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.
To determine the eutomers of potent GluN2B‐selective N‐methyl‐d ‐aspartate (NMDA) receptor antagonists with a 3‐benzazepine scaffold, 7‐benzyloxy‐3‐(4‐phenylbutyl)‐2,3,4,5‐tetrahydro‐1H‐3‐benzazepin‐1‐ols (S)‐ 2 and (R)‐ 2 were separated by chiral HPLC. Hydrogenolysis and subsequent methylation of the enantiomerically pure benzyl ethers of (S)‐ 2 and (R)‐ 2 provided the enantiomeric phenols (S)‐ 3 and (R)‐ 3 [3‐(4‐phenylbutyl)‐2,3,4,5‐tetrahydro‐1H‐3‐benzazepine‐1,7‐diol] and methyl ethers (S)‐ 4 and (R)‐ 4 . All enantiomers were obtained with high enantiomeric purity (≥99.7 % ee). The absolute configurations were determined by CD spectroscopy. R‐configured enantiomers turned out to be the eutomers in receptor binding studies and two‐electrode voltage clamp experiments. The most promising ligand of this compound series is the R‐configured phenol (R)‐ 3 , displaying high GluN2B affinity (Ki=30 nm ), high inhibition of ion flux (IC50=61 nm ), and high cytoprotective activity (IC50=93 nm ). Whereas the eudismic ratio in the receptor binding assay is 25, the eudismic ratio in the electrophysiological experiment is 3.  相似文献   

2.
(2′S)‐2′‐Deoxy‐2′‐C‐methyluridine and (2′R)‐2′‐deoxy‐2′‐C‐methyluridine were incorporated in the 3′‐overhang region of the sense and antisense strands and in positions 2 and 5 of the seed region of siRNA duplexes directed against Renilla luciferase, whereas (2′S)‐2′‐deoxy‐2′‐C‐methylcytidine was incorporated in the 6‐position of the seed region of the same constructions. A dual luciferase reporter assay in transfected HeLa cells was used as a model system to measure the IC50 values of 24 different modified duplexes. The best results were obtained by the substitution of one thymidine unit in the antisense 3′‐overhang region by (2′S)‐ or (2′R)‐2′‐deoxy‐2′‐C‐methyluridine, reducing IC50 to half of the value observed for the natural control. The selectivity of the modified siRNA was measured, it being found that modifications in positions 5 and 6 of the seed region had a positive effect on the ON/OFF activity.  相似文献   

3.
(R)‐4‐Hydroxymethyl‐2‐phenyl‐2‐oxazoline (R)‐ 1 ) was prepared from (L)‐serine. The respective tosylate ((S)‐ 2 ) was converted into sulfides (S)‐ 4 and (S)‐ 5 , and sulfone (S)‐ 6 , useful starting materials for the elaboration of additional chiral centers. A previously reported [ α]D 25 value for (R)‐ 4 is corrected.  相似文献   

4.
GY Han  GQ Shi  LT Qu  JY Yuan  FE Chen  PY Wu 《Polymer International》2004,53(10):1554-1560
N‐Substituted pyrrole derivatives with chiral side groups have been synthesized and electrochemically polymerized in acetonitrile containing tetrabutylammonium perchlorate (TBAClO4) and (S)‐(+)‐camphor‐10‐sulfonic acid ((S)‐(+)‐CSA) or (R)‐(?)‐camphor‐10‐sulfonic acid ((R)‐(?)‐CSA). The resulting N‐substituted polypyrrole films were characterized by cyclic voltammetry, infrared, Raman and X‐ray photoelectron (XPS) spectroscopies. XPS results demonstrated that the as‐grown polymer films are preferably doped by CSA anions when the monomer and the CSA anion have the same optical rotation dispersion (ORD). Furthermore, the conductivities of the polymers synthesized in the media containing CSA with the same ORD of the corresponding monomers were measured to be about 2–10 times higher than those of polymers obtained from electrolytes without CSA. Copyright © 2004 Society of Chemical Industry  相似文献   

5.
In the presence of Na2CO3 (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐3‐(1,3‐dioxo‐butyl)oxymethyl‐1,2,3,4‐tetrahydrocarboline ( 1 ) were transformed into (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐3‐hydroxymethyl‐1,2,3,4‐tetrahydrocarboline ( 2 ), which were cyclized to (6S)‐3‐acetyl‐6‐hydroxymethyl‐4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizine ( 4 ), via(6S,12bS)‐ and (6S,12bR)‐3‐acetyl‐2‐hydroxyl‐6‐hydroxymethyl‐1,2,3,4,6,7,12,12b‐octahydro‐4‐oxoindolo[2,3‐a]quinoline ( 3 ). (6S)‐ 4 was coupled with Boc‐Gly, Boc‐L‐Asp(β‐benzyl ester), or Boc‐L‐Gln to give 6‐amino acid substituted (6S)‐3‐acetyl‐4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizines 5a , 5b , or 5c , respectively. After the removal of Boc from (6S)‐ 5a (6S)‐3‐acetyl‐6‐glycyl‐4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizine ( 6 ) was obtained. The anticancer activities of (6S)‐ 5 and (6S)‐ 6 in vitro were tested.  相似文献   

6.
Acyclic nucleoside phosphonates (ANPs) are a promising class of antimalarial therapeutic drug leads that exhibit a wide variety of Ki values for Plasmodium falciparum (Pf) and human hypoxanthine‐guanine‐(xanthine) phosphoribosyltransferases [HG(X)PRTs]. A novel series of ANPs, analogues of previously reported 2‐(phosphonoethoxy)ethyl (PEE) and (R,S)‐3‐hydroxy‐2‐(phosphonomethoxy)propyl (HPMP) derivatives, were designed and synthesized to evaluate their ability to act as inhibitors of these enzymes and to extend our ongoing antimalarial structure–activity relationship studies. In this series, (S)‐3‐hydroxy‐2‐(phosphonoethoxy)propyl (HPEP), (S)‐2‐(phosphonomethoxy)propanoic acid (CPME), or (S)‐2‐(phosphonoethoxy)propanoic acid (CPEE) are the acyclic moieties. Of this group, (S)‐3‐hydroxy‐2‐(phosphonoethoxy)propylguanine (HPEPG) exhibits the highest potency for PfHGXPRT, with a Ki value of 0.1 μM and a Ki value for human HGPRT of 0.6 μM . The crystal structures of HPEPG and HPEPHx (where Hx=hypoxanthine) in complex with human HGPRT were obtained, showing specific interactions with active site residues. Prodrugs for the HPEP and CPEE analogues were synthesized and tested for in vitro antimalarial activity. The lowest IC50 value (22 μM ) in a chloroquine‐resistant strain was observed for the bis‐amidate prodrug of HPEPG.  相似文献   

7.
The enantiomers of the potent σ1 ligand fluspidine ( 1 ) were prepared by using chiral preparative HPLC. Synthesis of racemic tosylate 2 and subsequent separation of enantiomers yielded (R)‐ 2 and (S)‐ 2 in excellent enantiomeric purities. The fluspidine enantiomers (R)‐ 1 and (S)‐ 1 were synthesized from (R)‐ 2 and (S)‐ 2 by nucleophilic substitution with tetra‐n‐butylammonium fluoride, affording (R)‐ 1 with 99.6 % ee and (S)‐ 1 with 96.4 % ee. Tosylates (R)‐ 2 and (S)‐ 2 can also serve as precursors for the radiosynthesis of enantiomerically pure radiotracers [18F](R)‐ 1 and [18F](S)‐ 1 . The absolute configuration of the pure enantiomers was elucidated by comparison of their CD spectra with a calculated CD spectrum of a simplified model compound. In receptor binding studies, both enantiomers displayed very high σ1 receptor affinity and selectivity against the σ2 receptor. (R)‐Fluspidine ((R)‐ 1 ) is the eutomer, with a Ki value of 0.57 nM and a eudysmic ratio of 4. Incubation of (R)‐ 1 and (S)‐ 1 with rat liver microsomes led to the identification of seven and eight metabolites, respectively. Although the S‐configured enantiomer formed additional metabolite (S)‐ 1‐3 , it is metabolically more stable than (R)‐ 1 .  相似文献   

8.
A hybrid approach was applied for the design of an inhibitor of trypsin‐like serine proteases. Compound 16 [(R,R)‐ and (R,S)‐diphenyl (4‐(1‐(4‐amidinobenzylamino)‐1‐oxo‐3‐phenylpropan‐2‐ylcarbamoyl)phenylamino)(4‐amidinophenyl)methylphosphonate hydrochloride], prepared in a convergent synthetic procedure, possesses a phosphonate warhead prone to react with the active site serine residue in a covalent, irreversible manner. Each of the two benzamidine moieties of 16 can potentially be accommodated in the S1 pocket of the target enzyme, but only the benzamidine close to the phosphonate group would then promote an irreversible interaction. The Janus‐faced inhibitor 16 was evaluated against several serine proteases and caused a pronounced inactivation of human thrombin with a second‐order rate constant (kinac/Ki) of 59 500 M ?1 s?1. With human matriptase, 16 showed preference for a reversible mode of inhibition (IC50=2.6 μM ) as indicated by linear progress curves and enzyme reactivation.  相似文献   

9.
Methyl (1S,3S and 1R,3S)‐1‐(2, 2‐dimethoxyethyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxylate ( 3 ) was hydrolyzed in the presence of sodium hydroxide to give (1S,3S and 1R,3S)‐1‐(2,2‐dimethoxyethyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxylic acid ( 4 ), which was reduced with LiAlH4 to provide (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐3‐hydroxymethyl‐1,2,3,4‐tetrahydrocarbolines ( 10 ), and then amidated in ammonia containing methanol to obtain (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxamide ( 14 ). Acylation of (1S,3S and 1R,3S)‐ 3 , (1S,3S and 1R,3S)‐ 4 , (1S,3S)‐ 10 , (1R, 3S)‐ 10 , (1S, 3S)‐ 14 and (1R,3S)‐ 14 afforded the corresponding methyl (1S,3S and 1R,3S)‐1‐(2,2‐dimethoxyethyl)‐ 2‐(1,3‐dioxobutyl)‐1,2,3,4‐tetrahydrocarbolines‐3‐carboxylate ( 6 ), (1S,3S and 1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxylic acid ( 5 ), (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐3‐(1,3‐dioxobutyl)oxymethyl‐1,2,3,4‐tetrahydrocarboline ( 11 ), (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxamide ( 15 ), respectively. After Aldol reaction, dehydration and dehydrogenation the desired (6S)‐6‐substituted 4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizines 8 , 9 , 12 , 13 , and 16 were obtained. Their anticancer activities in vitro were investigated.  相似文献   

10.
BACKGROUND: The hydrophobic phenyl n‐propyl ketone was used as a model compound to examine alcohol dehydrogenase activity in Saccharomyces cerevisiae mediated cell culture. Parameters such as pH, hexane‐to‐water volume percentage, and the amount of cofactor Zn2+ ion for either cell growth or reduction were studied to see their effect on the enantioselectivity toward the product R‐(+)‐ or S‐(?)‐1‐phenyl‐1‐butanol. RESULTS: The pH for cell growth in aqueous culture was 7.0, while the pH for reduction in the aqueous portion of the biphasic culture was 5.0. Without Zn2+ ion the biphasic cultures of middle to high hexane‐to‐water volume percentage exhibited an R‐(+)‐1‐phenyl‐1‐butanol enantiomeric excess of 53.7% to > 99%. Without Zn2+ ion the biphasic cultures at low hexane‐to‐water volume percentage possessed an S‐(?)‐1‐phenyl‐1‐butanol enantiomeric excess of 14.5–46.5%. Exclusively, the enantioselectivity for biphasic cultures containing Zn2+ ion was an S‐(?)‐1‐phenyl‐1‐butanol enantiomeric excess of 27.5% to > 99%. Reduction mediated in aqueous culture with varied amount of Zn2+ ion by the yeast Candida utilis also showed an S‐(?)‐1‐phenyl‐1‐butanol enantiomeric excess of 79.2–95.4%. CONCLUSION: The enantioselectivity of S. cerevisiae mediated biphasic culture reduction of phenyl n‐propyl ketone can be manipulated through the cofactor Zn2+ ion and the hexane volume percentage of the biphasic culture. Copyright © 2008 Society of Chemical Industry  相似文献   

11.
Extension of a structure–activity relationship study of the antitumor cytotoxicity of lipidic dialkynylcarbinols (DACs) is envisaged by formal methinylogation of one of the ethyndiyl moieties of the DAC warhead into the corresponding allenylalkynylcarbinol (AllAC) counterpart. External AllACs were directly obtained by methinylation of the parent DACs with formaldehyde in either the racemic or scalemic series. Isomers containing external progargyl and propynyl motifs were also prepared. Internal AllACs were obtained as racemic statistical mixtures of stereoisomers in two steps from the key C5‐DAC rac‐TIPS‐C≡C‐CH(OH)‐C≡CH and aldehydes. Kinetic resolution of the (S)‐C5‐DAC in 97 % ee and (R)‐C5‐DAC in 99 % ee was achieved by sequential lipase‐mediated acetylation/hydrolysis using the Candida antartica lipase (Novozyme 435). The four internal AllAC stereoisomers were prepared by asymmetric methinylation with (R)‐ or (S)‐diphenylprolinol as chiral auxiliary. Cytotoxicity assays on HCT116 cancer cells showed that the most active (eutomeric) external or internal AllAC exhibits an S configuration, a fatty chain length of n=12, and a 50 % inhibitory concentration IC50≈1.0 μm .  相似文献   

12.
Baeyer–Villiger monooxygenase (BVMO)‐mediated regiodivergent conversions of asymmetric ketones can lead to the formation of “normal” or “abnormal” lactones. In a previous study, we were able to change the regioselectivity of a BVMO by mutation of the active‐site residues to smaller amino acids, which thus created more space. In this study, we demonstrate that this method can also be used for other BVMO/substrate combinations. We investigated the regioselectivity of 2‐oxo‐Δ3‐4,5,5‐trimethylcyclopentenylacetyl‐CoA monooxygenase from Pseudomonas putida (OTEMO) for cis‐bicyclo[3.2.0]hept‐2‐en‐6‐one ( 1 ) and trans‐dihydrocarvone ( 2 ), and we were able to switch the regioselectivity of this enzyme for one of the substrate enantiomers. The OTEMO wild‐type enzyme converted (?)‐ 1 into an equal (50:50) mixture of the normal and abnormal products. The F255A/F443V variant produced 90 % of the normal product, whereas the W501V variant formed up to 98 % of the abnormal product. OTEMO F255A exclusively produced the normal lactone from (+)‐ 2 , whereas the wild‐type enzyme was selective for the production of the abnormal product. The positions of these amino acids were equivalent to those mutated in the cyclohexanone monooxygenases from Arthrobacter sp. and Acinetobacter sp. (CHMOArthro and CHMOAcineto) to switch their regioselectivity towards (+)‐ 2 , which suggests that there are hot spots in the active site of BVMOs that can be targeted with the aim to change the regioselectivity.  相似文献   

13.
In this study we explored the pharmaceutically underexploited ATPase domain of DNA gyrase (GyrB) as a potential platform for developing novel agents that target Mycobacterium tuberculosis. In this effort a combination of ligand‐ and structure‐based pharmacophore modeling was used to identify structurally diverse small‐molecule inhibitors of the mycobacterial GyrB domain based on the crystal structure of the enzyme with a pyrrolamide inhibitor (PDB ID: 4BAE ). Pharmacophore modeling and subsequent in vitro screening resulted in an initial hit compound 5 [(E)‐5‐(5‐(2‐(1H‐benzo[d]imidazol‐2‐yl)‐2‐cyanovinyl)furan‐2‐yl)isophthalic acid; IC50=4.6±0.1 μm ], which was subsequently tailored through a combination of molecular modeling and synthetic chemistry to yield the optimized lead compound 24 [(E)‐3‐(5‐(2‐cyano‐2‐(5‐methyl‐1H‐benzo[d]imidazol‐2‐yl)vinyl)thiophen‐2‐yl)benzoic acid; IC50=0.3±0.2 μm ], which was found to display considerable in vitro efficacy against the purified GyrB enzyme and potency against the H37Rv strain of M. tuberculosis. Structural handles were also identified that will provide a suitable foundation for further optimization of these potent analogues.  相似文献   

14.
The microbial transformations of (+)‐ and (?)‐bornyl acetate were investigated using the plant parasitic fungus, Glomerella cingulata. As a result, (+)‐ and (?)‐bornyl acetate were converted to (+)‐ and (?)‐5‐exo‐hydroxybornyl acetate, (+)‐ and (?)‐5‐oxobornyl acetate and (+)‐ and (?)‐borneol respectively. The structures of the metabolic products were determined by spectroscopic data. © 2001 Society of Chemical Industry  相似文献   

15.
The diastereomeric 1,4‐diphosphine ligands, (S,S,S,S)‐ 1a , (R,S,S,R)‐ 1b and (R,S,S,S)‐ 1c , with the imidazolidin‐2‐one backbone were synthesized, and utilized for an investigation of the effects of backbone chirality on the enantioselectivity in the Rh(I)‐catalyzed hydrogenation of various functionalized olefinic substrates. It was found that the catalytic efficiencies are largely dependent on the configurations of the α‐carbons to phosphine. Thus, the Rh complex of the pseudo‐C2‐symmetrical diphosphine, (R,S,S,S)‐ 1c , showed excellent enantioselectivities (93.0–98.6% ees) in the hydrogenations of a broad spectrum of substrates, and especially in the hydrogenations of methyl α‐(N‐acetyamino)‐β‐arylacrylates (95.3–97.0% ees). However, the enantioselectivities obtained with the C2‐symmetrical (R,S,S,R)‐ 1b were largely dependent on the substrate (19.8–97.3% ees). The Rh complex of ligand 1a having the (S,S,S,S)‐configuration showed the lowest catalytic efficiency for all of the substrates examined (0–84.8% ees).  相似文献   

16.
Concurrent inhibition of aromatase and steroid sulfatase (STS) may provide a more effective treatment for hormone‐dependent breast cancer than monotherapy against individual enzymes, and several dual aromatase–sulfatase inhibitors (DASIs) have been reported. Three aromatase inhibitors with sub‐nanomolar potency, better than the benchmark agent letrozole, were designed. To further explore the DASI concept, a new series of letrozole‐derived sulfamates and a vorozole‐based sulfamate were designed and biologically evaluated in JEG‐3 cells to reveal structure–activity relationships. Amongst achiral and racemic compounds, 2‐bromo‐4‐(2‐(4‐cyanophenyl)‐2‐(1H‐1,2,4‐triazol‐1‐yl)ethyl)phenyl sulfamate is the most potent DASI (aromatase: IC50=0.87 nM ; STS: IC50=593 nM ). The enantiomers of the phenolic precursor to this compound were separated by chiral HPLC and their absolute configuration determined by X‐ray crystallography. Following conversion to their corresponding sulfamates, the S‐(+)‐enantiomer was found to inhibit aromatase and sulfatase most potently (aromatase: IC50=0.52 nM ; STS: IC50=280 nM ). The docking of each enantiomer and other ligands into the aromatase and sulfatase active sites was also investigated.  相似文献   

17.
A search for the large‐scale preparation of (5S)‐5,6‐(isopropylidenedioxy)‐3‐oxohexanoates ( 2 ) – a key intermediate in the synthesis of pharmacologially important statins – starting from (S)‐malic acid is described. The synthesis of the required initial compound methyl (3S)‐3,4‐(isopropylidenedioxy)butanoate ( 1 ) by Moriwake’s reduction of dimethyl (S)‐malate ( 3 ) has been improved. Direct 2‐C chain elongation of ester 1 using the lithium enolate of tert‐butyl acetate has been shown to be successful at a 3‐ to 5‐fold excess of the enolate. Unfortunately, the product, tert‐butyl (5S)‐5,6‐(isopropylidenedioxy)‐3‐oxohexanoate ( 2a ) is unstable during distillation. Ethyl (5S)‐5,6‐(isopropylidenedioxy)‐3‐oxohexanoate ( 2b ) was prepared alternatively on a multigram scale from (3S)‐3,4‐(isopropylidenedioxy)butanoic acid ( 7 ) by activation with N,N′‐carbonyldiimidazole and subsequent reaction with Mg(OOCCH2COOEt)2. A convenient pathway for the in situ preparation of the latter is also described. Ethyl ester ( 2b ) can be advantageously purified by distillation. The stereochemistry of the catalytic hydrogenation of β‐keto ester ( 2b ) to ethyl (5S)‐5,6‐(isopropylidenedioxy)‐3‐hydrohyhexanoate (syn‐ 6 and anti‐ 6 ) has been studied using a number of homogeneous achiral and chiral Rh(I) and Ru(II) complexes with phosphine ligands. A comparison of Rh(I) and Ru(II) catalysts with (S)‐ and (R)‐BINAP as chiral ligands revealed opposite activity in dependence on the polarity of the solvent. No influence of the chiral backbone of substrate 2b on the enantioselectivity was noted. A ratio of syn‐ 6 /anti‐ 6 =2.3 was observed with an achiral (Ph3P)3RuCl2 catalyst. Ru[(R)‐Tol‐BINAP]Cl2 neutralized with one equivalent of AcONa afforded the most efficient catalytic system for the production of optically pure syn‐(5S)‐5,6‐isopropylidenedioxy‐3‐hydroxyhexanoate (syn‐ 6 ) at a preparative substrate/catalyst ratio of 1000:1.  相似文献   

18.
The enantioselective cobalt‐catalyzed [6+2] cycloadditions of cycloheptatriene 1 with alkynes 2 is reported. Chiral phosphoramidites based on 3,3′‐disubstituted (R)‐BINOL appeared to be efficient ligands, affording the corresponding cycloadducts with good yields and up to 92 % ee. A vibrational circular dichroism study afforded the absolute configuration of new chiral (+)‐(1S,6R)‐7‐phenyl[4.2.1]bicyclo‐ nonatriene 3a and (−)‐(1S,6R)‐7‐trimethylsilyl[4.2.1]bicyclononatriene 3c .  相似文献   

19.
Palladium‐catalysed monophosphorylation of (R)‐2,2′‐bisperfluoroalkanesulfonates of BINOL (RF=CF3 or C4F9) by a diaryl phosphinate [Ar2P(O)H] followed by phosphine oxide reduction (Cl3SiH) then lithium diisopropylamide‐mediated anionic thia‐Fries rearrangement furnishes enantiomerically‐pure (R)‐2′‐diarylphosphino‐2′‐hydroxy‐3′‐perfluoralkanesulfonyl‐1,1′‐binaphthalenes [(R)‐ 8ab and (R)‐ 8g–j ], which can be further diversified by Grignard reagent (RMgX)‐mediated CF3‐displacement [→(R)‐ 8c–f ]. Coupling of (R)‐ 8a–j with (S)‐1,1′‐binaphthalene‐2,2′‐dioxychlorophosphine (S)‐ 9 generates 3′‐sulfonyl BINAPHOS ligands (R,S)‐ 10a–j in good yields (43–82%). These new ligands are of utlility in the asymmetric hydrophosphonylation of styrene ( 1 ) by 4,4,5,5‐tetramethyl‐1,3,2‐dioxaphospholane 2‐oxide ( 2 ), for which a combination of the chiral ligands with either [Pd(Cp)(allyl)] or [Pd(allyl)(MeCN)2]+/NaCH(CO2Me)2 proves to be a convenient and active pre‐catalyst system. A combination of an electron‐rich phosphine moiety and an electron‐deficient 3′‐sulfone moiety provides the best enantioselectivity to date for this process, affording the branched 2‐phenethenephosphonate, (−)‐iso‐ 3 , in up to 74% ee with ligand (R,S)‐ 10i , where Ar=p‐anisyl and the 3′‐SO2R group is triflone.  相似文献   

20.
2,3‐Benzodiazepine derivatives : 1‐(4‐Aminophenyl)‐3,5‐dihydro‐3‐N‐ethylcarbamoyl‐5‐methyl‐7,8‐methylenedioxy‐4H‐2,3‐benzodiazepin‐4‐one was synthesized, and its enantiomers were separated by chiral HPLC. Pharmacological evaluation of each enantiomer showed that (S)‐(?)‐ 5 appears to be more potent than its optical antipode (R)‐(+)‐ 5 in an AMPA receptor binding assay.

  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