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 62 毫秒
1.
An anti‐selective Mannich reaction of aldehydes with N‐sulfonyl imines has been developed by using a 4‐hydroxypyrrolidine in combination with an external Brønsted acid. The catalyst design is based on three elements: the α‐substituent of the pyrrolidine, the 4‐hydroxy group, and the Brønsted acid, the combination of which is essential for high chemical and stereochemical efficiency. The reaction works with aromatic aldehyde‐derived imines, which have rarely been employed in previously reported enamine‐based anti‐Mannich reactions. Additionally, both N‐tosyl and N‐nosyl imines can be successfully used and the Mannich adducts can be easily reduced or oxidized, and after N‐deprotection the corresponding β‐amino acids and β‐amino alcohols can be obtained with good yields. The results also show that this ternary catalytic system may be practical in other enamine‐based reactions.  相似文献   

2.
Catalytic asymmetric Michael and Mannich‐type reactions of glycine Schiff bases with chiral two‐center organocatalysts, tartrate‐derived diammonium salts (TaDiASs), are described. On the basis of conformational studies, optimized TaDiASs with a 2,6‐disubstituted cyclohexane spiroacetal were newly designed. These TaDiASs catalyzed the asymmetric Michael and Mannich‐type reactions of glycine Schiff bases with higher enantioselectivity than previous catalysts. In the Mannich‐type reaction, aromatic N‐Boc‐protected imines (Boc=tert‐butoxycarbonyl) as well as enolizable alkyl imines were applicable. As a synthetic application of the catalytic asymmetric Mannich‐type reaction with the optimized TaDiASs, we developed a catalytic asymmetric total synthesis of (+)‐nemonapride, which is an antipsychotic agent.  相似文献   

3.
A polystyrene‐cross‐linking tricyclohexylphosphine (PS‐TCP) was synthesized through radical emulsion polymerization of 4‐tert‐butylstyrene as a monomer and tris(trans‐4‐styrylcyclohexyl)phosphine as a threefold cross‐linker. The PS‐TCP showed enhanced ligand performance compared to the corresponding polystyrene‐triphenylphosphine hybrid PS‐TPP and tricyclohexylphosphine in Pd‐catalyzed Suzuki–Miyaura and Buchwald–Hartwig reactions of aryl chlorides.  相似文献   

4.
Chloroacylation reactions of cross‐linked polystyrene (CPS) microspheres were conducted with two kinds of ω‐chloroacyl chlorides, chloroacetyl chloride, and 4‐chlorobutyryl chloride as reagent, respectively, and the chloromethylation reaction of CPS microspheres was also performed using 1,4‐bis (chloromethoxy) butane as reagent, obtaining three kinds of modified CPS microspheres on which the exchangeable chlorine atoms were introduced. Subsequently, the reactions of these modified microspheres with triphenylphosphine were carried out, respectively, and three kinds of quaternary phosphonium salt (QPS)‐type triphase catalysts (TPC) were prepared. These catalysts were used in the N‐alkylation reaction of phthalimide, namely the reaction of phthalimide in water phase with 1‐bromobutane in organic phase, resulting in N‐butylphthalimide. The effects of main reaction conditions on the triphase‐transfer catalysis reaction were examined, and the relationship between the structure and catalytic activity for these TPC was investigated. The experimental results indicate that the prepared QPS‐type TPC are effective for the N‐alkylation reaction of phthalimide carried out between oil phase and water phase. The polarity of the organic solvent and the temperature affect the reaction rate greatly. The result of the reaction in nitrobenzene having the highest polarity among the used several solvents is the best. It will make the reaction to speed up to raise temperature. The chemical structures of the TPC have crucial influences on the catalytic activity of the TPC. The catalyst with a longer spacer arm, which links the catalytic group to the matrix microspheres, has higher activity. The bonding density of QPS group on the polymer microspheres affects the hydrophilic and hydrophobic property of the TPC and, accordingly, affects the catalytic activity greatly. There is a maximum conversion of 1‐bromobutane as the bonding density of QPS group on TPC is 0.94 mmol/g. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 677–686, 2011  相似文献   

5.
A Resin‐linker‐vector (RLV) strategy is described for the radiosynthesis of tracer molecules containing the radionuclide 18F, which releases the labelled vector into solution upon nucleophilic substitution of a polystyrene‐bound arylsulfonate linker with [18F]‐fluoride ion. Three model linker‐vector molecules 7 a – c containing different alkyl spacer groups were assembled in solution from (4‐chlorosulfonylphenyl)alkanoate esters, exploiting a lipase‐catalysed chemoselective carboxylic ester hydrolysis in the presence of the sulfonate ester as a key step. The linker‐vector systems were attached to aminomethyl polystyrene resin through amide bond formation to give RLVs 8 a – c with acetate, butyrate and hexanoate spacers, which were characterised by using magic‐angle spinning (MAS) NMR spectroscopy. On fluoridolysis, the RLVs 8 a , b containing the longer spacers were shown to be more effective in the release of the fluorinated model vector (4‐fluorobutyl)phenylcarbamic acid tert‐butyl ester ( 9 ) in NMR kinetic studies and gave superior radiochemical yields (RCY≈60 %) of the 18F‐labelled vector. The approach was applied to the synthesis of the radiopharmaceutical O‐(2‐[18F]‐fluoroethyl)‐L ‐tyrosine ([18F]‐FET), delivering protected [18F]‐FET in >90 % RCY. Acid deprotection gave [18F]‐FET in an overall RCY of 41 % from the RLV.  相似文献   

6.
The palladium‐catalyzed asymmetric [4+3] cyclization of trimethylenemethane donors with benzofuran‐derived azadienes furnishes chiral benzofuro[3,2‐b]azepine frameworks in high yields (up to 98 %) with exclusive regioselectivities and excellent stereoselectivities (up to >20:1 d.r., >99 % ee). This catalytic asymmetric [4+3] cyclization of Pd‐trimethylenemethane can enrich the arsenal of Pd‐TMM reactions in organic synthesis. In addition, this strategy provides an alternative approach to chiral azepines by a transition‐metal‐catalyzed asymmetric [4+3] cyclization.  相似文献   

7.
The mechanism of a typical Petasis‐type boronic mannich reaction (the styrylboronic acid, dibenzylamine, and α‐hydroxylpropionaldehyde) has been investigated using density functional theory calculations. According to our calculations, the reaction is most likely to proceed through the following steps: 1) the nucleophilic addition of the amine to the aldehyde to form the carbinolamine; 2) the dehydration of the carbinolamine; 3) the formation of the tetra‐coordinated borate intermediate; 4) the C? C bond formation by the intramolecular transfer of the styryl group; 5) the hydrolysis of the resulting intermediate to give the final products. The highest point on the energy profile is the transition state for the C? C bond formation (118.8 kJ·mol?1 above the reactants in ethanol). Our results can give reasonable explanations on some experimental facts observed for many Petasis‐type boronic Mannich reactions.  相似文献   

8.
Aryl ketones represent problematic substrates for asymmetric Mannich reactions due to a large steric hindrance exhibited by such compound species. A highly enantioselective direct Mannich reaction of aryl methyl ketones with cyclic imine benzo[e][1,2,3]oxathiazine 2,2‐dioxides could be successfully carried out utilizing a combination of cinchona alkaloid‐derived primary amines with trifluoroacetic acid (TFA); the primary amines feature a superior catalytic efficacy over secondary amines with a variety of sterically hindered carbonyl compounds as substrates. The reaction proceeded well with various cyclic imines in 89–97 % ee and with various aryl methyl ketones in 85–98 % ee. Moreover, the aryl carbonyl of a Mannich product could be transformed to ketoxime, which further undergoes a Beckmann rearrangement to produce an amide compound while maintaining enantioselectivity.  相似文献   

9.
In this study, dithiomalonates (DTMs) were demonstrated to be exceptionally efficient Mannich donors in terms of reactivity and stereoselectivity in cinchona‐based‐squaramide‐catalyzed enantioselective Mannich reactions of diverse imines or α‐amidosulfones as imine surrogates. Owing to the superior reactivity of DTMs as compared to conventional malonates, the catalyst loading could be reduced to 0.1 mol % without the erosion of enantioselectivity (up to 99 % ee). Furthermore, by the use of a DTM, even some highly challenging primary alkyl α‐amidosulfones were smoothly converted into the desired adducts with excellent enantioselectivity (up to 97 % ee), whereas the use of a malonate or monothiomalonate resulted in no reaction under identical conditions. The synthetic utility of the chiral Mannich adducts obtained from primary alkyl substrates was highlighted by the organocatalytic, coupling‐reagent‐free synthesis of the antidiabetic drug (?)‐(R)‐sitagliptin.  相似文献   

10.
[BMIM]Cl catalyzed three component Mannich‐type reaction of 4‐(quinazolin‐4‐yloxy)benzenamine and aldehyde with dialkyl phosphite under microwave irradiation has been described. The salient features of the reaction leading to new α‐aminophosphonates include shorter reaction time and good yields. The method is environmentally friendly and does not require toxic catalysts or solvents. To the best of our knowledge, this is the first report for [BMIM]Cl induced one‐pot synthesis of α‐aminophosphonate derivatives.  相似文献   

11.
The one‐pot, three‐component Mannich reactions of thiopyran‐4‐one 1 with different aromatic aldehydes and aniline derivatives in the presence of catalytic quantities of ZrOCl2.8H2O (15 mol %) led to rapid and high yield formation of various 3‐methylamino substituted derivatives of 1 at room temperature. Spectroscopic and X‐ray analyses of the products suggested the formation of the anti stereoisomers as the major product of the reactions.  相似文献   

12.
《化学:亚洲杂志》2017,12(9):963-967
A highly efficient direct asymmetric three‐component Mannich reaction of an N ‐acylpyrazole, an aldehyde and a primary or secondary amine, was enabled by a rhodium‐based Lewis‐acid catalyst with metal‐centered chirality. Excellent enantioselectivities were achieved for a variety of substrates at a typical catalyst loading of merely 0.5 mol % (23 examples, up to 98 % ee ).  相似文献   

13.
Allylation, trans‐ and cis‐crotylation of (2E,4E)‐2,4‐dimethylhexadienal, a representative α,β,γ,δ‐unsaturated aldehyde, was carried out under different catalytic and stoichiometric conditions. The reactions catalyzed by organocatalysts TRIP‐PA and N,N′‐dioxides gave the best results with respect to yields, asymmetric induction, and catalyst load in comparison to other procedures. The developed methodology was applied in the enantioselective synthesis of (5R,6S)‐(+)‐pteroenone, a defensive metabolite (ichthyodeterrent) of the Antarctic pteropod Clione antarctica.  相似文献   

14.
将(2S,5S)-吡咯烷-2,5-二羧酸用于催化羟基丙酮与苯胺和苯甲醛的直接不对称Mannich反应,以较好的收率(64~95%)和最高96% ee得到了syn-1,2-氨基醇化合物。  相似文献   

15.
A twofold interpenetrated pillared–bilayer framework, {[Zn3( L )2( L2 )(DMF)] ? (18DMF)(6H2O)}n ( 1 ), has been synthesized from the ligands tris(4′‐carboxybiphenyl)amine ( H3L ) and 1,2‐bis(4‐pyridyl)ethylene ( L2 ). The structure contains [Zn3(COO)6] secondary building units (SBUs), in which three ZnII ions are almost linear with carboxylate bridging. This framework undergoes reversible pillar linker substitution reactions at the terminal ZnII centers with three different dipyridyl linkers of different lengths to afford three daughter frameworks, 2 – 4 . Frameworks 2 – 4 are interconvertible through reversible linker substitution reactions. Also, competitive linker‐exchange experiments show preferential incorporation of linker L3 in the parent framework 1 . The larger linker L5 does not undergo such substitution reactions and framework 5 , which contains this linker, can be synthesized solvothermally as a twofold interpenetrated structure. Interestingly, when framework 5 is dipped in a solution of L3 in DMF, linker substitution takes place as before, but linker L5 now moves and diagonally binds two ZnII centers to afford 6 as a nonpenetrated single framework. This linker place exchange reaction is unprecedented. All of these reactions take place in a single‐crystal to single‐crystal (SC‐SC) manner, and have been observed directly through X‐ray crystallography. In addition, each 3D framework undergoes complete copper(II) transmetalation.  相似文献   

16.
Two types of polymer‐supported nanometal catalysts with high catalytic activity and recyclability in water have been developed. One catalyst was composed of linear polystyrene‐stabilized metal nanoparticles (PS‐MtNPs). A palladium catalyst (PS‐PdONPs) was prepared in water by the thermal decomposition of Pd(OAc)2 in the presence of polystyrene. The degree of immobilization of Pd, but not the size of the Pd nanoparticles, was dependent on the molecular weight and cross‐linking of the polystyrene. The PS‐PdONPs exhibited high catalytic activity for Suzuki, Heck, and Sonogashira coupling reactions in water and they could be recycled without loss of activity. Linear polystyrene was also suitable as a stabilizer for in situ generated PdNPs and PtNPs. The second catalyst was a polyion complex that was composed of poly[4‐chloromethylstyrene‐co‐(4‐vinylbenzyl)tributylammonium chloride] and poly(acrylic acid)‐stabilized PdNPs (PIC‐PdNPs). Aggregation and redispersion of PIC‐PdNPs were easily controlled by adjusting the pH value of the solution.  相似文献   

17.
An aldehyde that is not fluorescent responsive toward a chiral diamine has been converted to a sensitive fluorescence enhancement sensor through incorporation of an additional hydrogen bonding unit to increase the structural rigidity of the reaction product of the aldehyde with the diamine. This new chiral aldehyde is synthesized in one step from the reaction of (S)‐3‐formylBINOL with salicyl chloride. When treated with trans‐1,2‐cyclohexanediamine in ethanol, it shows greatly enhanced fluorescence at λ=410 nm with good enantioselectivity. NMR and mass spectroscopic methods are used to investigate the reaction of the chiral aldehyde with the diamine. This study has revealed a two‐stage reaction mechanism including a fast imine formation and a slow ester cleavage.  相似文献   

18.
The catalytic asymmetric [3+2] cycloaddition of α‐ketoesters with 2‐nitrovinylindoles and 2‐nitrovinyl‐ pyrroles has been established. This strategy allowed the construction of structurally diverse pyrrolo[1,2‐a]indoles bearing three contiguous stereocenters in generally high yields and good to excellent stereoselectivities (up to 98% yield, > 98 : 2 dr, 99% ee). The efficient synthesis of tetracyclic psychotropic compound analogue via the derivatization of cycloadduct showed the great synthetic potential of this strategy.  相似文献   

19.
Highly efficient and diastereodivergent aza‐Diels–Alder reactions have been developed to access either diastereomeric series of benzofuran‐fused δ‐lactams and dihydropyridines in nearly perfect stereoselectivity (d.r. >20:1, >99 % ee for all examples). The complementarity of N‐heterocyclic carbene and chiral amine as the catalyst was demonstrated for the first time, together with an excellent level of catalytic efficiency (1 mol % loading).  相似文献   

20.
The photochemistry of two isostructural metal–organic frameworks based on 5‐amino/5‐formamidoisophthalate (CAU‐10‐NH2/NHCHO) or mixed‐linker 5‐amino/5‐formamido‐ and 5‐nitroisophthalate (CAU‐10‐NO2/NH2/NHCHO) has been studied using laser flash photolysis. 355 nm excitation of CAU‐10‐NH2/NHCHO leads to a transient absorption spectrum characterized by a broad continuous absorption from 380 to 800 nm that was attributed to the presence of holes (440 nm) and electrons (600 nm) based on iPrOH and N2O quenching, respectively. In contrast, no transients were observed for the isostructural mixed‐linker CAU‐10‐NO2/NH2/NHCHO, data that is compatible with the uniform distribution of linkers 5‐amino/5‐formamido/5‐nitroisophthalate as charge‐transfer complex pairs. The same effect of quenching of 5‐aminoisophthalate transients by 5‐nitroisophthalate was also observed in aqueous solution (pH 9) but with much lower strength. Using a simple Stern–Volmer formalism allowed the estimation of the interaction of 5‐aminoisophthalate with 5‐nitroisophthalate in MOF to be 5.2×104 times stronger than in the aqueous phase.  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