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.
《Comptes Rendus Chimie》2002,5(2):111-118
The sequential treatment of trans-(C6F5)(p-tol3P)2PtC≡CC≡CC≡CSiEt3 (2, acetone solution) with n-Bu4N+ F in wet THF (to generate a PtC≡CC≡CC≡CH complex), ClSiMe3 (fluoride ion scavenger), excess HC≡CSiEt3, excess O2, and CuCl/TMEDA (0.20–0.25 equiv in acetone; Hay cross-coupling conditions) gives trans-(C6F5)(p-tol3P)2PtC≡CC≡CC≡CC≡CSiEt3 (4, 30%), as previously reported. When an analogous reaction is conducted with SiMe4 in place of ClSiMe3, the side-product trans-(C6F5)(p-tol3P)2PtC≡CC≡CC≡CC(Me)2OSiEt3 is obtained (ca 28%), the structure of which is established crystallographically. For comparison, crystal structures of three related complexes, 4, 2, and trans-(p-tol)(Ph3P)2PtC≡CC≡CC≡CSiEt3, are determined.  相似文献   

2.
The platinum polyynyl complexes trans-(C6F5)(p-tol3P)2Pt(C≡C)n/2H undergo oxidative homocoupling (O2, CuCl/TMEDA) to diplatinum polyynediyl complexes trans, trans-(C6F5)(p-tol3P)2Pt(C≡C)nPt(Pp-tol3)2(C6F5) (n=4, 2 ; 6, 5 ; 8, 8 ; 92–97 %) as reported previously. When related reactions are conducted in the presence of CuI adducts of the 1,10-phenanthroline-based macrocycles 2,9-(1,10-phenanthrolinediyl)(p-C6H4O(CH2)6O)2(1,3-C6H4) ( 10 , 33-membered) or 2,9-(1,10-phenanthrolinediyl)(p-C6H4O(CH2)6O)2(2,7-naphthalenediyl) ( 11 , 35-membered), excess K2CO3, and I2 (oxidant), rotaxanes are isolated that feature a Pt(C≡C)nPt axle that has been threaded through the macrocycle ( 2⋅10 , 9 %; 5⋅10 , 41 %; 5⋅11 , 28 %; 8⋅10 , 12 %; 8⋅11 , 9 %). Their crystal structures are determined and analyzed in detail, particularly with respect to geometric perturbations and the degree of steric sp carbon chain insulation. NMR spectra show a number of shielding effects. UV/Vis spectra do not indicate significant electronic interactions between the Pt(C≡C)nPt axles and macrocycles, although cyclic voltammetry data suggest rapid reactions following oxidation.  相似文献   

3.
[(BDI)Mg+][B(C6F5)4] ( 1 ; BDI=CH[C(CH3)NDipp]2; Dipp=2,6-diisopropylphenyl) was prepared by reaction of (BDI)MgnPr with [Ph3C+][B(C6F5)4]. Addition of 3-hexyne gave [(BDI)Mg+ ⋅ (EtC≡CEt)][B(C6F5)4]. Single-crystal X-ray analysis, NMR investigations, Raman spectra, and DFT calculations indicate a significant Mg-alkyne interaction. Addition of the terminal alkynes PhC≡CH or Me3SiC≡CH led to alkyne deprotonation by the BDI ligand to give [(BDI-H)Mg+(C≡CPh)]2 ⋅ 2 [B(C6F5)4] ( 2 , 70 %) and [(BDI-H)Mg+(C≡CSiMe3)]2 ⋅ 2 [B(C6F5)4] ( 3 , 63 %). Addition of internal alkynes PhC≡CPh or PhC≡CMe led to [4+2] cycloadditions with the BDI ligand to give {Mg+C(Ph)=C(Ph)C[C(Me)=NDipp]2}2 ⋅ 2 [B(C6F5)4] ( 4 , 53 %) and {Mg+C(Ph)=C(Me)C[C(Me)=NDipp]2}2 ⋅ 2 [B(C6F5)4] ( 5 , 73 %), in which the Mg center is N,N,C-chelated. The (BDI)Mg+ cation can be viewed as an intramolecular frustrated Lewis pair (FLP) with a Lewis acidic site (Mg) and a Lewis (or Brønsted) basic site (BDI). Reaction of [(BDI)Mg+][B(C6F5)4] ( 1 ) with a range of phosphines varying in bulk and donor strength generated [(BDI)Mg+ ⋅ PPh3][B(C6F5)4] ( 6 ), [(BDI)Mg+ ⋅ PCy3][B(C6F5)4] ( 7 ), and [(BDI)Mg+ ⋅ PtBu3][B(C6F5)4] ( 8 ). The bulkier phosphine PMes3 (Mes=mesityl) did not show any interaction. Combinations of [(BDI)Mg+][B(C6F5)4] and phosphines did not result in addition to the triple bond in 3-hexyne, but during the screening process it was discovered that the cationic magnesium complex catalyzes the hydrophosphination of PhC≡CH with HPPh2, for which an FLP-type mechanism is tentatively proposed.  相似文献   

4.
The synthesis of a unique series of heteromultinuclear transition metal compounds is reported. Complexes 1‐I‐3‐Br‐5‐(FcC≡C)‐C6H3 ( 4 ), 1‐Br‐3‐(bpy‐C≡C)‐5‐(FcC≡C)‐C6H3 ( 6 ), 1,3‐(bpy‐C≡C)2‐5‐(FcC≡C)‐C6H3 ( 7 ), 1‐(XC≡C)‐3‐(bpy‐C≡C)‐5‐(FcC≡C)‐C6H3 ( 8 , X = SiMe3; 9 , X = H), 1‐(HC≡C)‐3‐[(CO)3ClRe(bpy‐C≡C)]‐5‐(FcC≡C)‐C6H3 ( 11 ), 1‐[(Ph3P)AuC≡C]‐3‐[(CO)3ClRe(bpy‐C≡C)]‐5‐(FcC≡C)‐C6H3 ( 13 ), 1‐[(Ph3P)AuC≡C]‐3‐(bpy‐C≡C)‐5‐(FcC≡C)‐C6H3 ( 14 ), [1‐[(Ph3PAuC≡C]‐3‐[{[Ti](C≡CSiMe3)2}Cu(bpy‐C≡C)]‐5‐(FcC≡C)‐C6H3]PF6 ( 16 ), and [1,3‐[(tBu2bpy)2Ru(bpy‐C≡C)]2‐5‐(FcC≡C)‐C6H3](PF6)4 ( 18 ) (Fc = (η5‐C5H4)(η5‐C5H5)Fe, bpy = 2,2′‐bipyridiyl‐5‐yl, [Ti] = (η5‐C5H4SiMe3)2Ti) were prepared by using consecutive synthesis methodologies including metathesis, desilylation, dehydrohalogenation, and carbon–carbon cross‐coupling reactions. In these complexes the corresponding metal atoms are connected by carbon‐rich bridging units comprising 1,3‐diethynyl‐, 1,3,5‐triethynylbenzene and bipyridyl units. They were characterized by elemental analysis, IR and NMR spectroscopy, and partly by ESI‐TOF mass spectrometry., The structures of 4 and 11 in the solid state are reported. Both molecules are characterized by the central benzene core bridging the individual transition metal complex fragments. The corresponding acetylide entities are, as typical, found in a linear arrangement with representative M–C, C–CC≡C and C≡C bond lengths.  相似文献   

5.
Terminal alkyne coupling reactions promoted by rhodium(I) complexes of macrocyclic NHC-based pincer ligands—which feature dodecamethylene, tetradecamethylene or hexadecamethylene wingtip linkers viz. [Rh(CNC-n)(C2H4)][BArF4] (n=12, 14, 16; ArF=3,5-(CF3)2C6H3)—have been investigated, using the bulky alkynes HC≡CtBu and HC≡CAr’ (Ar’=3,5-tBu2C6H3) as substrates. These stoichiometric reactions proceed with formation of rhodium(III) alkynyl alkenyl derivatives and produce rhodium(I) complexes of conjugated 1,3-enynes by C−C bond reductive elimination through the annulus of the ancillary ligand. The intermediates are formed with orthogonal regioselectivity, with E-alkenyl complexes derived from HC≡CtBu and gem-alkenyl complexes derived from HC≡CAr’, and the reductive elimination step is appreciably affected by the ring size of the macrocycle. For the homocoupling of HC≡CtBu, E-tBuC≡CCH=CHtBu is produced via direct reductive elimination from the corresponding rhodium(III) alkynyl E-alkenyl derivatives with increasing efficacy as the ring is expanded. In contrast, direct reductive elimination of Ar'C≡CC(=CH2)Ar’ is encumbered relative to head-to-head coupling of HC≡CAr’ and it is only with the largest macrocyclic ligand studied that the two processes are competitive. These results showcase how macrocyclic ligands can be used to interrogate the mechanism and tune the outcome of terminal alkyne coupling reactions, and are discussed with reference to catalytic reactions mediated by the acyclic homologue [Rh(CNC-Me)(C2H4)][BArF4] and solvent effects.  相似文献   

6.
Treatment of the mono(salicylaldiminato)titanium complexes {3-But-2-(O)C6H3CHN(Ar)}TiCl3(THF) (Ar = C6H5, 2,4,6-Me3C6H2 or C6F5) with the potassium β-enaminoketonates (C6H5)NC(CH3)C(H)C(R)OK (R = CH3, CF3) yielded the first examples of heteroligated (salicylaldiminato) (β-enaminoketonato)titanium dichloride complexes. The complex {3-But-2-(O)C6H3CHN(C6H5)}{(C6H5)NC(CH3)C(H)C(CH3)O}TiCl2 was structurally characterized by X-ray diffraction and has an orientation with trans-O,O,cis-Cl,Cl, cis-N,N distorted octahedral geometry. These complexes polymerize ethene when activated with MAO; the highest productivity, 5650 kg PE (mol metal)−1 h−1 atm−1, was afforded by {3-But-2-(O)C6H3CHN(C6F5)}{(C6H5)NC(CH3)C(H)C(CF3)O}TiCl2 at 60 °C.  相似文献   

7.
Transparent orange crystals of [Yb(MeCp)2(O2CC6F5)]2 and [Yb(MeCp)2(O2C‐o‐HC6F4)]2 were obtained by oxidation of Yb(MeCp)2 with M(O2CR) (M = 1/2 Hg, Tl; R = C6F5, o‐HC6F4) in tetrahydrofuran. They have a dimeric structure with bridging bidentate (O, O')‐benzoate groups and eight coordinated ytterbium. Both crystallise isotypic in the orthorhombic space group Pbca. Room temperature as well as low temperature single crystal X‐ray investigations show the o‐H/F positions in [Yb(MeCp)2(O2C‐o‐HC6F4)]2 not to be ordered.  相似文献   

8.
The crucial factor of the reaction of 2,6-di-tert-butylphenol with alkali hydroxides is temperature, depending on which two types of potassium or sodium 2,6-di-tert-butylphenoxides are formed. These types exhibit different catalytic activity in the alkylation of 2,6-di-tert-butylphenol with methyl acrylate. More active forms of 2,6-But 2C6H3OK or 2,6-But 2C6H3ONa are synthesized at temperatures higher than 160 °C and are predominantly the monomers, which dimerize on cooling. The data of 1H NMR, electronic, and IR spectra for the corresponding forms of 2,6-But 2C6H3OK and 2,6-But 2C6H3ONa isolated in the individual state are in agreement with cyclohexadienone structure. In DMSO or DMF, the dimeric forms of 2,6-di-tert-butylphenoxides react with methyl acrylate to form methyl 3-(4-hydroxy-3,5-di-tert-butylphenyl)propionate in 64–92% yield. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2138–2143, December, 2006.  相似文献   

9.
Tetrakis(p‐tolyl)oxalamidinato‐bis[acetylacetonatopalladium(II)] ([Pd2(acac)2(oxam)]) reacted with Li–C≡C–C6H5 in THF with formation of [Pd(C≡C–C6H5)4Li2(thf)4] ( 1a ). Reaction of [Pd2(acac)2(oxam)] with a mixture of 6 equiv. Li–C≡C–C6H5 and 2 equiv. LiCH3 resulted in the formation of [Pd(CH3)(C≡C–C6H5)3Li2(thf)4] ( 2 ), and the dimeric complex [Pd2(CH3)4(C≡C–C6H5)4Li4(thf)6] ( 3 ) was isolated upon reaction of [Pd2(acac)2(oxam)] with a mixture of 4 equiv. Li–C≡C–C6H5 and 4 equiv. LiCH3. 1 – 3 are extremely reactive compounds, which were isolated as white needles in good yields (60–90%). They were fully characterized by IR, 1H‐, 13C‐, 7Li‐NMR spectroscopy, and by X‐ray crystallography of single crystals. In these compounds Li ions are bonded to the two carbon atoms of the alkinyl ligand. 1a reacted with Pd(PPh3)4 in the presence of oxygen to form the already known complexes trans‐[Pd(C≡C–C6H5)2(PPh3)2] and [Pd(η2‐O2)(PPh3)2]. In addition, 1a is an active catalyst for the Heck coupling reaction, but less active in the catalytic Sonogashira reaction.  相似文献   

10.
Sequential treatment of 2‐C6H4Br(CHO) with LiC≡CR1 (R1=SiMe3, tBu), nBuLi, CuBr?SMe2 and HC≡CCHClR2 [R2=Ph, 4‐CF3Ph, 3‐CNPh, 4‐(MeO2C)Ph] at ?50 °C leads to formation of an intermediate carbanion (Z)‐1,2‐C6H4{CA(=O)C≡CBR1}{CH=CH(CH?)R2} ( 4 ). Low temperatures (?50 °C) favour attack at CB leading to kinetic formation of 6,8‐bicycles containing non‐classical C‐carbanion enolates ( 5 ). Higher temperatures (?10 °C to ambient) and electron‐deficient R2 favour retro σ‐bond C?C cleavage regenerating 4 , which subsequently closes on CA providing 6,6‐bicyclic alkoxides ( 6 ). Computational modelling (CBS‐QB3) indicated that both pathways are viable and of similar energies. Reaction of 6 with H+ gave 1,2‐dihydronaphthalen‐1‐ols, or under dehydrating conditions, 2‐aryl‐1‐alkynylnaphthlenes. Enolates 5 react in situ with: H2O, D2O, I2, allylbromide, S2Me2, CO2 and lead to the expected C ‐E derivatives (E=H, D, I, allyl, SMe, CO2H) in 49–64 % yield directly from intermediate 5 . The parents (E=H; R1=SiMe3, tBu; R2=Ph) are versatile starting materials for NaBH4 and Grignard C=O additions, desilylation (when R1=SiMe) and oxime formation. The latter allows formation of 6,9‐bicyclics via Beckmann rearrangement. The 6,8‐ring iodides are suitable Suzuki precursors for Pd‐catalysed C?C coupling (81–87 %), whereas the carboxylic acids readily form amides under T3P® conditions (71–95 %).  相似文献   

11.
In the comproportionation reaction of CuIIX2 and Cu0 with isopropylacetylene (iPr−C≡C−H), the ethynediide species C22− is generated via concomitant C−H/C−C bond cleavage of the iPr−C≡C−H precursor under moderate temperature to direct the formation of CuI mixed ethynediide/isopropylethynide nanoclusters (potentially explosive). The active ethynediide dianion C22− exhibits chameleon‐like templating behavior to form C2@Cum (m =6 ( 3 , 4 ), 7 ( 2 , 4 ), 8 ( 1 )) central structural units for successive formation of {C22−⊂Cu24} ( 1 , 2 ), {6 C22−⊂Cu48} ( 3 ), and {18 C22−⊂Cu92} ( 4 ) complexes. Bearing the highest C22− content, complex 4 features an unprecedented nanoscale Cu2C2 kernel. Furthermore, 1 – 3 exhibit structure‐controlled photoluminescence in the solid state.  相似文献   

12.
Enhanced reactivity is shown by uncoordinated C≡C bonds in the proximity of a metal in phosphanylacetylene complexes. cis-[Pt(C6F5)2(thf)2] reacts with [M(C6F5)2(PPh2C≡CPh)2] (M=Pt, Pd) to form binuclear complexes containing the novel 2,3-bis(diphenylphosphanyl)-1,3-butadien-1-yl bridging ligand. Substitution of the solvent ligands with, for example, PPh2H (see picture) provides species that could be characterized by X-ray crystallography.  相似文献   

13.
The reaction of alkynyldifluoroboranes RC≡CBF2 (R = (CH3)3C, CF3, (CF3)2CF) with organyliodine difluoride R′IF2 bearing electron‐withdrawing polyfluoroorganyl groups R′ = C6F5, (CF3)2CFCF=CF, C4F9, and CF3CH2 leads to the corresponding alkynyl(organyl)iodonium salts [(RC≡C)(R′)I][BF4]. This approach uses a widely applicable method as demonstrated for a representative series of polyfluorinated aryl‐, alkenyl‐, and alkyliodine difluorides. Generally, these syntheses proceed with good yields and deliver pure iodonium salts. The distinct electrophilic nature of their [(RC≡C)(R′)I]+ cations is deduced from multinuclear magnetic resonance data. Within the series of new iodonium salts [CF3C≡C(C4F9)I][BF4] is an intrinsic unstable one and decomposed forming CF3C≡CI and C4F10.  相似文献   

14.
Reactions of [Ru{C=C(H)-1,4-C6H4C≡CH}(PPh3)2Cp]BF4 ([ 1 a ]BF4) with hydrohalic acids, HX, results in the formation of [Ru{C≡C-1,4-C6H4-C(X)=CH2}(PPh3)2Cp] [X=Cl ( 2 a-Cl ), Br ( 2 a-Br )], arising from facile Markovnikov addition of halide anions to the putative quinoidal cumulene cation [Ru(=C=C=C6H4=C=CH2)(PPh3)2Cp]+. Similarly, [M{C=C(H)-1,4-C6H4-C≡CH}(LL)Cp ]BF4 [M(LL)Cp’=Ru(PPh3)2Cp ([ 1 a ]BF4); Ru(dppe)Cp* ([ 1 b ]BF4); Fe(dppe)Cp ([ 1 c ]BF4); Fe(dppe)Cp* ([ 1 d ]BF4)] react with H+/H2O to give the acyl-functionalised phenylacetylide complexes [M{C≡C-1,4-C6H4-C(=O)CH3}(LL)Cp’] ( 3 a – d ) after workup. The Markovnikov addition of the nucleophile to the remote alkyne in the cations [ 1 a–d ]+ is difficult to rationalise from the vinylidene form of the precursor and is much more satisfactorily explained from initial isomerisation to the quinoidal cumulene complexes [M(=C=C=C6H4=C=CH2)(LL)Cp’]+ prior to attack at the more exposed, remote quaternary carbon. Thus, whilst representative acetylide complexes [Ru(C≡C-1,4-C6H4-C≡CH)(PPh3)2Cp] ( 4 a ) and [Ru(C≡C-1,4-C6H4-C≡CH)(dppe)Cp*] ( 4 b ) reacted with the relatively small electrophiles [CN]+ and [C7H7]+ at the β-carbon to give the expected vinylidene complexes, the bulky trityl ([CPh3]+) electrophile reacted with [M(C≡C-1,4-C6H4-C≡CH)(LL)Cp’] [M(LL)Cp’=Ru(PPh3)2Cp ( 4 a ); Ru(dppe)Cp* ( 4 b ); Fe(dppe)Cp ( 4 c ); Fe(dppe)Cp* ( 4 d )] at the more exposed remote end of the carbon-rich ligand to give the putative quinoidal cumulene complexes [M{C=C=C6H4=C=C(H)CPh3}(LL)Cp’]+, which were isolated as the water adducts [M{C≡C-1,4-C6H4-C(=O)CH2CPh3}(LL)Cp’] ( 6 a–d ). Evincing the scope of the formation of such extended cumulenes from ethynyl-substituted arylvinylene precursors, the rather reactive half-sandwich (5-ethynyl-2-thienyl)vinylidene complexes [M{C=C(H)-2,5-cC4H2S-C≡CH}(LL)Cp’]BF4 ([ 7 a – d ]BF4 add water readily to give [M{C≡C-2,5-cC4H2S-C(=O)CH3}(LL)Cp’] ( 8 a – d )].  相似文献   

15.
The most common method for the deprotection of TBDMS ethers utilizes stoichiometric amounts of tetrabutylammonium fluoride, n-Bu4N+F (TBAF), which is highly corrosive and toxic. We have developed a mild and chemoselective method for the deprotection of TBDMS, TES, and TIPS ethers using iron(III) tosylate as a catalyst. Phenolic TBDMS ethers, TBDPS ethers and the BOC group are not affected under these conditions. Iron(III) tosylate is an inexpensive, commercially available, and non-corrosive reagent.  相似文献   

16.
The dialkyl malonate derived 1,3-diphosphines R2C(CH2PPh2)2 (R= a , Me; b , Et; c , n-Bu; d , n-Dec; e , Bn; f , p-tolCH2) are combined with (p-tol3P)2PtCl2 or trans-(p-tol3P)2Pt((C≡C)2H)2 to give the chelates cis-(R2C(CH2PPh2)2)PtCl2 ( 2 a – f , 94–69 %) or cis-(R2C(CH2PPh2)2)Pt((C≡C)2H)2 ( 3 a – f , 97–54 %). Complexes 3 a – d are also available from 2 a – d and excess 1,3-butadiyne in the presence of CuI (cat.) and excess HNEt2 (87–65 %). Under similar conditions, 2 and 3 react to give the title compounds [(R2C(CH2PPh2)2)[Pt(C≡C)2]4 ( 4 a – f ; 89–14 % (64 % avg)), from which ammonium salts such as the co-product [H2NEt2]+ Cl are challenging to remove. Crystal structures of 4 a , b show skew rhombus as opposed to square Pt4 geometries. The NMR and IR properties of 4 a – f are similar to those of mono- or diplatinum model compounds. However, cyclic voltammetry gives only irreversible oxidations. As compared to mono-platinum or Pt(C≡C)2Pt species, the UV-visible spectra show much more intense and red-shifted bands. Time dependent DFT calculations define the transitions and principal orbitals involved. Electrostatic potential surface maps reveal strongly negative Pt4C16 cores that likely facilitate ammonium cation binding. Analogous electronic properties of Pt3C12 and Pt5C20 homologs and selected equilibria are explored computationally.  相似文献   

17.
A new series of mono- and bis-alkynyl CoIII(TIM) complexes (TIM=2,3,9,10-tetramethyl-1,4,8,11-tetraazacyclotetradeca-1,3,8,10-tetraene) is reported herein. The trans-[Co(TIM)(C2R)Cl]+ complexes were prepared from the reaction between trans-[Co(TIM)Cl2]PF6 and HC2R (R=tri(isopropyl)silyl or TIPS ( 1 ), -C6H4-4-tBu ( 2 ), -C6H4-4-NO2 ( 3 a ), and N-mesityl-1,8-naphthalimide or NAPMes ( 4 a )) in the presence of Et3N. The intermediate complexes of the type trans-[Co(TIM)(C2R)(NCMe)](PF6)(OTf), 3 b and 4 b , were obtained by treating 3 a and 4 a , respectively, with AgOTf in CH3CN. Furthermore, bis-alkynyl trans-[Co(TIM)(C2R)2]PF6 complexes, 3 c and 4 c , were generated following a second dehydrohalogenation reaction between 3 b and 4 b , respectively, and the appropriate HC2R in the presence of Et3N. These new complexes have been characterized using X-ray diffraction ( 2 , 3 a , 4 a , and 4 c ), IR, 1H NMR, UV/Vis spectroscopy, fluorescent spectroscopy ( 4 c ), and cyclic voltammetry.  相似文献   

18.
The dimesitylpropargylphosphanes mes2P?CH2?C≡C?R 6 a (R=H), 6 b (R=CH3), 6 c (R=SiMe3) and the allene mes2P?C(CH3)=C=CH2 ( 8 ) were reacted with Piers’ borane, HB(C6F5)2. Compound 6 a gave mes2PCH2CH=CH(B(C6F5)2] ( 9 a ). In contrast, addition of HB(C6F5)2 to 6 b and 6 c gave mixtures of 9 b (R=CH3) and 9 c (R=SiMe3) with the regioisomers mes2P?CH2?C[B(C6F5)2]=CRH 2 b (R=CH3) and 2 c (R=SiMe3), respectively. Compounds 2 b , c underwent rapid phosphane/borane (P/B) frustrated Lewis pair (FLP) reactions under mild conditions. Compound 2 c reacted with nitric oxide (NO) to give the persistent FLP NO radical 11 . The systems 2 b , c cleaved dihydrogen at room temperature to give the respective phosphonium/hydridoborate products 13 b , c . Compound 13 c transferred the H+/H? pair to a small series of enamines. Compound 13 c was also a metal‐free catalyst (5 mol %) for the hydrogenation of the enamines. The allene 8 reacted with B(C6F5)3 to give the zwitterionic phosphonium/borate 17 . The ‐PPh2‐substituted mes2P‐propargyl system 6 d underwent a typical 1,2‐P/B‐addition reaction to the C≡C triple bond to form the phosphetium/borate zwitterion 20 . Several products were characterized by X‐ray diffraction.  相似文献   

19.
1H NMR method showed that in systems based on triisobutylaluminum (TIBA) and triphenylcyclopropenylium [Ph3C3]+[B(C6F5)4]–(CPB) or triphenylmethylium [Ph3C]+[B(C6F5)4]–(TB) perfluorophenylborates in a toluene–dichloromethane mixture the Friedel–Crafts process occurs with the formation of ditolylmethane (DTM) accompanied by the complete decomposition of TIBA to form isobutane. 19F NMR spectroscopy showed that the [B(C6F5)4]–anion decomposes in the systems to form B(C6F5)3 and HC6F5. The short-living [AlBu2 i]+ cation formed in the reaction of perfluorophenylborates with TIBA is assumed to be the species initiating the process. It has been shown that CPB is less reactive than TB. The addition of a stoichiometric amount of Ph2CCpFluHfMe2 exerts no effect on the process with the CPB-containing system but inhibits the reaction in the case of TB.  相似文献   

20.
A novel β-diketiminate stabilized gallium hydride, (DippL)Ga(Ad)H (where ( Dipp L)={HC(MeCDippN)2}, Dipp=2,6-diisopropylphenyl and Ad=1-adamantyl), has been synthesized and shown to undergo insertion of carbon dioxide into the Ga−H bond under mild conditions. In this case, treatment of the resulting κ1-formate complex with triethylsilane does not lead to regeneration of the hydride precursor. However, when combined with B(C6F5)3, (DippL)Ga(Ad)H catalyses the reductive hydrosilylation of CO2. Under stoichiometric conditions, the addition of one equivalent of B(C6F5)3 to (DippL)Ga(Ad)H leads to the formation of a 3-coordinate cationic gallane complex, partnered with a hydroborate anion, [(DippL)Ga(Ad)][HB(C6F5)3]. This complex rapidly hydrometallates carbon dioxide and catalyses the selective reduction of CO2 to the formaldehyde oxidation level at 60 °C in the presence of Et3SiH (yielding H2C(OSiEt3)2). When catalysis is undertaken in the presence of excess B(C6F5)3, appreciable enhancement of activity is observed, with a corresponding reduction in selectivity: the product distribution includes H2C(OSiEt3)2, CH4 and O(SiEt3)2. While this system represents proof-of-concept in CO2 hydrosilylation by a gallium hydride system, the TOF values obtained are relatively modest (max. 10 h−1). This is attributed to the strength of binding of the formatoborate anion to the gallium centre in the catalytic intermediate (DippL)Ga(Ad){OC(H)OB(C6F5)3}, and the correspondingly slow rate of the turnover-limiting hydrosilylation step. In turn, this strength of binding can be related to the relatively high Lewis acidity measured for the [(DippL)Ga(Ad)]+ cation (AN=69.8).  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