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 187 毫秒
1.
The first catalytic asymmetric synthesis of highly functionalized difluoromethylated cyclopropanes is described. The method, based on a rhodium‐catalyzed cyclopropanation of difluoromethylated olefins, gives access to a broad range of cyclopropanes bearing ester, ketone, or nitro functional groups. By using Rh2((S )‐BTPCP)4 as a catalyst, the corresponding products were obtained in high yields and high diastereo‐ and enantioselectivities (up 20:1 d.r. and 99 % ee ). This methodology allowed preparation of enantioenriched difluoromethylcyclopropanes for the first time.  相似文献   

2.
The catalytic asymmetric synthesis of highly functionalized cyclopropanes from α-substituted allyl sulfones and silanes is reported. The reaction, using α-aryl diazoacetates or diacceptor diazo reagents, catalyzed by a chiral rhodium complex (Rh2((S)-BTPCP)4), furnished the corresponding cyclopropanes in moderate to high yields (27–97 %), high diastereoselectivities (68 : 32 to 20 : 1 d.r.) and moderate to excellent ee (40–99 %). This methodology offers a privileged access to an underexplored class of enantioenriched cyclopropanes with a high level of functionality, an asset for further post-functionalization and their incorporation into more complex structure.  相似文献   

3.
The first successful example of a catalytic asymmetric cyclopropanation with α-diazopropionates is described. The cyclopropanation reaction of 1-aryl-substituted and related conjugated alkenes with tert-butyl α-diazopropionate has been achieved by catalysis with dirhodium(II) tetrakis[N-tetrabromophthaloyl-(S)-tert-leucinate], Rh2(S-TBPTTL)4, providing the corresponding cyclopropane products containing a quarternary stereogenic center in good to high yields and with high diastereo- and enantioselectivities (trans:cis = 90:10 to >99:1, 81-93% ee).  相似文献   

4.
The heteroleptic dirhodium paddlewheel catalyst 7 with a chiral carboxylate/acetamidate ligand sphere is uniquely effective in asymmetric [2+1] cycloadditions with α-diazo-α-trimethylstannyl (silyl, germyl) acetate. Originally discovered as a trace impurity in a sample of the homoleptic parent complex [Rh2((R)-TPCP)4] ( 5 ), it is shown that the protic acetamidate ligand is quintessential for rendering 7 highly enantioselective. The -NH group is thought to lock the ensuing metal carbene in place via interligand hydrogen bonding. The resulting stannylated cyclopropanes undergo “stereoretentive” cross coupling, which shows for the first time that even chiral quarternary carbon centers can be made by the Stille–Migita reaction.  相似文献   

5.
The heteroleptic dirhodium paddlewheel catalyst 7 with a chiral carboxylate/acetamidate ligand sphere is uniquely effective in asymmetric [2+1] cycloadditions with α‐diazo‐α‐trimethylstannyl (silyl, germyl) acetate. Originally discovered as a trace impurity in a sample of the homoleptic parent complex [Rh2((R)‐TPCP)4] ( 5 ), it is shown that the protic acetamidate ligand is quintessential for rendering 7 highly enantioselective. The ‐NH group is thought to lock the ensuing metal carbene in place via interligand hydrogen bonding. The resulting stannylated cyclopropanes undergo “stereoretentive” cross coupling, which shows for the first time that even chiral quarternary carbon centers can be made by the Stille–Migita reaction.  相似文献   

6.
The first catalytic asymmetric hetero-Diels–Alder (HDA) reaction between 3-tert-butyldimethylsilyloxy-1-dimethylamino-1,3-pentadiene (4-methyl-substituted Rawal’s diene) and aldehydes is described. With 3 mol % of dirhodium(II) tetrakis[N-benzene-fused-phthaloyl-(S)-piperidinonate], Rh2((S)-BPTPI)4, the cycloaddition reaction proceeded exclusively in an endo fashion and gave, after a novel sequential treatment with dimethyl acetylenedicarboxylate and acetyl chloride, the corresponding 2,3-cis-disubstituted dihydropyranones with up to 98% ee and perfect diastereoselectivity. The utility of this catalytic protocol was demonstrated by an asymmetric synthesis of the (−)-cis-aerangis lactone.  相似文献   

7.
Summary : Cluster-containing monomers were obtained and characterized. Mono- and disubstituted products were obtained under mild conditions via the interaction of Rh6(CO)16 with 4-vinylpyridine (4-VPy). Substitution of labile acetonitrile ligand in Rh6(CO)15NCMe by allyldiphenylphosphine (AlPPh2) yields Rh6(CO)14(µ,η2-PPh2CH2CHCH2) with formation of π-complex. The copolymerization of cluster-containing monomers synthesized with traditional monomers has been studied. It was found that Rh6- containing monomers change neither the ligand surroundings nor the structure of cluster monomer framework during polymerization reaction. Polymer-immobilized clusters were found to be active in hydrogenation reactions of cyclohexene.  相似文献   

8.
The reaction of Rh4(CO)12 with Pd(PBu t 3)2 yielded the high nuclearity bimetallic hexarhodium-tripalladium cluster complex Rh6(CO)16[Pd(PBu t 3)]3, 10, in 11% yield. Compound 10 was converted to the hexarhodium-tetrapalladium cluster Rh6(CO)16[Pd(PBu t 3)]4, 11, in 62% yield by reaction with an additional quantity of Pd(PBu t 3)2. Both compounds were characterized crystallographically. Structurally, both compounds consist of an octahedral cluster of six rhodium atoms with sixteen carbonyl ligands analogous to that of the known compound Rh6(CO)16. Compound 10 also contains three Pd(PBu t 3) groups that bridge three Rh–Rh bonds along edges of the Rh6 octahedron to give an overall D3 symmetry to the Rh6Pd3 cluster. Compound 11 contains four edge bridging Pd(PBu t 3) groups distributed across the Rh6 octahedron to give an overall D2d symmetry to the Rh6Pd4 cluster. Each Rh–Pd connection in both compounds contains a bridging carbonyl ligand that helps to stabilize the bond between the Pd(PBu t 3) groups and the Rh atoms. Both compounds can be regarded as Pd(PBu t 3) adducts of Rh6(CO)16.  相似文献   

9.
Parallel effort : Stereodivergent parallel kinetic resolution of a racemic mixture of dienes using Davies' [Rh2{(S)‐dosp}4] or [Rh2{(R)‐dosp}4] catalysts promotes a tandem vinyl diazoacetate cyclopropanation/Cope rearrangement sequence to afford two diastereomeric, enantioenriched cycloheptadienes, which correspond to the natural antipodes of the title diterpenoids (see scheme).

  相似文献   


10.
6‐(Diazomethyl)‐1,3‐bis(methoxymethyl)uracil ( 5 ) was prepared from the known aldehyde 3 by hydrazone formation and oxidation. Thermolysis of 5 and deprotection gave the pyrazolo[4,3‐d]pyrimidine‐5,7‐diones 7a and 7b . Rh2(OAc)4 catalyzed the transformation of 5 into to a 2 : 1 (Z)/(E) mixture of 1,2‐diuracilylethenes 9 (67%). Heating (Z)‐ 9 in 12n HCl at 95° led to electrocyclisation, oxidation, and deprotection to afford 73% of the pyrimido[5,4‐f]quinazolinetetraone 12 . The Rh2(OAc)4‐catalyzed reaction of 5 with 3,4‐dihydro‐2H‐pyran and 2,3‐dihydrofuran gave endo/exo‐mixtures of the 2‐oxabicyclo[4.1.0]heptane 13 (78%) and the 2‐oxabicyclo[3.1.0]hexane 15 (86%), Their treatment with AlCl3 or Me2AlCl promoted a vinylcyclopropane–cyclopentene rearrangement, leading to the pyrano‐ and furanocyclopenta[1,2‐d]pyrimidinediones 14 (88%) and 16 (51%), respectively. Similarly, the addition product of 5 to 2‐methoxypropene was transformed into the 5‐methylcyclopenta‐pyrimidinedione 18 (55%). The Rh2(OAc)4‐catalyzed reaction of 5 with thiophene gave the exo‐configured 2‐thiabicyclo[3.1.0]hexane 19 (69%). The analoguous reaction with furan led to 8‐oxabicyclo[3.2.1]oct‐2‐ene 20 (73%), and the reaction with (E)‐2‐styrylfuran yielded a diastereoisomeric mixture of hepta‐1,4,6‐trien‐3‐ones 21 (75%) that was transformed into the (1E,4E,6E)‐configured hepta‐1,4,6‐trien‐3‐one 21 (60%) at ambient temperature.  相似文献   

11.
Transformation of N‐alkylated anilines to N‐aryloxamates was studied using ethyl 2‐diazoacetoacetate as an alkylating agent and dirhodium tetraacetate (Rh2(OAc)4) as the catalyst. The general applicability of the reaction as a synthetic method for N‐aryloxamates was studied with a number of substituted N‐alkylated anilines. The results revealed that the oxamate was formed by a radical reaction with molecular O2 and Rh2(OAc)4 as initiator.  相似文献   

12.
The dirhodium complex bis­(benzonitrile)tetra­kis[μ‐4‐(diethyl­amino)benzoato‐κ2O:O′]dirhodium(II)(RhRh) benzonitrile disolvate, [Rh2(C11H14NO2)4(C7H5N)2]·2C7H5N, lies about an inversion centre. The dirhodium complex (methanol)tetra­kis(μ‐4‐nitro­benzoato‐κ2O:O′)(pyridine)dirhodium(II)(RhRh) dichloro­methane solvate, [Rh2(C7H4NO4)4(C5H5N)(CH4O)]·CH2Cl2, lies in a general position in the unit cell, but the complexes dimerize around an inversion centre via O—H⋯O hydrogen bonding of the axial MeOH to a carboxyl­ate O atom. In the latter crystal structure, π–π stacking inter­actions between the bridging 4‐nitro­benzoate ligands and the axial pyridine ligand are observed between adjacent mol­ecules.  相似文献   

13.
The synthesis and X-ray structure of the binuclear complex tetrakis[(4S)-4-phenyloxazolidin-2-one]-dirhodium(II) ([Rh2{(4S)-phox}4]) are reported. Structure-selectivity comparisons are made for typical metal carbene transformations, such as inter- and intramolecular cyclopropane formation, intermolecular cyclopropene formation and intramolecular C–H insertions of diazoacetates and diazoacetamides. The enantioselectivity achieved in the [Rh2{(4S)-phox}4]-catalyzed reactions is intermediate between that of [Rh2{(5S)-mepy}4] and [Rh2{(4R)-bnox}4], which were described previously (mepy = methyl 5-oxopyrrolidine-2-carboxylate; bnox = 4-benzyloxazolidin-2-one). In contrast to other catalyzed intermolecular cyclopropane formations, those using [Rh2{(4S)-phox}4] result preferentially in formation of the cis-cyclopropane.  相似文献   

14.
Various ligands, such as (Z)‐1‐phenyl‐2‐[(4S)‐4‐phenyl‐4,5‐dihydro‐1,3‐oxazol‐2‐yl]ethen‐1‐ol ((S)‐ 1a ) and (Z)‐1‐phenyl‐2‐[(4S)‐4‐phenyl‐4,5‐dihydro‐1,3‐thiazol‐2‐yl]ethen‐1‐ol ((S)‐ 1c ), were investigated as auxiliaries for the asymmetric synthesis of chiral ruthenium(II) complexes. The reaction of these chiral auxiliary ligands with [RuCl2(dmso)4], 2,2′‐bipyridine (bpy, 2.2 equiv), and triethylamine (10 equiv) in DMF/PhCl (1:8) at 140 °C for several hours diastereoselectively provided the complexes Λ‐[Ru(bpy)2{(S)‐ 1a ? H}] (Λ‐(S)‐ 2a , 52 % yield, 56:1 d.r.) and Λ‐[Ru(bpy)2{(S)‐ 1c ? H}] (Λ‐(S)‐ 2c , 48 % yield, >100:1 d.r.) in a single step after purification. Both Λ‐(S)‐ 2a and Λ‐(S)‐ 2c could be converted into Λ‐[Ru(bpy)3](PF6)2 by replacing the bidentate enolato ligands with bpy, under retention of configuration, induced by either NH4PF6 as a weak acid (from Λ‐(S)‐ 2a : 73 % yield, 22:1 e.r.; from Λ‐(S)‐ 2c : 77 % yield, 22:1 e.r.), TFA as a strong acid (from Λ‐(S)‐ 2a : 72 % yield, 52:1 e.r.; from Λ‐(S)‐ 2c : 85 % yield, 25:1 e.r.), methylation with Meerwein′s salt (from Λ‐(S)‐ 2a : 59 % yield, 46:1 e.r.; from Λ‐(S)‐ 2c : 86 % yield, 37:1 e.r.), ozonolysis (from Λ‐(S)‐ 2a : 56 % yield, 22:1 e.r.; from Λ‐(S)‐ 2c : 43 % yield, 6.3:1 e.r.), or oxidation with a peroxy acid (from Λ‐(S)‐ 2a : 72 % yield, 45:1 e.r.; from Λ‐(S)‐ 2c : 79 % yield, 8.5:1 e.r.). This study shows that, except for the reaction with NH4PF6, oxazoline‐enolato complex Λ‐(S)‐ 2a provides Λ‐[Ru(bpy)3](PF6)2 with higher enantioselectivities than analogous thiazoline‐enolato complex Λ‐(S)‐ 2c , which might be due to the higher coordinative stability of the thiazoline‐enolato complex, thus requiring more prolonged reaction times. Thus, this study provides attractive new avenues for the asymmetric synthesis of non‐racemic ruthenium(II)‐polypyridyl complexes without the need for using a strong acid or a strong methylating reagent, as has been the case in all previously reported auxiliary methods from our group.  相似文献   

15.
Chenli Fan  Yin Zuo 《合成通讯》2013,43(21):2782-2792
Abstract

A convenient and efficient procedure for the synthesis of 2-substituted-6,7-dihydrobenzo[d]oxazol-4(5H)-ones and 2-aryl-6,7-dihydrobenzofuran-4(5H)-ones through a Rh2(OAc)4-catalyzed C≡X (X?=?N, C) insertion of cyclic 2-diazo-1,3-diketones with nitriles and aromatic alkynes has been developed. This reaction uses readily available starting materials and stable cyclic 2-diazo-1,3-diketone compounds, with desired products formed in good to high yields. A tentative mechanism involving a C≡X bond insertion and 1,5-dipolar electrocyclization/ring opening and cyclization sequence for this reaction is proposed.  相似文献   

16.
Rh17S15‐Rh catalysts supported on acid‐treated carbon black were prepared from RhCl3 by a facile method using sulfur source ((NH4)2S2O3) and reducing agent (NaBH4), followed by an additional thermal treatment at 650 °C in argon. The prepared catalyst comprised an Rh17S15 single crystalline phase and a zero‐valent metal (Rh) phase supported on a conductive carbon. By XRD characterization, the constituent ratio of Rh17S15 to Rh in the electrocatalysts, ranging from 51–95 %, varied with the increase of amount of (NH4)2S2O3 or NaBH4. Morphologies of the resulting catalysts were characterized by transmission electron microscopy (TEM) technique. Most of particles were found to have a distribution of agglomerates ranging in size from 10 to 50 nm. In studying the effect of the constituent phases of chalcogenide electrocatalysts on oxygen reduction reaction activity, it is paramount to understand and optimize the structure sensitivity of the reaction, which will aid in determining the optimal ratio of Rh17S15 to Rh of the electrocatalyst. Activity and stability of the prepared catalysts were addressed using a series of cyclic voltammetry (CV) experiments in 1 M HCl electrolyte, in which the electrocatalyst of 95 % of Rh17S15 was found to be the most stable. The rotating disk electrode (RDE) experiment indicated the sulfide catalyst with 82 % of Rh17S15 showed the better performance for the ORR, which was discussed based on the compromise between the stability for the constituent phase of Rh17S15 in 1 M HCl and enhanced activity found for Rh phase.  相似文献   

17.
Eight dinuclear rhodium(II) complexes containing various, (primarily, polyfunctional) N-donor ligands in the trans position with respect to the Rh-Rh bond were synthesized and characterized by X-ray diffraction. In the Chinese-lantern dinuclear rhodium(II) pivalates, RhII 2 (μ-OOCCMe3)4(L)2 (L is 2,3-diaminopyridine (2), 7,8-benzoquinoline (4), 2,2′:6′,2″-terpyridine (5), N-phenyl-o-phenylenediamine (7)), and RhII 2 (μ-OOCCMe3)4L1L2 (3, L1 is 2-phenylpyridine, L2 = MeCN), the steric effects of the axial ligands are most strongly reflected in the Rh-N(L) and Rh-Rh bond lengths. The introduction of chelating ligands containing a conformationally rigid chelate ring leads to the cleavage of two carboxylate bridges to form the dinuclear double-bridged structure RhII 2 (μ- OOCCMe3)2(OCCMe3)22-L3)2, where L3 is 8-amino-2,4-dimethylquinoline (6). The reaction of complex 7 containing coordinated N-phenyl-o-phenylenediamine with pyrrole-2,5-dialdehyde afforded the new RhII 2(μ-OOCCMe3)4(L4)2 complex (8) containing 5-(1-phenyl-1-H-benzimidazol-2-yl)-1H-pyrrole-2-carbaldehyde (L4) in the axial positions of the dirhodium tetracarboxylate fragment. The coordinated diamine differs in reactivity from the free diamine. The reaction of the former with the above dialdehyde affords the [1+1]-condensation product, viz., 5-{(E)-[(2-anilinophenyl)imino]methyl}-1-H-pyrrole-2-carbaldehyde, whereas the reaction of unsubstituted o-phenylenediamine gives 5-{(E)-[(2-aminophenyl)imino]methyl}-1-H-pyrrole-2-carbaldehyde (L5) . The reaction of the latter with RhII 2(μ-OOCCMe3)4(H2O)2 affords the dinuclear complex RhII 2(μ-OOCCMe3)2(OOCCMe3)22-L5)2 (9), which is an analog of complex 6 containing only two bridging carboxylate groups.__________Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 581–591, March, 2005.  相似文献   

18.
The sodium salts of anionic chiral cobalt(III) complexes (CCC?Na+) have been found to be efficient catalysts of the asymmetric Povarov reaction of easily accessible dienophiles, such as 2,3‐dihydrofuran, ethyl vinyl ether, and an N‐protected 2,3‐dihydropyrrole, with 2‐azadienes. Ring‐fused tetrahydroquinolines with up to three contiguous stereogenic centers were thus obtained in high yields, excellent diastereoselectivities (endo/exo up to >20:1), and high enantioselectivities (up to 95:5 e.r.).  相似文献   

19.
Regio- and stereoselective distal allylic/benzylic C−H functionalization of allyl and benzyl silyl ethers was achieved using rhodium(II) carbenes derived from N-sulfonyltriazoles and aryldiazoacetates as carbene precursors. The bulky rhodium carbenes led to highly site-selective functionalization of less activated allylic and benzylic C−H bonds even in the presence of electronically preferred C−H bonds located α to oxygen. The dirhodium catalyst Rh2(S-NTTL)4 is the most effective chiral catalyst for triazole-derived carbene transformations, whereas Rh2(S-TPPTTL)4 works best for carbenes derived from aryldiazoacetates. The reactions afford a variety of δ-functionalized allyl silyl ethers with high diastereo- and enantioselectivity. The utility of the present method was demonstrated by its application to the synthesis of a 3,4-disubstituted l -proline scaffold.  相似文献   

20.
Preparation and Vibrational Spectra of Nonahalogenodirhodates(III), [Rh2ClnBr9-n]3?, n = 0–9 The pure nonahalogenodirhodates(III), A3[Rh2ClnBr9-n], A = K, Cs, (TBA); n = 0–4, 9, have been prepared. They are formed from the monomer chlorobromorhodates(III), [RhClnBr6-n]3?, n = 0–6, which are bridged to confacial bioctahedral complexes by ligand abstraction in less polar organic solvents. From the mixtures the complexions are separated by ion exchange chromatography on DEAE-cellulose. The solid, air-stable, air-stable, K-, Cs- and (TBA)-salts of [Rh2ClnBr9-n]3?, n = 0–4, are green, of [Rh2Cl9]3? are brown. The IR and Raman spectra of [Rh2Br9]3? and [Rh2Cl9]3? are assigned according to the point group D3h. The chlorobromodirhodates exist as mixtures of geometrical and structural isomers, which belong to different point groups. The vibrational spectra exhibit bands in characteristic regions; at high wavenumbers stretching vibrations with terminal ligands v(Rh—Clt): 360–320, v(Rh—Brt): 280–250; in a middle region with bridging ligands v(Rh—Clb): 300–270, v(Rh—Brb): 210–170 cm?1; the deformation bands are observed at distinct lower frequencies. The terminal ligands are fixed very strong, and the distance between v(Rh—Xt) and v(Rh—Xb) increases with decreasing size of the cations.  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