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 acetylcholinesterase inhibition by enantiomers of exo‐ and endo‐2‐norbornyl‐Nn‐butylcarbamates shows high stereoselelectivity. For the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐exo‐2‐norbornyl‐Nn‐butylcarbamates, the R‐enantiomer is more potent than the S‐enantiomer. But, for the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates, the S‐enantiomer is more potent than the R‐enantiomer. Optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates are synthesized from condensations of optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norborneols with n‐butyl isocyanate, respectively. Optically pure norborneols are obtained from kinetic resolutions of their racemic esters by lipase catalysis in organic solvent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

2.
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   

3.
Three new indole diketopiperazine alkaloids, 11‐methylneoechinulin E and variecolorin M, and (+)‐variecolorin G, along with 12 known analogs, were isolated from a soft coral‐associated epiphytic fungus Aspergillus sp. EGF 15‐0‐3. The structures of the new compounds were unambiguously established by extensive spectroscopic analyses including HR‐ESI‐MS, 1D and 2D NMR spectroscopy and optical rotation measurements. The absolute configurations of (+)‐ and (?)‐variecolorin G were determined by experimental and quantum‐chemical ECD investigations and single‐crystal X‐ray diffraction analysis. Variecolorin G is a pair of enantiomeric mixtures with a ratio of 1 : 2. Moreover, (+)‐neoechinulin A is firstly reported as a natural product. The cytotoxic activities of all the isolated compounds against NCI‐H1975 gefitinib resistance (NCI‐H1975/GR) cell lines were preliminarily evaluated by MTT method.  相似文献   

4.
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   

5.
A new set of optically active 2,3,3‐ and 2,7,7‐trimethyl‐substituted γ‐aminonorbornan‐2‐ols have been obtained from 2‐methylenenorbornane‐1‐carbonitriles derived from (+)‐camphor and (?)‐fenchone and probed as chiral ligands for the enantioselective addition of diethylzinc to benzaldehyde. This has allowed the study of the structural factors influencing the chirality transfer, such as variation of the relative configuration at C(2) and steric hindrance at C(2), C(3), and C(7) positions of norbornane, which result in the observance of the important role played by the gem‐dimethyl position in γ‐aminonorbornan‐2‐exo‐ols. An empirical rationalization of the obtained experimental results has been realized on the basis of energetically‐favored diastereomeric Noyori‐like transition states. Chirality 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

6.
The resolution methods applying (?)‐(4R,5R)‐4,5‐bis(diphenylhydroxymethyl)‐2,2‐dimethyldioxolane (“TADDOL”), (?)‐(2R,3R)‐α,α,α',α'‐tetraphenyl‐1,4‐dioxaspiro[4.5]decan‐2,3‐dimethanol (“spiro‐TADDOL”), as well as the acidic and neutral Ca2+ salts of (?)‐O,O'‐dibenzoyl‐ and (?)‐O,O'‐di‐p‐toluoyl‐(2R,3R)‐tartaric acid were extended for the preparation of 1‐n‐butyl‐3‐methyl‐3‐phospholene 1‐oxide in optically active form. In one case, the intermediate diastereomeric complex could be identified by single‐crystal X‐ray analysis. The absolute P‐configuration of the enantiomers of the phospholene oxide was also determined by comparing the experimentally obtained and calculated CD spectra. Chirality 26:174–182, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
8.
For the first time, a method for enantiomer resolution of the anticonvulsant Galodif (1‐((3‐chlorophenyl)(phenyl)methyl) urea) by chiral HPLC was developed, whereas the enantiomeric composition of 1‐((3‐chlorophenyl)(phenyl)methyl) amine—precursor in Galodif synthesis—cannot be resolved by this method. However, starting 1‐((3‐chlorophenyl)(phenyl)methyl) amine quantitatively forms diastereomeric N‐((3‐chlorophenyl)(phenyl)methyl)‐1‐camphorsulfonamides in reaction with chiral (1R)‐(+)‐ or (1S)‐(?)‐camphor‐10‐sulfonyl chlorides. The diastereomeric ratio of obtained camphorsulfonamides can be easily determined by NMR 1H and 13C spectroscopy. The DFT calculations of specific rotation of Galodif enantiomers showed good agreement with experimental data. The absolute configuration of enantiomers was proposed for the first time.  相似文献   

9.
The biological activity of monoterpenoid indole alkaloids (MIAs) has led to their use in cancer treatment and other medical applications. Their biosynthesis has involved the formation of reactive intermediates by responsible enzymes to elaborate several different chemical scaffolds. Modification of scaffolds through different substitution reactions has produced chemically diverse MIAs and related biological activities. The present study characterizes the three‐step pathway involved in the formation of (+)‐echitovenine, the major O‐acetylated MIA of Catharanthus roseus roots, and differentiates it from a parallel pathway involved in the formation of hörhammericine. Separate hydrolases convert a common reactive MIA intermediate to aspidosperma skeletons of opposite specific rotations, that is (+)‐vincadifformine and (?)‐tabersonine, respectively. The formation of (+) minovincinine from (+) vincadifformine 19‐hydroxylase (V19H) is catalyzed by a root‐specific cytochrome P450 with high amino acid sequence similarity to the leaf‐specific tabersonine‐3‐hydroxylase involved in vindoline biosynthesis. Similarly, O‐acetylation of (+)‐minovincinine to form (+) echitovenine involves minovincinine‐O‐acetytransferase. The substrate specificity of V19H and MAT for their respective (+)‐enantiomers defines the separate enantiomer‐specific pathway involved in (+)‐echitovenine biosynthesis and differentiates it from a parallel (?)‐enantiomer‐specific pathway involved in the formation of hörhammericine from (?)‐tabersonine.  相似文献   

10.
(?)‐5‐Epieremophilene, an epimer of the versatile sesquiterpene (+)‐valencene, is an inaccessible natural product catalyzed by three sesquiterpene synthases (SmSTPSs1‐3) of the Chinese medicinal herb Salvia miltiorrhiza, and its biological activity remains less explored. In this study, three metabolically engineered Escherichia coli strains were constructed for (?)‐5‐epieremophilene production with yields of 42.4–76.0 mg/L in shake‐flask culture. Introducing an additional copy of farnesyl diphosphate synthase (FDPS) gene through fusion expression of SmSTPS1‐FDPS or dividing the FDP synthetic pathway into two modules resulted in significantly improved production, and ultimately 250 mg of (?)‐5‐epieremophilene were achieved. Biological assay indicated that (?)‐5‐epieremophilene showed significant antifeedant activity against Helicoverpa armigera (EC50=1.25 μg/cm2), a common pest of S. miltiorrhiza, implying its potential defensive role in the plant. The results provided an ideal material supply for studying other potential biological activities of (?)‐5‐epieremophilene, and also a strategy for manipulating terpene production in engineered E. coli using synthetic biology.  相似文献   

11.
(R)‐(+) and (S)‐(?)‐1‐phenylethylamine have been shown to promote highly diastereoselective and complementary enantioselective formal [3 + 2]carbocyclization reactions between 2,3‐butanedione and conjugated nitroalkenes with formation of enantiomerically rich 2‐hydroxy‐3‐nitrocyclopentanone derivatives. The reactions were carried out both in solvent and under solvent‐free conditions. The absolute configurations of the products were assigned by X‐ray and circular dichroism spectra analyses. Chirality 24:1005–1012, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

12.
Al Arsh Basheer  Imran Ali 《Chirality》2018,30(9):1088-1095
The environmental stereoselective uptake and degradation of (±)‐o,p‐DDD pesticide stereomers in water‐sediment system are described. The results were analyzed by artificial neural network model. The optimized experimental parameters were concentration of o,p‐DDD streamers (7.0 μg L?1), experimental time (60 min), pH (6), dose (5.0 g L?1), and temperature (25°C). The maximum uptake and degradation were 87% and 85% and 33.0% and 30.5% for (?)‐ and (+)‐stereomers of o,p‐DDD in 15‐day time. Both uptake and degraded phenomenon showed first‐order rate reaction. Thermodynamic variables indicated exothermic nature of uptake and degradation processes. The uptake and degradation were slightly higher for (?)‐stereomer than (+)‐stereomer of o,p‐DDD. It was assumed that both uptake and degradation processes are accountable for the removal of the streomers of o,p‐DDD from earth's ecosystem, but the uptake is responsible for major contribution. The magnitudes of relative errors obtained by artificial neural network model were in the range of ±0.2 to 3.5, indicating good applicability of the experimental data. The results are very useful to control the environmental contamination due to the chiral o,p‐DDD pesticide as its two enantiomers have different ecological toxicities.  相似文献   

13.
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   

14.
Optically active 3,3′‐dimethyl‐2,2′‐diamino‐1,1′‐binaphthyl (DM‐DABN) and 3,3′‐dimethyl‐2‐amino‐2′‐hydroxybinaphthyl (DM‐NOBIN) derivatives were synthesized by Cu‐(?)‐sparteine complex‐catalyzed enantioselective homo‐ and hetero‐coupling of 2‐naphthylamine, respectively. The difference in enantioselectivity was observed by changing the concentration of oxygen. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

15.
Evidence is accumulating to suggest that 3,4‐methylenedioxymethamphetamine (MDMA) has neurotoxic and neuroinflammatory properties. MDMA is composed of two enantiomers with different biological activities. In this study, we evaluated the in vivo effects of S(+)‐MDMA, R(?)‐MDMA, and S(+)‐MDMA in combination with R(?)‐MDMA on microglial and astroglial activation compared with racemic MDMA, by assessment of complement type 3 receptor (CD11b) and glial fibrillary acidic protein (GFAP) immunoreactivity in the mouse striatum, nucleus accumbens, motor cortex, and substantia nigra. Motor activity and body temperature were also measured, to elucidate the physiological modifications paired with the observed glial changes. Similar to racemic MDMA (4 × 20 mg/kg), S(+)‐MDMA (4 × 10 mg/kg) increased both CD11b and GFAP in the striatum, although to a lower degree, whereas R(?)‐MDMA (4 × 10 mg/kg) did not induce any significant glial activation. Combined administration of S(+) plus R(?)‐MDMA did not induce any further activation compared with S(+)‐MDMA. In all other areas, only racemic MDMA was able to slightly activate the microglia, but not the astroglia, whereas enantiomers had no effect, either alone or in combination. Racemic MDMA and S(+)‐MDMA similarly increased motor activity and raised body temperature, whereas R(?)‐MDMA affected neither body temperature nor motor activity. Interestingly, the increase in body temperature was correlated with glial activation. The results show that no synergism, but only additivity of effects, is caused by the combined administration of S(+)‐ and R(?)‐MDMA, and underline the importance of investigating the biochemical and behavioral properties of the two MDMA enantiomers to understand their relative contribution to the neuroinflammatory and neurotoxic effects of MDMA.  相似文献   

16.
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   

17.
Two pairs of new non‐brominated racematic pyrrole derivatives, (±)‐nakamurine D ( 1 ) and (±)‐nakamurine E ( 2 ), two new diterpene alkaloids, isoagelasine C ( 16 ) and isoagelasidine B ( 21 ), together with 13 known pyrrole derivatives ((±)‐ 3  –  15 ), five known diterpene alkaloids ( 17  –  20 , 22 ) were isolated from the South China Sea sponge Agelas nakamurai . The racemic mixtures, compounds 1  –  4 , were resolved into four pairs of enantiomers, (+)‐ 1 and (?)‐ 1 , (+)‐ 2 and (?)‐ 2 , (+)‐ 3 and (?)‐ 3 , and (+)‐ 4 and (?)‐ 4 , by chiral HPLC . The structures and absolute configurations were elucidated on the basis of comprehensive spectroscopic analyses, quantum chemical calculations, quantitative measurements of molar rotations, application of van't Hoff 's principle of optical superposition, and comparison with the literature data. The NMR and MS data of compound 3 are reported for the first time, as the structure was listed in SciFinder Scholar with no associated reference. These non‐brominated pyrrole derivatives were found in this species for the first time. Compound 18 showed valuable cytotoxicities against HL ‐60, K562, and HCT ‐116 cell lines with IC 50 values of 12.4, 16.0, and 19.8 μm , respectively. Compounds 16  –  19 , 21 , and 22 showed potent antifungal activities against Candida albicans with MIC values ranging from 0.59 to 4.69 μg/ml. Compounds 16  –  19 exhibited moderate antibacterial activities against Proteusbacillus vulgaris (MIC values ranging from 9.38 to 18.75 μg/ml).  相似文献   

18.
A simple, efficient, and economical method based on the combination of the exceptional behavior of o,o′‐dibenzoyl‐ or o,o′‐di‐p‐toluyl‐(2R,3R)‐tartaric acid in chiral recognition processes, and the coordination ability of calcium or magnesium ion was developed for the resolution of phospholene oxides 1 . The calcium or magnesium salt of (?)‐o,o′‐dibenzoyl‐(2R,3R)‐tartaric acid 2 , 4 ‐ 6 or calcium hydrogen (?)‐o,o′‐di‐p‐toluyl‐(2R,3R)‐tartrate 3 may form crystalline diastereomeric coordination complexes with the appropriate antipode of substituted 3‐phospholene oxides 1 that makes possible efficient resolutions. Optically active phospholene oxides 1 were prepared directly by simply crystallization and digestion of the corresponding diastereomeric complexes so formed. Thermal behavior of the crystalline diastereomeric complexes was studied by simultaneous TG/DTA. The novel method may be of more general value in respect of the resolution of tertiary phosphine oxides. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

19.
In our effort towards measuring the parity violation energy difference between two enantiomers, a simple chiral oxorhenium complex 5 bearing enantiopure 2‐mercaptocyclohexan‐1‐ol has been prepared as a potential candidate species. Vibrational circular dichroism revealed a chiral environment surrounding the rhenium atom, even though the rhenium is not a stereogenic center itself, and enabled to assign the (1S,2S)‐(?) and (1R,2R)‐(+) absolute configuration for 5 . For both compound 5 and complex 4 , previously studied by us and bearing a propane‐2‐olato‐3‐thiolato ligand, relativistic calculations predict parity violating vibrational frequency differences of a few hundreds of millihertz, above the expected sensitivity attainable by a molecular beam Ramsey interferometer that we are constructing.  相似文献   

20.
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   

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