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.
A bis‐cyclometalated chiral‐at‐metal rhodium complex catalyzes the Diels–Alder reaction between N‐Boc‐protected 3‐vinylindoles (Boc=tert‐butyloxycarbonyl) and β‐carboxylic ester‐substituted α,β‐unsaturated 2‐acyl imidazoles with good‐to‐excellent regioselectivity (up to 99:1) and excellent diastereoselectivity (>50:1 d.r.) as well as enantioselectivity (92–99 % ee) under optimized conditions. The rhodium catalyst serves as a chiral Lewis acid to activate the 2‐acyl imidazole dienophile by two‐point binding and overrules the preferred regioselectivity of the uncatalyzed reaction.  相似文献   

2.
Asymmetric catalysis under almost‐neutral reaction conditions is key for the efficient synthesis of optically active polar molecules. We have developed catalytic enantioselective reactions of acyclic or cyclic alkenyl esters by using an (S)‐BINOL‐derived chiral tin‐dibromide reagent that possesses a bulky aryl group at the 3 or 3′ position as the chiral pre‐catalyst in the presence of a sodium alkoxide and an alcohol, in which a chiral tin alkoxide bromide is generated in situ and recycled with the assistance of an alcohol. In this Personal Account, we describe three types of asymmetric transformation that proceed through a chiral tin enolate: 1) The asymmetric aldol reaction of alkenyl esters or unsaturated lactones with aldehydes or isatins; 2) the asymmetric three‐component Mannich‐type reaction of alkenyl esters and related cycloaddition reactions; and 3) the asymmetric N‐nitroso aldol reaction of unsaturated lactones with nitrosoarenes.  相似文献   

3.
Chemoselective control over N/O selectivity is an intriguing issue in nitroso chemistry. Recently, we reported an unprecedented asymmetric α‐amination reaction of β‐ketocarbonyl compounds that proceeded through the catalytic coupling of enamine carbonyl groups with in‐situ‐generated carbonyl nitroso moieties. This process was facilitated by a simple chiral primary and tertiary diamine that was derived from tert‐leucine. This reaction featured high chemoselectivity and excellent enantioselectivity for a broad range of substrates. Herein, a computational study was performed to elucidate the origins of the enantioselectivity and N/O regioselectivity. We found that a bidentate hydrogen‐bonding interaction between the tertiary N+? H and nitrosocarbonyl groups accounted for the high N selectivity, whilst the enantioselectivity was determined by Si‐facial attack on the (E)‐ and (Z)‐enamines in a Curtin–Hammett‐type manner. The bidentate hydrogen‐bonding interaction with the nitrosocarbonyl moieties reinforced the facial selectivity in this process.  相似文献   

4.
The steric effects of substituents on five‐membered rings are less pronounced than those on six‐membered rings because of the difference in bond angles. Thus, the regioselectivities of reactions of five‐membered heteroarenes that occur with selectivities dictated by steric effects, such as the borylation of C?H bonds, have been poor in many cases. We report that the silylation of five‐membered‐ring heteroarenes occurs with high sterically derived regioselectivity when catalyzed by the combination of [Ir(cod)(OMe)]2 (cod=1,5‐cyclooctadiene) and a phenanthroline ligand or a new pyridyl‐imidazoline ligand that further increases the regioselectivity. The silylation reactions with these catalysts produce high yields of heteroarylsilanes from functionalization at the most sterically accessible C?H bonds of these rings under conditions that the borylation of C?H bonds with previously reported catalysts formed mixtures of products or products that are unstable. The heteroarylsilane products undergo cross‐coupling reactions and substitution reactions with ipso selectivity to generate heteroarenes that bear halogen, aryl, and perfluoroalkyl substituents.  相似文献   

5.
We report a dual function asymmetric catalysis by a chiral phosphoric acid catalyst that controls both enantioselective addition of an achiral α‐vinyl allylboronate to aldehydes and pseudo‐axial orientation of the α‐vinyl group in the transition state. The reaction produces dienyl homoallylic alcohols with high Z‐selectivities and enantioselectivities. Computational studies revealed that minimization of steric interactions between the alkyl groups of the diol on boron and the chiral phosphoric acid catalyst influence the orientation of α‐vinyl substituent of the allylboronate reagent to occupy a pseudo‐axial position in the transition state.  相似文献   

6.
The preparation of three new chiral derivatives of DMPU (N,N′‐dimethylpropyleneurea) is described (Schemes 24); one type of derivative carries 1‐phenylethyl or 1‐cyclohexylethyl groups at the N‐atoms of the tetrahydropyrimidin‐2(1H)‐one ring ( 2 and 4 ), another type of derivative is substituted at C(4) and C(6) of the heterocyclic ring ( 7 ). The potential of these chiral Lewis bases as promoters in the regio‐ and/or enantioselective addition of 2‐(1,3‐dithianyl)lithium to cyclohex‐2‐en‐1‐one was explored; they are all unable to effect enantioselective addition; the derivatives with branched substituents at the N‐atoms do not shift the addition mode from 1,2 to 1,4, while the 3,4,5,6‐tetrahydro‐1,3,4,6‐tetramethylpyrimidin‐2(1H)‐one does (Scheme 5). The results provide useful information regarding the nature of the nucleophilic organolithium reagent: obviously, the steric hindrance to Li complexation on the CO O‐atom of the tetrahydropyrimidin‐2(1H)‐one by branched substituents at N‐atoms (cf. X‐ray crystal structure of 2 in the Fig.) prevents solvent‐separated‐ion‐pair (SSIP) formation; this was confirmed by PM3 and B3LYP/3‐21‐G(d)//PM3 calculations (Scheme 6).  相似文献   

7.
Chiral, enantiomerically pure Pd‐catalysts were used to control the regioselectivity of nucleophilic attack in allylic substitutions with optically active 1,3‐disubstituted allyl acetates (Schemes 4 – 6). In contrast to reactions with achiral catalysts, where the regioselectivity is determined by the steric and electronic effects of the allylic substituents, chiral catalysts allow selective preparation of either one of the two regioisomeric products, depending on which enantiomer of the catalyst is employed. It is not necessary to start from an enantiomerically pure substrate, because the major and minor enantiomers are converted to different regioisomers (not to enantiomeric products; see Scheme 3), resulting in products of very high ee, even when the starting material is only of moderate enantiomer purity.  相似文献   

8.
A novel method for the catalytic asymmetric dearomatization by visible‐light‐activated [2+2] photocycloaddition with benzofurans and one example of a benzothiophene is reported, thereby providing chiral tricyclic structures with up to four stereocenters including quaternary stereocenters. The benzofurans and the benzothiophene are functionalized at the 2‐position with a chelating N‐acylpyrazole moiety which permits the coordination of a visible‐light‐activatable chiral‐at‐rhodium Lewis acid catalyst. Computational molecular modeling revealed the origin of the unusual regioselectivity and identified the heteroatom in the heterocycle to be key for the regiocontrol.  相似文献   

9.
A chiral imidazolidine‐containing NCN/Pd‐OTf catalyst ( C4 ) promoted the nucleophilic addition of unprotected indoles to N‐Boc imines. Using sulfinyl amines as the N‐Boc imine precursors, the combined use of C4 with K2CO3 activated the NH indoles to give chiral 3‐indolyl methanamines with up to 98 % ee. Compared with conventional acid‐catalyzed Friedel–Crafts reactions, this reaction proceeds under mildly basic conditions and is advantageous for the use of acid‐sensitive substrates.  相似文献   

10.
Proton abstraction of Ntert‐butoxycarbonyl‐piperidine (N‐Boc‐piperidine) with sBuLi and TMEDA provides a racemic organolithium that can be resolved using a chiral ligand. The enantiomeric organolithiums can interconvert so that a dynamic resolution occurs. Two mechanisms for promoting enantioselectivity in the products are possible. Slow addition of an electrophile such as trimethylsilyl chloride allows dynamic resolution under kinetic control (DKR). This process occurs with high enantioselectivity and is successful by catalysis with substoichiometric chiral ligand (catalytic dynamic kinetic resolution). Alternatively, the two enantiomers of this organolithium can be resolved under thermodynamic control with good enantioselectivity (dynamic thermodynamic resolution, DTR). The best ligands found are based on chiral diamino‐alkoxides. Using DTR, a variety of electrophiles can be used to provide an asymmetric synthesis of enantiomerically enriched 2‐substituted piperidines, including (after Boc deprotection) the alkaloid (+)‐β‐conhydrine. The chemistry was extended, albeit with lower yields, to the corresponding 2‐substituted seven‐membered azepine ring derivatives.  相似文献   

11.
Copper‐catalyzed enantioselective allyl–allyl coupling between allylboronates and either Z‐acyclic or cyclic allylic phosphates using a new chiral N‐heterocyclic carbene ligand, bearing a phenolic hydroxy, is reported. This reaction occurs with exceptional SN2′‐type regioselectivities and high enantioselectivities to deliver chiral 1,5‐diene derivatives with a tertiary stereogenic center at the allylic/homoallylic position.  相似文献   

12.
Asymmetric benzannulation of 1‐arylalkynes, where the aryl group is an ortho‐substituted aromatic group, with 2‐(cyanomethyl)phenylboronate was catalyzed by a rhodium complex coordinated with a chiral diene ligand to give high yields of axially chiral 2‐aminobiaryls with greater than 90 % ee.  相似文献   

13.
Chloroprene (=2‐chlorobuta‐1,3‐diene; 4b ) and electron‐rich dienes such as 2‐methoxy‐( 4c ), 2‐acetoxy‐( 4d ), and 2‐(phenylseleno)buta‐1,3‐diene ( 4e ) refused to equilibrate with the corresponding sultines 5 or 6 between −80 and −10° in the presence of excess SO2 and an acidic promoter. Isoprene ( 4a ) and 2‐(triethylsilyl)‐( 4f ), 2‐phenyl‐( 4g ), and 2‐(2‐naphthyl)buta‐1,3‐diene ( 4i ) underwent the hetero‐Diels‐Alder additions with SO2 at low temperature. In contrast, 2‐(1‐naphthyl)buta‐1,2‐diene ( 4h ) did not. With dienes 4a, 4g , and 4i , the hetero‐Diels‐Alder additions with SO2 gave the corresponding 4‐substituted sultine 5 with high regioselectivity. In the case of 4g +SO2⇄ 5g , the energy barrier for isomerization of 5g to 5‐phenylsultine ( 6g ) was similar to that of the cheletropic addition of 4g to give 3‐phenylsulfolene ( 7g ). The hetero‐Diels‐Alder addition of 4f gave a 1 : 4 mixture of the 4‐(triethylsilyl)sultine ( 5f ) and 5‐(triethylsilyl)sultine ( 6f ). The preparation of the two new dienes 4h and 4i is reported.  相似文献   

14.
Why are linear (E)‐enones such challenging substrates in the Rh‐catalyzed asymmetric arylation with boronic acids, which is one of the most important asymmetric catalysis methods? DFT computations show that these substrates adopt a specific conformation in which the largest substituent is antiperiplanar to RhI π‐complexed with the C?C bond within the enantioselectivity‐determining carborhodation transition state. Additionally, for such structures, there is a strong, but not exclusive, preference for scis enone conformation. This folding minimizes steric interactions between the substrate and the ligand, and hence reduces the enantioselectivity. This idea is further confirmed by investigating three computation‐only substrate “probes”, one of which is capable of double asymmetric induction, and a recent computationally designed 1,5‐diene ligand. On average, excellent agreement between predicted and experimental enantioselectivity was attained by a three‐pronged approach: 1) thorough conformational search within ligand and substrate subunits to locate the most preferred carborhodation transition state; 2) including dispersion interaction and long‐range corrections by SMD/ωB97xD/DGDZVP level of theory; and 3) full substrate and ligand modeling. Based on the results, a theory‐enhanced enantioselectivity model that is applicable to both chiral diene and diphosphane ligands is proposed.  相似文献   

15.
The regioselectivity of the interaction between isomeric 4,5‐dihydro‐1,4‐benzothiazepin‐3(2H)‐one 1,1‐dioxide and 1,5‐dihydro‐4,1‐benzothiazepin‐2(3H)‐one 4,4‐dioxide derivatives with the Vilsmeier reagent and DMFDMA (N,N‐dimethylformamide dimethylacetal) has been investigated. The structures of synthesized compounds are confirmed by 1H, 13C NMR, elemental analysis, and X‐ray data.  相似文献   

16.
An HPLC method was established for enantioseparation of (R,S)‐atenolol (ATE) and determination of enantiomers in rat plasma. Marfey's reagent (1‐fluoro‐2,4‐dinitrophenyl‐5‐L‐alanine amide, FDNP‐L‐Ala‐NH2, MR) was used as chiral derivatizing reagent with detection of diastereomers at 340 nm. It was shown that the R‐isomer eluted before the S‐isomer. The method was validated for linearity, repeatability, limits of detection and limit of quantification (LOQ). Recovery of ATE at LOQ was 92.8% for (R)‐ATE and 92.6% for (S)‐ATE. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
Bifunctional thiourea 1 a catalyzes aza‐Henry reaction of nitroalkanes with N‐Boc‐imines to give syn‐β‐nitroamines with good to high diastereo‐ and enantioselectivity. Apart from the catalyst, the reaction requires no additional reagents such as a Lewis acid or a Lewis base. The N‐protecting groups of the imines have a determining effect on the chirality of the products, that is, the reaction of N‐Boc‐imines gives R adducts as major products, whereas the same reaction of N‐phosphonoylimines furnishes the corresponding S adducts. Various types of nitroalkanes bearing aryl, alcohol, ether, and ester groups can be used as nucleophiles, providing access to a wide range of useful chiral building blocks in good yield and high enantiomeric excess. Synthetic versatility of the addition products is demonstrated by the transformation to chiral piperidine derivatives such as CP‐99,994.  相似文献   

18.
The self‐sufficient cytochrome P450 monooxygenase CYP505E3 from Aspergillus terreus catalyzes the regioselective in‐chain hydroxylation of alkanes, fatty alcohols, and fatty acids at the ω‐7 position. It is the first reported P450 to give regioselective in‐chain ω‐7 hydroxylation of C10–C16 n‐alkanes, thereby enabling the one step biocatalytic synthesis of rare alcohols such as 5‐dodecanol and 7‐tetradecanol. It shows more than 70 % regioselectivity for the eighth carbon from one methyl terminus, and displays remarkably high activity towards decane (TTN≈8000) and dodecane (TTN≈2000). CYP505E3 can be used to synthesize the high‐value flavour compound δ‐dodecalactone via two routes: 1) conversion of dodecanoic acid into 5‐hydroxydodecanoic acid (24 % regioselectivity), which at low pH lactonises to δ‐dodecalactone, and 2) conversion of 1‐dodecanol into 1,5‐dodecanediol (55 % regioselectivity), which can be converted into δ‐dodecalactone by horse liver alcohol dehydrogenase.  相似文献   

19.
The ‘t‐amino effect’ of amino‐nitroso compounds was documented by preparing the (dialkylamino)‐nitroso pyrimidines 4 – 18 , and cyclising them under thermal conditions in high yields to the purine derivatives 19 – 32 . The reactivity of the amino‐nitroso‐pyrimidines, particularly of 17 derived from diethyl iminodiacetate, and of 19 , derived from 1‐phenylimidazolidine, correlates with the stability of the intermediate azomethine ylide. Thermolysis of the amino‐nitroso‐pyrimidines 34 – 37 , possessing dialkylamino substituents at C(4) and C(6), proceeded by protiodenitrosation, leading to 38 – 41 .  相似文献   

20.
‘Keteniminium triflates’ (=N‐alk‐1‐enylideneaminium trifluoromethanesulfonates; [MeN(Ts)CHC NR1(R2)]+TfO) generated in situ from Nα‐tosylsarcosinamides MeN(Ts)CH2CONR1(R2) unexpectedly react with cyclopenta‐1,3‐diene and cyclohexa‐1,3‐diene to give Diels–Alder reactions across the CC bond of the cumulene. The use of Nα‐tosylsarcosinamides derived from chiral pyrrolidines allows the direct preparation of trinorbornenone derivative 6 in high enantiomer purity.  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