxypropyl]‐2H‐quinolizine‐1,2,3,7,9‐pentol ( 21 ), its (1S,2S,3R,7R,8S,9S,9aR) stereoisomer (−)‐ 22 , and to its (1S,2S,3R,7R,8S,9R,9aR) stereoisomer (+)‐ 23 , respectively (Schemes 6 and 7). The polyhydroxylated quinolizidines (−)‐ 22 and (+)‐ 23 adopt `trans‐azadecalin' structures with chair/chair conformations in which H−C(9a) occupies an axial position anti‐periplanar to the amine lone electron pair. Quinolizidines 21 , (−)‐ 22 , and (+)‐ 23 were tested for their inhibitory activities toward 25 commercially available glycohydrolases. Compound 21 is a weak inhibitor of β‐galactosidase from jack bean, of amyloglucosidase from Aspergillus niger, and of β‐glucosidase from Caldocellum saccharolyticum. Stereoisomers (−)‐ 22 and (+)‐ 23 are weak but more selective inhibitors of β‐galactosidase from jack bean.  相似文献   

3.
(S)‐(?)‐Ofloxacin and (R)‐(+)‐ofloxacin concentrations in the plasma of Pagrosomus major after drug treatment were detected by chiral high‐performance liquid chromatography, and various pharmacokinetic parameters were calculated from these data. The elimination half‐life of (S)‐(?)‐ofloxacin was significantly shorter than that of the (R)‐(+) enantiomer. (S)‐(?)‐Ofloxacin also had a significantly lower maximum plasma concentration, area under the concentration–time curve from zero to infinity, and mean residence time than (R)‐(+)‐ofloxacin. However, the apparent volume of distribution and total body clearance of (S)‐(?)‐ofloxacin were greater than those of (R)‐(+)‐ofloxacin. The ratio of the (S)‐(?)‐ to (R)‐(+)‐ofloxacin plasma concentration was always <1.0. Together, these data suggest that (S)‐(?)‐ofloxacin was preferentially excreted and (R)‐(+)‐ofloxacin was preferentially absorbed. Although the difference in pharmacokinetic parameters was small, the metabolic behavior of the ofloxacin enantiomers in P. major was enantioselective. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

4.
Enzymatic resolution of racemic 1,4,5,6‐tetrachloro‐2‐(hydroxymethyl)‐7,7‐dimethoxybicyclo[2.2.1]hept‐5‐ene (rac‐ 1 ) using various lipases in vinyl acetate as acetyl source was studied. The obtained enantiomerically enriched (+)‐(1,4,5,6‐tetrachloro‐7,7‐dimethoxybicyclo[2.2.1]hept‐5‐en‐2‐yl)methyl acetate ((+)‐ 2 ; 94% ee), upon treatment with Na in liquid NH3, followed by Amberlyst‐15 resin in acetone, provided (−)‐5‐(hydroxymethyl)bicyclo[2.2.1]hept‐2‐en‐7‐one ((−)‐ 7 ), which is a valuable precursor for the synthesis of carbasugar derivatives. Subsequent Baeyer–Villiger oxidation afforded a nonseparable mixture of bicyclic lactones, which was subjected to LiAlH4 reduction and then acetylation. The resultant compounds (−)‐ 11 and (+)‐ 12 were submitted to a cis‐hydroxylation reaction, followed by acetylation, to afford the novel carbasugar derivatives (1S,2R,3S,4S,5S)‐4,5‐bis(acetoxymethyl)cyclohexane‐1,2,3‐triyl triacetate ((−)‐( 13 )) and (1R,3R,4R,6R)‐4,6‐bis(acetoxymethyl)cyclohexane‐1,2,3‐triyl triacetate ((−)‐( 14 )), respectively, with pseudo‐C2‐symmetric configuration. The absolute configuration of enantiomerically enriched unreacted alcohol (−)‐ 1 (68% ee) was determined by X‐ray single‐crystal analysis by anchoring optically pure (R)‐1‐phenylethanamine. Based on the configurational correlation between (−)‐ 1 and (+)‐ 2 , the absolute configuration of (+)‐ 2 was determined as (1R,2R,4S).  相似文献   

5.
The (−)‐ and (+)‐β‐irones ((−)‐ and (+)‐ 2 , resp.), contaminated with ca. 7 – 9% of the (+)‐ and (−)‐transα‐isomer, respectively, were obtained from racemic α‐irone via the 2,6‐trans‐epoxide (±)‐ 4 (Scheme 2). Relevant steps in the sequence were the LiAlH4 reduction of the latter, to provide the diastereoisomeric‐4,5‐dihydro‐5‐hydroxy‐transα‐irols (±)‐ 6 and (±)‐ 7 , resolved into the enantiomers by lipase‐PS‐mediated acetylation with vinyl acetate. The enantiomerically pure allylic acetate esters (+)‐ and (−)‐ 8 and (+)‐ and (−)‐ 9 , upon treatment with POCl3/pyridine, were converted to the β‐irol acetate derivatives (+)‐ and (−)‐ 10 , and (+)‐ and (−)‐ 11 , respectively, eventually providing the desired ketones (+)‐ and (−)‐ 2 by base hydrolysis and MnO2 oxidation. The 2,6‐cis‐epoxide (±)‐ 5 provided the 4,5‐dihydro‐4‐hydroxy‐cisα‐irols (±)‐ 13 and (±)‐ 14 in a 3 : 1 mixture with the isomeric 5‐hydroxy derivatives (±)‐ 15 and (±)‐ 16 on hydride treatment (Scheme 1). The POCl3/pyridine treatment of the enantiomerically pure allylic acetate esters, obtained by enzymic resolution of (±)‐ 13 and (±)‐ 14 , provided enantiomerically pure cisα‐irol acetate esters, from which ketones (+)‐ and (−)‐ 22 were prepared (Scheme 4). The same materials were obtained from the (9S) alcohols (+)‐ 13 and (−)‐ 14 , treated first with MnO2, then with POCl3/pyridine (Scheme 4). Conversely, the dehydration with POCl3/pyridine of the enantiomerically pure 2,6‐cis‐5‐hydroxy derivatives obtained from (±)‐ 15 and (±)‐ 16 gave rise to a mixture in which the γ‐irol acetates 25a and 25b and 26a and 26b prevailed over the α‐ and β‐isomers (Scheme 5). The (+)‐ and (−)‐cisγ‐irones ((+)‐ and (−)‐ 3 , resp.) were obtained from the latter mixture by a sequence involving as the key step the photochemical isomerization of the α‐double bond to the γ‐double bond. External panel olfactory evaluation assigned to (+)‐β‐irone ((+)‐ 2 ) and to (−)‐cisγ‐irone ((−)‐ 3 ) the strongest character and the possibility to be used as dry‐down note.  相似文献   

6.
The title compound (±)‐ 1 , a recently discovered, valuable, floral‐type odorant, has been synthesized by a straightforward procedure (Scheme 1). To determine the properties of the enantiomers of 1 , their separation by preparative HPLC and the determination of their absolute configuration by X‐ray crystallography were carried out (Figure). Furthermore, the analogues 2 – 6 were synthesized, either from differently methylated 2‐methylindan‐1‐ones (Schemes 2 and 3) or, in the case of the 2,4,6‐trimethylated homologue 6 , by a completely different synthetic approach (Scheme 4). An evaluation of (+)‐(S)‐ 1 , (−)‐(R)‐ 1 , and (±)‐ 1 showed only minor differences in terms of odor (Table).  相似文献   

7.
Discrepancies between chiroptical data from the literature and our determination of the structure of the title compounds (+)‐ 5 and (+)‐ 9a were resolved by an unambiguous assignment of their absolute configuration. Accordingly, the dextrorotatory cis‐3‐hydroxy esters have (3R,4R)‐ and the laevorotatory enantiomers (3S,4S)‐configuration. The final evidences were demonstrated on both enantiomers (+)‐ and (?)‐ 5 by biological reduction of 4 by bakers' yeast and stereoselective [RuII(binap)]‐catalyzed hydrogenations of 4 (Scheme 2), by the application of the NMR Mosher method on (+)‐ and (?)‐ 5 (Scheme 3), as well as by the transformation of (+)‐ 5 into a common derivative and chiroptical correlation (Scheme 4).  相似文献   

8.
A liquid chromatography–photodiode array (LC‐PDA) method using a chiral analytical column was developed to determine the plasma levels of enantiomers of methadone and its chiral metabolite, 2‐ethylidene‐1,5‐dimethyl‐3,3‐diphenylpyrrolidine (EDDP), without the standard compounds of R‐form or S‐form enantiomers. This method was established by the characteristics of recombinant cytochrome P‐450 (CYP) isozymes, where CYP2C19 prefers to metabolize R‐methadone and CYP2B6 prefers to metabolize S‐methadone. We incubated the racemic methadone standard with either enzyme for 24 h. We identified the retention times of R‐ and S‐methadone to be around 10.72 and 14.46 min, respectively. Furthermore, we determined the retention times of R‐ and S‐EDDP to be approximately 6.76 and 7.72 min, respectively. No interferences were shown through the retention times of morphine, buprenorphine and diazepam. With the high recovery rate of a solid‐phase extraction procedure, this method was applied in analyzing plasma concentrations of seven methadone maintenance patients where R‐ and S‐methadone and R‐ and S‐EDDP were 233.4 ± 154.9 and 185.9 ± 136.3 ng/mL and 84.4 ± 99.4 and 37.6 ± 22.9 ng/mL, respectively. These data suggest that the present method can be applied for routine assay for plasma methadone and EDDP concentrations for patients under treatment. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
Ketamine is an N‐methyl‐d ‐aspartate receptor antagonist that is usually used clinically as a racemic mixture. Its two enantiomers exhibit different pharmacological activities. To determine whether the enantiomers have different pharmacokinetic profiles, a chiral liquid chromatography–tandem mass spectrometry method was developed and validated for the determination of ketamine enantiomers in dog plasma. The enantiomers of ketamine were extracted from 50 μL of plasma by methyl tert‐butyl ether. Adequate chromatographic retention and baseline resolution of the enantiomers were achieved within a runtime of 5 min on a chiral column coated with polysaccharide derivatives, using a gradient mobile phase of acetonitrile and 10 mm ammonium bicarbonate aqueous solution. Ketamine enantiomers were detected by mass spectrometry with multiple reaction monitoring mode using the transitions of m/z 238.3 → 125.9 for the analytes and m/z 237.1 → 194.1 for carbamazepine (internal standard). The method was linear over the concentration range from 0.5 to 500 ng/mL for each enantiomer. The lower limit of quantification (LLOQ) for each enantiomer was 0.5 ng/mL. The intra‐ and inter‐day precision was <7.3% and 8.5% for R‐ and S‐ketamine, respectively. The accuracy was 92.9–110.4% for R‐ketamine and 99.8–102.4% for S‐ketamine. The method was successfully applied to characterize the stereoselective pharmacokinetic profiles of ketamine in beagle dogs.  相似文献   

10.
A process‐scale stereoselective synthesis of nature‐identical (−)‐(S,S)‐7‐hydroxycalamenal (=(−)‐(5S,8S)‐5,6,7,8‐tetrahydro‐3‐hydroxy‐5‐methyl‐8‐(1‐methylethyl)naphthalene‐2‐carbaldehyde; (−)‐ 1a ) in 96% enantiomeric excess (ee) with the aid of chiral Ru complexes has been developed. The key step was the enantioselective hydrogenation of easily accessible 2‐(4‐methoxyphenyl)‐3‐methylbut‐2‐enoic acid ( 10 ) to (+)‐ 11 in a 86% ee (Scheme 5 and Table 1). A substantial increase in optical purity (96% ee) was achieved by induced crystallization of the intermediate (+)‐3,4‐dihydro‐4‐(1‐methylethyl)‐7‐methoxy‐2H‐naphthalen‐1‐one ((+)‐ 3 ). Computational conformation analysis carried out on the analog (−)‐ 9 rationalized the high diastereoselectivity achieved in the catalytic hydrogenation of the CC bond.  相似文献   

11.
In order to accurately investigate the preclinical pharmacokinetics of (R)‐(+)‐rabeprazole sodium injection, a reliable high‐performance liquid chromatography (HPLC) method was developed using a Chiral‐AGP column to prove that there is no chiral bioconversion of (R)‐(+)‐rabeprazole to (S)‐(?)‐rabeprazole in beagle dogs after single intravenous administration of (R)‐(+)‐rabeprazole sodium injection. An HPLC–tandem mass spectrometry (HPLC‐MS/MS) method for analysis of (R)‐(+)‐rabeprazole was developed and validated, and used to acquire the pharmacokinetic parameters in beagle dogs. (R)‐(+)‐Rabeprazole and internal standard omeprazole were extracted from plasma samples by protein precipitation and separated on a C18 column using methanol–5 mm ammonium acetate as mobile phase. Detection was performed using a turbo‐spray ionization source and mass spectrometric positive multi‐reaction monitoring mode. The linear relationship was achieved in the range from 2.5 to 5000 ng/mL. The method also afforded satisfactory results in terms of sensitivity, specificity, precision, accuracy and recovery as well as the stability of the analyte under various conditions, and was successfully applied to a preclinical pharmacokinetic study in beagle dogs after single intravenous administrations of (R)‐(+)‐rabeprazole sodium injection at 0.33, 2 and 6 mg/kg. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

12.
A highly sensitive, specific and enantioselective assay has been validated for the quantitation of OTX015 enantiomers [(+)‐OTX015 and (−)‐OTX015] in mice plasma on LC–MS/MS‐electrospray ionization as per regulatory guidelines. Protein precipitation was used to extract (±)‐OTX015 enantiomers and internal standard (IS) from mice plasma. The active [(−)‐OTX015] and inactive [(+)‐OTX015] enantiomers were resolved on a Chiralpak‐IA column using an isocratic mobile phase (0.2% ammonia/acetonitrile 20 : 80, v /v) at a flow rate of 1.2 mL/min. The total run time was 6.0 min. (+)‐OTX015, (−)‐OTX015 and IS eluted at 3.34, 4.08 and 4.77 min, respectively. The MS/MS ion transitions monitored were m/z 492 → 383 for OTX015 and m/z 457 → 401 for IS. The standard curves for OTX015 enantiomers were linear (r 2 > 0.998) in the concentration range 1.03–1030 ng/mL. The inter‐ and intraday precisions were in the range 2.20–13.3 and 8.03–12.1% and 3.80–14.4 and 8.97–13.6% for (+)‐OTX015 and (−)‐OTX015, respectively. Both the enantiomers were found to be stable in a battery of stability studies. This novel method has been applied to the study of stereoselective oral pharmacokinetics of (−)‐OTX015 and unequivocally demonstrated that (−)‐OTX015 does not undergo chiral inversion to its antipode in vivo in mice.  相似文献   

13.
The reaction of 1‐(trimethylsilyloxy)cyclopentene ( 9 ) with (±)‐1,3,5‐triisopropyl‐2‐(1‐(RS)‐{[(1E)‐2‐methylpenta‐1,3‐dienyl]oxy}ethyl)benzene ((±)‐ 4a ) in SO2/CH2Cl2 containing (CF3SO2)2NH, followed by treatment with Bu4NF and MeI gave a 3.0 : 1 mixture of (±)‐(2RS)‐2{(1RS,2Z,4SR)‐2‐methyl‐4‐(methylsulfonyl)‐1‐[(RS)‐1‐(2,4,6‐triisopropylphenyl)ethoxy]pent‐2‐en‐1‐yl}cyclopentanone ((±)‐ 10 ) and (±)‐(2RS)‐2‐{(1RS,2Z)‐2‐methyl‐4‐[(SR)‐methylsulfonyl]‐1‐[(SR)‐1‐(2,4,6‐triisopropylphenyl)ethoxy]pent‐2‐en‐1‐yl}cyclopentanone ((±)‐ 11 ). Similarly, enantiomerically pure dienyl ether (−)‐(1S)‐ 4a reacted with 1‐(trimethylsilyloxy)cyclohexene ( 12 ) to give a 14.1 : 1 mixture of (−)‐(2S)‐2‐{(1S,2Z,4R)‐2‐methyl‐4‐(methylsulfonyl)‐1‐[(S)‐1‐(2,4,6‐triisopropylphenyl)ethoxy]pent‐2‐enyl}cyclohexanone ((−)‐ 13a ) and its diastereoisomer 14a with (1S,2R,4R) or (1R,2S,4S) configuration. Structures of (±)‐ 10 , (±)‐ 11 , and (−)‐ 13a were established by single‐crystal X‐ray crystallography. Poor diastereoselectivities were observed with the (E,E)‐2‐methylpenta‐1,3‐diene‐1‐ylethers (+)‐ 4b and (−)‐ 4c bearing ( 1 S )‐1‐phenylethyl and (1S)‐1‐(pentafluorophenyl)ethyl groups instead of the Greene's auxiliary ((1S)‐(2,4,6‐triisopropylphenyl)ethyl group). The results demonstrate that high α/βsyn and asymmetric induction (due to the chiral auxiliary) can be obtained in the four‐component syntheses of the β‐alkoxy ketones. The method generates enantiomerically pure polyfunctional methyl sulfones bearing three chiral centers on C‐atoms and one (Z)‐alkene moiety.  相似文献   

14.
Complete resolution of hydroxyeicosatetraenoic acid (HETE) enantiomers was achieved using hydroxypropyl‐γ‐cyclodextrin (HP‐γ‐CD)‐modified MEKC. The optimum running conditions were determined to be utilizing a 30 mM phosphate–15 mM borate buffer (pH 9.0) containing 30 mM HP‐γ‐CD and 75 mM SDS as the BGE, application of +30 kV as the effective voltage, and carrying out the experiment at 15°C. The eluents were detected at 235 nm. The method was used successfully for the simultaneous separations of (S)‐ and (R)‐enantiomers of regioisomeric 8‐, 11‐, 12‐, and 15‐HETEs. Subsequently, the optimized method was applied to evaluate the stereochemistry of 8‐ and 12‐HETEs from the marine red algae, Gracilaria vermiculophylla and Gracilaria arcuata, respectively. The 8‐HETE was found to be a mixture of 98% (R)‐enantiomer and 2% (S)‐enantiomer, while the 12‐HETE was a mixture of 98% (S)‐enantiomer and 2% (R)‐enantiomer. The present study demonstrates that the HP‐γ‐CD‐modified MEKC method is simple and sensitive and provides unambiguous information on the configuration of natural and synthetic HETEs.  相似文献   

15.
The cyclic [2R,S(R)]‐bornane‐10,2‐sulfinamide (−)‐ 2b , an analogue of Oppolzer`s camphor‐derived sultam (−)‐ 2a , was synthesized by reduction of the known N‐alkylidenesulfinamide (+)‐ 1b with NaBH4. The uncatalyzed [4+2] cycloaddition of cyclopentadiene to the methyl ester (−)‐ 3b of the N‐fumaroylsulfinamide, obtained from (−)‐ 2b , proceeds with lower endo and π‐facial selectivity as compared to dienophiles (−)‐ 3a , c . In contrast to these latter, the diastereoselectivity is reversed either in apolar CCl4 or in the presence of TiCl4. This inversion is explained by a competitive C(α)‐si addition on the reactive anti‐s‐trans conformer.  相似文献   

16.
Enantiomerically pure (+)‐(1S,4S,5S,6S)‐6‐endo‐(benzyloxy)‐5‐exo‐{[(tert‐butyl)dimethylsilyl]oxy}‐7‐oxabicyclo[2.2.1]heptan‐2‐one ((+)‐ 5 ) and its enantiomer (−)‐ 5 , obtained readily from the Diels‐Alder addition of furan to 1‐cyanovinyl acetate, can be converted with high stereoselectivity into 8‐oxabicyclo[3.2.1]octane‐2,3,4,6,7‐pentol derivatives (see 23 – 28 in Scheme 2). A precursor of them, (1R,2S,4R,5S,6S,7R,8R)‐7‐endo‐(benzyloxy)‐8‐exo‐hydroxy‐3,9‐dioxatricyclo[4.2.1.02,4]non‐5‐endo‐yl benzoate ((−)‐ 19 ), is transformed into (1R,2R,5S, 6S,7R,8S)‐6‐exo,8‐endo‐bis(acetyloxy)‐2‐endo‐(benzyloxy)‐4‐oxo‐3,9‐dioxabicyclo[3.3.1]non‐7‐endo‐yl benzoate ((−)‐ 43 ) (see Scheme 5). The latter is the precursor of several protected 2,6‐anhydrohepturonic acid derivatives such as the diethyl dithioacetal (−)‐ 57 of methyl 3,5‐di‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐D ‐glycero‐D ‐galacto‐hepturonate (see Schemes 7 and 8). Hydrolysis of (−)‐ 57 provides methyl 3,5‐di‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐D ‐glycero‐D ‐galacto‐hepturonate 48 that undergoes highly diastereoselective Nozaki‐Oshima condensation with the aluminium enolate resulting from the conjugate addition of Me2AlSPh to (1S,5S,6S,7S)‐7‐endo‐(benzyloxy)‐6‐exo‐{[(tert‐butyl)dimethylsilyl]oxy}‐8‐oxabicyclo[3.2.1]oct‐3‐en‐2‐one ((−)‐ 13 ) derived from (+)‐ 5 (Scheme 12). This generates a β‐C‐mannopyranoside, i.e., methyl (7S)‐3,5‐di‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐7‐C‐[(1R,2S,3R,4S,5R,6S,7R)‐6‐endo‐(benzyloxy)‐7‐exo‐{[(tert‐butyl)dimethylsilyl]oxy}‐4‐endo‐hydroxy‐2‐exo‐(phenylthio)‐8‐oxabicyclo[3.2.1]oct‐3‐endo‐yl]‐L ‐glycero‐D ‐manno‐heptonate ((−)‐ 70 ; see Scheme 12), that is converted into the diethyl dithioacetal (−)‐ 75 of methyl 3‐O‐acetyl‐2,6‐anhydro‐4,5‐dideoxy‐4‐C‐{[methyl (7S)‐3,5,7‐tri‐O‐acetyl‐2,6‐anhydro‐4‐O‐benzoyl‐L ‐glycero‐D ‐manno‐heptonate]‐7‐C‐yl}‐5‐C‐(phenylsulfonyl)‐L ‐glycero‐D ‐galacto‐hepturonate ( 76 ; see Scheme 13). Repeating the Nozaki‐Oshima condensation to enone (−)‐ 13 and the aldehyde resulting from hydrolysis of (−)‐ 75 , a (1→3)‐C,C‐linked trisaccharide precursor (−)‐ 77 is obtained.  相似文献   

17.
A highly enantiospecific, azide‐free synthesis of (?)‐(R)‐ and (+)‐(S)‐piperidin‐3‐ol in excellent yield was developed. The key step of the synthesis involves the enantiospecific ring openings of enantiomerically pure (R)‐ and (S)‐2‐(oxiran‐2‐ylmethyl)‐1H‐isoindole‐1,3(2H)‐diones with the diethyl malonate anion and subsequent decarboxylation.  相似文献   

18.
Enantio‐ and stereoselective cyclopolymerization of hexa‐1,5‐diene was achieved by enantiomerically pure dichloro zirconium(IV) pre‐catalysts 2 possessing chiral [OSSO]‐type bis(phenolate) ligands (−)‐ 1 and (+)‐ 1 in combination with dried methylaluminoxane (dMAO) as an activator. The corresponding activities were recorded with quite high values up to 1,960 g mmol( 2 )–1 h–1, which are extremely larger than those of the related complexes. The microstructure analysis for the PMCPs furnished by pre‐catalysts (Λ,S,S)‐ 2 and (Δ,R,R)‐ 2 showed good isotacticity factors (α = 75−78%) and relatively high proportions of trans‐cyclopentane rings (σ = 14−21%). These enantiomeric PMCPs exhibited large specific optical rotations ([α]D = +28 to +32° from (Λ,S,S)‐ 2 , −26 to −34° from (Δ,R,R)‐ 2 ).

  相似文献   


19.
《Tetrahedron: Asymmetry》2005,16(17):2954-2958
Laterally lithiated (S)-(−)- and (R)-(+)-o-toluamides 6 with a chiral auxiliary derived from (S)- and (R)-phenylalaninol, respectively, were used as the building blocks and chirality inductors in the asymmetric modification of the Pomeranz–Fritsch–Bobbitt synthesis of isoquinoline alkaloids. Their addition to imine 2 proceeded with partial cyclization, giving isoquinolones (+)-7 and (−)-7 along with acyclic products, (−)-8 and (+)-8, respectively. LAH-reduction of (+)-7 and (−)-7, followed by cyclization, afforded both enantiomers of the alkaloid, (S)-(−)- and (R)-(+)-O-methylbharatamine 5, in 32% and 40% overall yield and with 88% and 73% ees, respectively.  相似文献   

20.
Optical resolution of racemic 5‐oxo‐1‐phenyl‐pyrazolidine‐3‐carboxylic acid 2 with L‐amino acid methyl ester via the diastereomers formation was investigated. Treatment of racemic 5‐oxo‐1‐phenyl‐pyrazolidine‐3‐carboxylic acid 2 with L‐valine methyl ester gave diastereomers with a total yield of 86%. The diastereomeric dipeptides can be easily separated by flash column chromatography. Acidic cleavage of the derived diastereomers gave both the optically pure (+)‐(R)‐ and (‐)‐(S)‐5‐oxo‐1‐phenyl‐pyrazolidine‐3‐carboxylic acid ((+)‐(R)‐ 2 and (‐)‐(S)‐ 2 ) with a total yield of 94% and 95%, respectively.  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