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 78 毫秒
1.
In recent years, C2-symmetric chiral bis (oxazoline) ligand-metal complexes have received a great of attention through their use in various catalytic processes[1-2]. Since the early 1990s, many impressive enantioselective carbon-carbon bond forming reactions, aziridination reactions, hydrosilylations, oxidations and reductions have been recorded using bis (oxazoline)-metal complexes with wide structural diversity[3-4]. The latest review is intended to focus on the recent developments of bis(oxazoline) ligand-metal catalyzed asymmetric reactions and their applications in organic synthesis. Herein we intend to report the synthesis of a new type of chiral bis(oxazoline), 1,3-bis(oxazolinyl)benzene. Application of their asymmetric catalysis is under working.  相似文献   

2.
Propylene copolymers with different polar groups were synthesised using rac‐Et[1‐Ind]2ZrCl2/MAO as the catalyst system. 10‐Undecen‐1‐ol, 10‐undecenoyl chloride, 10‐undecenoic acid, 2‐(9‐decen‐1‐yl)‐1,3‐oxazoline, 2‐(9‐decen‐1‐yl)‐4,4‐dimethyl‐1,3‐oxazoline, and 2‐[4‐(10‐undecene‐1‐oxy)phenyl]‐1,3‐oxazoline were used as comonomers. The addition of water to the 10‐undecenoyl chloride copolymer solution led to an acid‐functionalised copolymer. In the case of 2‐(9‐decen‐1‐yl)‐1,3‐oxazoline and its homopolymers, polymerisation temperature was varied. Up to 0.61 mol‐% comonomer were incorporated into the poly(propylene)s. The catalyst activities for 10‐undecen‐1‐ol, 10‐undecenoyl chloride and 10‐undecenoic acid were much higher than for the oxazoline comonomers.  相似文献   

3.
The enantioselective Lewis acid-catalyzed Diels-Alder reaction of 3-(2-propenoyl)-1,3-oxazolidin-2-one 8 with cyclopentadiene was examined using a series of chiral mox ligands 2-6, deferring in the side chain at 2-position of the chiral oxazoline and in the nature of the substituent at the chiral center (4-position) of the oxazoline ring, and a combination of N-[(1R)-2-chloro-1-phenylethyl]-2-[(4R)-4-phenyl-4,5-dihydrooxazol-2-yl]butyramide 2-MgI(2)-I(2) was the most efficient catalyst.  相似文献   

4.
A novel type of heterocycle, viz., 2,3a-disubstituted 5,6-dihydro-3aH-[1,3]oxazolo[3,2-b][1,2,4]oxadiazoles, was generated by an intermolecular PtII-mediated 1,3-dipolar cycloaddition (1,3-DCA) between the oxazoline N-oxide C(Me)2CH2OC(R)=N+(O-) (R = Me, Et) and coordinated nitriles in the complexes trans/cis-[PtCl2(R'CN)2] [R' = Me, Et, CH2Ph, Ph, N(C5H10)]. The reaction is unknown for free RCN and oxazoline N-oxides, but under PtII-mediated conditions, it proceeds smoothly (CH2Cl2, 20-25 degrees C, 18-20 h) and gives pure complexes [PtCl2{N=C(R')ONC(R)OCH2CMe2}2] [R/R' = Me/Me, 1; Me/Et, 2; Me/CH2Ph, 3; Me/Ph, 4; Me/N(C5H10), 5; Et/Me, 6; Et/Et, 7; Et/CH2Ph, 8; Et/Ph, 9; Et/N(C5H10), 10] in 42-84% yields after column chromatography. Compounds 1-10 were characterized by elemental analyses (C, H, N), FAB+-MS, IR, and 1H and 13C{1H} NMR spectroscopies, and X-ray diffraction (for 1, 2, 5, and 9). With the exception of benzonitrile complexes, 1,3-DCA of oxazoline N-oxides to the PtII-ligated nitriles occurred diastereoselectively and afforded mixtures of enantiomers. Depending on the substituents on nitriles, asymmetric atoms in both of the formed heterocyclic ligands have the same (SS/RR) or different (SR/RS) configurations. The heterocyclic ligands were liberated from 1-4 and 6-9 by treatment with excess ethane-1,2-diamine (en) in CH2Cl2 for 1 day at 20-25 degrees C (for R' = Me, Et, CH2Ph) and at 50 degrees C (for R' = Ph) to achieve the free organic species and the well-known [Pt(en)2](Cl)2; the products were separated, and 2,3a-disubstituted 5,6-dihydro-3aH-[1,3]oxazolo[3,2-b][1,2,4]oxadiazoles (11-18) were characterized by ESI+-MS and 1H and 13C{1H} NMR spectroscopies.  相似文献   

5.
2‐(4‐Carboxyphenyl)‐1,3‐oxazoline ( 2 a ), 2‐(3‐carboxyphenyl)‐1,3‐oxazoline ( 2 b ), and 2‐(6‐carboxynaphthyl‐2‐yl)‐1,3‐oxazoline ( 2 c ) were synthesized by reaction of monomethyl ester chlorides of aromatic dicarboxylic acids with 2‐chloroethylamine hydrochloride in the presence of triethylamine followed by cyclization with methanolic KOH. Thermal polymerization in bulk within a few minutes at 200–220°C resulted in new linear poly(ester amide)s 3 a – 3 c without significant side reactions. The polymerization occurred in the melt phase ( 2 b ) or in the solid state ( 2 a , 2 c ) and the resulting polymers are amorphous ( 3 b ) or semi‐crystalline ( 3 a , 3 c ). The polyaddition reactions were investigated by means of differential scanning calorimetry (DSC) and 1H NMR spectroscopy.  相似文献   

6.
Using two different zirconocene/MAO catalyst systems, propene was copolymerized with the comonomers 2‐(9‐decene‐1‐yl)‐1,3‐oxazoline and 2‐(4‐(10‐undecene‐1‐oxo)phenyl)‐1,3‐oxazoline, respectively. The catalysts used were rac‐Et[Ind]2ZrCl2 and rac‐Me2Si[2‐Me‐4, 5‐BenzInd]2ZrCl2. Up to 0.53 mol‐% oxazoline could be incorporated into polypropene. Oxazoline content, molecular weight, degree of isotacticity and melting behavior were dependent on the catalyst system, comonomer structure and comonomer concentration in the feed.  相似文献   

7.
Chiral bis(oxazoline) complexes of Cu(II), Zn(II) and Mg(II) have been immobilized on silica support via hydrogen-bonding interactions. Up to 93% ee is obtained in the Diels-Alder reaction between 3-((E)-2-butenoyl)-1,3-oxazolin-2-one and cyclopentadiene at room temperature with the heterogeneous bis(oxazoline) complexes, and the catalysts can be recycled without losing enantioselectivity. Experimental and theoretical studies show that the reversal of the absolute product configuration upon immobilization of the PhBOX-Cu(II) catalyst is triggered by the anion dissociation from Cu(II) onto the surface of the support.  相似文献   

8.
《Tetrahedron: Asymmetry》1998,9(23):4239-4247
Diastereoselective lithiation of (S)-2-ferrocenyl-4-(1-methylethyl)oxazoline, followed by addition of N2O4, gave (S)-2-[(pS)-2-nitroferrocenyl]-4-(1-methylethyl)oxazoline which was subsequently converted into derivatives of (pS)-2-aminoferrocenecarboxylic acid. The corresponding (pR)-derivatives were obtained through use of a removable TMS blocking group. The 2-nitroferrocenyloxazolines produced in this work underwent facile photo-decomplexation to give 2-nitrocyclopentadienyliden-1,3-oxazolidenes.  相似文献   

9.
Two classes of indole-phosphine oxazoline ligands have been prepared from readily available starting materials in good overall yields. These modular ligands include an indole skeleton with either a phosphine moiety or an oxazoline ring at the 2- or 3-position, respectively. The utility of these ligands was demonstrated in a catalytic asymmetric reaction: the palladium-catalyzed allylic alkylation of 1,3-diphenyl-2-propenyl acetate with dimethyl malonate was performed with enantioselectivities as high as 98%.  相似文献   

10.
Oxidation of polyfunctional threo-(1S,2S)-2-amino-1,3-propanediol derivatives with a 717 anion-exchange resin-supported bromine has been investigated. The result showed that oxidized products were in close relationship with the substituents at nitrogen in the starting materials. Its primary and secondary amine derivatives were oxidized in the presence of Na(2)HPO(4) to give essentially a substituted chiral oxazoline or C(3)-O acylated product in high yield, while oxidation of its N,N-dimethyl derivative mainly gave a chiral N-methyl oxidation-formylation product. This selective oxidation was first observed in 2-amino-1,3-propanediol chemistry.  相似文献   

11.
The influence of Hofmeister salts was investigated on the cloud point of three poly(2‐oxazoline)s, namely poly(2‐ethyl‐2‐oxazoline) [PEtOx], poly(2‐n‐propyl‐2‐oxazoline) [PnPropOx], and poly(2‐isopropyl‐2‐oxazoline) [PiPropOx]. In addition, a comb polymer based on oligo‐2‐ethyl‐2‐oxazoline side chains and a methacrylate backbone (POEtOxMA) was included in this investigation. It was found that the ionic response of the poly(2‐oxazoline)s strongly depends on their hydrophilicity. The comb polymer POEtOxMA revealed a strikingly similar response to the salts as linear PEtOx even though the cloud points of the polymers in water differ. This indicates that the architecture does not significantly influence the effect of the Hofmeister ions, even though there is a difference in the absolute cloud point.

  相似文献   


12.
《Tetrahedron: Asymmetry》2001,12(10):1475-1478
A chiral bis(oxazoline) was grafted on ArgoGel and used in the palladium-catalysed substitution of (±)-1,3-diphenyl-2-propenyl acetate with dimethyl malonate. The enantioselectivity was the same as that observed when the analogous monomeric catalyst was used (94–95% e.e.), despite the fact that the C2 symmetry of the ligand was affected when coupled to the polymer. The polymer-supported catalyst could be recycled several times after removal of precipitated Pd(0). The polymer-bound bis(oxazoline) was also applied in a zinc-catalysed Diels–Alder reaction but lower selectivity and reactivity than the monomer was observed.  相似文献   

13.
The cyclopropanation reaction of chromium Fischer carbene complexes with alkenyl oxazolines has been studied in both racemic and enantioselective fashions. The oxazolinyl group acts as both electron-acceptor substituent and chiral auxiliary. Achiral (4,4-dimethyloxazolin-2-yl)alkenes derived from trans-crotonic and trans-cinnamic acids 2a,b undergo the cyclopropanation reaction to give 4a-d,g with excellent diastereoselectivity (trans/cis ratio between 93:7 and >97:3), while those derived from acrylic and metacrylic acids 2c,d give the cyclopropanes 4e,f,h with much lower selectivity (trans/cis ratio between 68:32 and 83:17). The homogeneous catalytic hydrogenolysis of 4 leads in a selective manner to 5 or 6, depending on the nature of the R3 substituent. The removal of the oxazoline moiety is achieved by carboxybenzylation/hydrolysis and ester reduction, yielding monoprotected 1,4- and 1,3-diols 9 and 11, respectively. The alkenes derived from enantiopure (S)-valinol and (S)-tert-leucinol 3 led to cyclopropanes trans-12 with high relative and absolute stereocontrol. Using tert-leucinol as the auxiliary permits attaining total facial stereoselectivity (>98% ee). Reductive cleavage of the cyclopropane ring and removal of the auxiliary afford the enriched alcohols (3S,4S)-9 and (S)-11. The stereochemical outcome of the cyclopropanation reaction is rationalized by a trans approach of the s-cis conformer of the alkenyl oxazoline to the carbene complex involving the less hindered face of the oxazoline auxiliary and the re-face of the carbene complex.  相似文献   

14.
Optically active P-stereogenic phosphine/oxazoline bidentate ligands (1) were prepared and applied to palladium-catalyzed allylic substitution of 1,3-diphenyl-1-acetoxy-2-propene with dimethyl malonate. The absolute configuration of the allylation product was remarkably switched by changing a palladium/ligand ratio between 1/1 and 1/2.  相似文献   

15.
Crystal structures are reported for three isomeric compounds, namely 2‐(2‐hydroxy­phenyl)‐2‐oxazoline, (I), 2‐(3‐hydroxy­phenyl)‐2‐oxazoline, (II), and 2‐(4‐hydroxy­phenyl)‐2‐oxazoline, (III), all C9H9NO2 [systematic names: 2‐(4,5‐dihydro‐1,3‐oxazol‐2‐yl)phenol, (I), 3‐(4,5‐dihydro‐1,3‐oxazol‐2‐yl)phenol, (II), and 4‐(4,5‐dihydro‐1,3‐oxazol‐2‐yl)phenol, (III)]. In these compounds, the deviation from coplanarity of the oxazoline and benzene rings is dependent on the position of the hydroxy group on the benzene ring. The coplanar arrangement in (I) is stabilized by a strong intra­molecular O—H⋯N hydrogen bond. Surprisingly, the 2‐oxazoline ring in mol­ecule B of (II) adopts a 3T4 (C2TC3) conformation, while the 2‐oxazoline ring in mol­ecule A, as well as that in (I) and (III), is nearly planar, as expected. Tetra­mers of mol­ecules of (II) are formed and they are bound together via weak C—H⋯N hydrogen bonds. In (III), strong inter­molecular O—H⋯N hydrogen bonds and weak intra­molecular C—H⋯O hydrogen bonds lead to the formation of an infinite chain of mol­ecules perpendicular to the b direction. This paper also reports a theoretical investigation of hydrogen bonds, based on density functional theory (DFT) employing periodic boundary conditions.  相似文献   

16.
The origin of the reversal of the enantioselectivity in the palladium-catalyzed allylic alkylation of rac-1,3-diphenyl-2-propenyl acetate with dimethyl malonate anion using chiral dihydroxy bis(oxazoline) "BO" ligands derived from (1S,2S)-(+)-2-amino-1-phenyl-1,3-propanediol was investigated. To determine the structural effects of the dihydroxy BO ligand on this unique phenomenon, new homochiral dihydroxy BO ligands were prepared from L-threonine and L-serine and were assessed in the transformation. The results obtained with these novel BO ligands, compared with the one obtained by using the dihydroxy BO ligands derived from (1S,2S)-(+)-2-amino-1-phenyl-1,3-propanediol, reveal that the reversal in the enantioselectivity observed with the dihydroxy BO ligand depends on the structure of the ligand. The effect of different bases used to generate the dimethyl malonate anion was also examined. The results are discussed in terms of the interaction of one hydroxy group in the intermediate pi-allyl palladium complex with the dimethyl malonate anion.  相似文献   

17.
2‐Pyridylsulfone‐ and fluoroalkylated group‐activated olefins underwent highly efficient diastereo‐ and enantioselective 1,3‐dipolar cycloadditions across various aromatic and aliphatic nitrones in the presence of a chiral NiII/bis(oxazoline) catalyst. The process was tuned by 4 Å molecular sieves, chiral bis(oxazoline) ligands, reaction solvents, and temperature. A wide array of optically pure fluoroalkylated isoxazolidines were obtained, thus facilitating the asymmetric synthesis of an enantioenriched α‐trifluoromethylated γ‐amino alcohol in gram‐scale and a trifluoromethylated derivative of 1,3‐oxazinan‐2‐one with potential pharmaceutical interest. A stereochemical model, based on the absolute configuration of one adduct and some control experiments, was postulated to account for the observed endo‐ and enantioselectivity.  相似文献   

18.
A palladium‐catalyzed asymmetric dihydroxylation of 1,3‐dienes with catechols is developed using chiral pyridinebis(oxazoline) ligands. Various chiral 2‐substituted 1,4‐benzodioxanes could be synthesized with moderate to high yields and enantioselectivities from readily available starting materials. The reaction is proposed to proceed through a cascade Wacker‐type hydroxypalladation/asymmetric allylation process.  相似文献   

19.
Microwave reactions of 2-amino-2-methyl-1-propanol (2) or 2-aminoethanethiol hydrochloride (4) with readily available N-acylbenzotriazoles 1a-j in the presence of SOCl(2) produced 2-substituted 2-oxazolines 3a-j in 84-98% yields and 2-substituted thiazolines 5a-i in 85-97% yields, respectively. With use of this method chiral oxazoline 6, bisoxazoline 7, bisthiazoline 8, and 5,6-dihydro-4H-1,3-oxazines 9 or 10 have also been prepared in 82-96% yields. These results demonstrate a new application of N-acylbenzotriazoles in the preparation of oxazolines and thiazolines under mild conditions and short reaction times with microwave irradiation.  相似文献   

20.
《Tetrahedron: Asymmetry》2007,18(9):1094-1102
The synthesis of a novel series of aminophosphine–oxazoline and phosphoramidite–oxazoline is described. The efficacy of these aminophosphine–oxazoline ligands was investigated in the palladium catalysed asymmetric allylic alkylation of 1,3-diphenylprop-2-enyl acetate leading to a maximum of 38% ee at 64% conversion. Phosphoramidite–oxazoline ligands, however, gave ees of up to 87% at 71% conversion in the same reaction and also proved to be effective in the palladium catalysed asymmetric Suzuki coupling between 2-methylnaphthylboronic acid and 1-bromonaphthalene, leading to a maximum of 46% ee in 54% isolated yield at room temperature.  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