首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Remarkable effects of SiMe3 and SiEt3 para-substituents in the phenoxide-modified half-titanocenes, Cp*TiCl2(O-2,6-iPr2-4-R-C6H2) [R=SiMe3 ( 6 ), SiEt3 ( 7 )], toward the catalytic activities in ethylene copolymerizations with 2-methyl-1-pentene, 1-decene, 1-dodecene and with 9-decen-1-ol (DC-OH) have been demonstrated. The activities by 6 , 7 at 50 °C showed higher than those conducted at 25 °C in all cases in the presence of MAO cocatalyst. Efficient synthesis of high-molecular-weight (HMW) ethylene copolymers incorporating DC-OH (or 5-hexen-1-ol, HX-OH) has been attained in the copolymerization by 7 , which showed better DC-OH (HX-OH) incorporation at 50 °C to afford the HMW copolymers, poly(ethylene-co-DC-OH)s, with high activities (activity 1.21–3.81×105 kg-polymer mol−1-Ti h, Mn=6.55–10.0×104, DC-OH 2.3–3.6 mol %).  相似文献   

2.
Barium complexes ligated by bulky boryloxides [OBR2] (where R=CH(SiMe3)2, 2,4,6-iPr3-C6H2 or 2,4,6-(CF3)3-C6H2), siloxide [OSi(SiMe3)3], and/or phenoxide [O-2,6-Ph2-C6H3], have been prepared. A diversity of coordination patterns is observed in the solid state for both homoleptic and heteroleptic complexes, with coordination numbers ranging between 2 and 4. The identity of the bridging ligand in heteroleptic dimers [Ba(μ2-X1)(X2)]2 depends largely on the given pair of ligands X1 and X2. Experimentally, the propensity to fill the bridging position increases according to [OB{CH(SiMe3)2}2)]<[N(SiMe3)2]<[OSi(SiMe3)3]<[O(2,6-Ph2-C6H3)]<[OB(2,4,6-iPr3-C6H2)2]. This trend is the overall expression of 3 properties: steric constraints, electronic density and σ- and π-donating capability of the negatively charged atom, and ability to generate Ba ⋅ ⋅ ⋅ F, Ba ⋅ ⋅ ⋅ C(π) or Ba ⋅ ⋅ ⋅ H−C secondary interactions. The comparison of the structural motifs in the complexes [Ae{μ2-N(SiMe3)2}(OB{CH(SiMe3)2}2)]2 (Ae = Mg, Ca, Sr and Ba) suggest that these observations may be extended to all alkaline earths. DFT calculations highlight the largely prevailing ionic character of ligand-Ae bonding in all compounds. The ionic character of the Ae-ligand bond encourages bridging coordination, whereas the number of bridging ligands is controlled by steric factors. DFT computations also indicate that in [Ba(μ2-X1)(X2)]2 heteroleptic dimers, ligand predilection for bridging vs. terminal positions is dictated by the ability to establish secondary interactions between the metals and the ligands.  相似文献   

3.
Starting from the para‐phenylenediamine derivative HN(SiMe3)‐C6H4‐NH(SiMe3), a lithiation and subsequent borylation give [(MeO)2B]N(SiMe3)‐C6H4‐N(SiMe3)[B(OMe)2] ( 1 ), the hydridation of which yields Li2[(H3B)N(SiMe3)‐C6H4‐N(SiMe3)(BH3)] ( 2 ). Applying ZrCl4 upon 2 initiates a condensation to give the title compound [‐N(SiMe3)‐p‐C6H4‐N(SiMe3)‐BH‐]2, a hetero[3, 3]paracyclophane with two N‐B‐N chains that connect the para‐phenylene units. The product 3 crystallizes in the orthorhombic space group P212121.  相似文献   

4.
Organometallic Compounds of the Lanthanides. 155 [1] Synthesis and Characterization of New Lanthanocene Complexes containing Silylated Cyclopentadienyl Ligands The trichlorides of yttrium, samarium, and lutetium react with two equiv. of K[C5H4SiEt3] ( 1 ) to form the dimeric compounds [(η5‐C5H4SiEt3)2LnCl]2 (Ln = Y ( 2 ), Sm ( 3 ), Lu ( 4 )). These react with one equiv. of methyllithium to give the corresponding dimeric lanthanocenemethyl complexes [(η5‐C5H4SiEt3)2LnMe]2 (Ln = Y ( 5 ), Sm ( 6 ), Lu ( 7 )). The reaction between samarium trichloride and lutetium trichloride, respectively with two equiv. of K[1, 3‐C5H3(SiMe3)2] followed by one equiv. of methyllithium results in the formation of the monomeric methyl complexes [η5‐1, 3‐C5H3(SiMe3)2]2LnMe(THF) (Ln = Sm ( 8 ), Lu ( 9 )). The new compounds have been characterized by elemental analysis, mass spectrometry, 1H‐ and 13C{1H} NMR spectroscopy, as well as 1 — 7 by single crystal X‐ray structure analysis.  相似文献   

5.
Desactivation of Catalysts in the Polymerization of Acetylene by Bis(trimethylsilyl)acetylene Complexes of Titanocene or Zirconocene Unexpected inactive byproducts were observed in the catalytic polymerization of acetylene using metallocene alkyne complexes Cp2M(L)(η2-Me3SiC2SiMe3), 1 : M = Ti, without L; 2 : M = Zr, L = thf. The reaction of 1 was investigated in detail by NMR to give quantitatively at –20 °C the titanacyclopentadiene Cp2Ti–CH=CH–C(SiMe3)=C(SiMe3) ( 3 ). Around 0 °C 3 starts to rearrange to yield the dihydroindenyl complex 4 via coupling of one Cp-ligand with the titanacyclopentadiene. In the reaction of 2 under analogous conditions a zirconacyclopentadiene Cp2Zr–CH=CH–C(SiMe3)=C(SiMe3) ( 5 ) and the dimeric complex [Cp2Zr(C(SiMe3)=CH(SiMe3)]2[μ-σ(1,2)-C≡C] ( 6 ) were observed. Whereas 5 decomposes to a mixture of unidentified paramagnetic species, 6 was isolated and investigated by NMR spectroscopy and X-ray analysis. In the reaction of rac-(ebthi)Zr(η2-Me3SiC2SiMe3) (ebthi = ethylenbistetrahydroindenyl) with 2-ethynyl-pyridine the complex rac-(ebthi)ZrC(SiMe3)=CH(SiMe3)](σ-C≡CPy) 7 was obtained, which was investigated by an X-ray analysis.  相似文献   

6.
The bicyclic amido-substituted silicon(I) ring compound Si4{N(SiMe3)Mes}4 2 (Mes=Mesityl=2,4,6-Me3C6H2) features enhanced zwitterionic character and different reactivity from the analogous compound Si4{N(SiMe3)Dipp}4 1 (Dipp=2,6-iPr2C6H3) due to the smaller mesityl substituents. In a reaction with the N-heterocyclic carbene NHC (1,3,4,5-tetramethyl-imidazol-2-ylidene), we observe adduct formation to give Si4{N(SiMe3)Mes}4 ⋅ NHC ( 3 ). This adduct reacts further with the Lewis acid BH3 to yield the Lewis acid–base complex Si4{N(SiMe3)Mes}4 ⋅ NHC ⋅ BH3 ( 4 ). Coordination of AlBr3 to 2 leads to the adduct 5 . Calculated proton affinities and fluoride ion affinities reveal highly Lewis basic and very weak Lewis acidic character of the low-valent silicon atoms in 1 and 2 . This is confirmed by protonation of 1 and 2 with Brookharts acid yielding 6 and 7 . Reaction with diphenylacetylene only occurs at 111 °C with 2 in toluene and is accompanied by fragmentation of 2 to afford the silacyclopropene 8 and the trisilanorbornadiene species 9 .  相似文献   

7.
《中国化学会会志》2017,64(11):1340-1346
In this investigation, we describe substituent effect on the dipole moment, ionization potential, electron affinity, structure, frontier orbitals energy, in the trans‐Cl(OC)(H3P)3W(≡C‐para‐C6H4X) (X = H, F, SiH3, CN, NO2, SiMe3, CMe3, NH2, NMe2) complexes using MPW1PW91 quantum chemical calculations. The nature of chemical bond between the [Cl(OC)(H3P)3W] and [C‐para‐C6H4X]+ fragments was illustrated with energy decomposition analysis (EDA). Percentage composition in terms of the defined groups of frontier orbitals for these complexes was inspected to investigate the character in metal–ligand bonds. Quantum theory of atoms in molecules (QTAIM) was used for illustration of metal–ligand bonds in these complexes.  相似文献   

8.
The reactions of [M(NO)(CO)4(ClAlCl3)] (M=Mo, W) with (iPr2PCH2CH2)2NH, (PNHP) at 90 °C afforded [M(NO)(CO)(PNHP)Cl] complexes (M=Mo, 1a ; W, 1b ). The treatment of compound 1a with KOtBu as a base at room temperature yielded the alkoxide complex [Mo(NO)(CO)(PNHP)(OtBu)] ( 2a ). In contrast, with the amide base Na[N(SiMe3)2], the PNHP ligand moieties in compounds 1a and 1b could be deprotonated at room temperature, thereby inducing dehydrochlorination into amido complexes [M(NO)(CO)(PNP)] (M=Mo, 3a ; W, 3b ; PNP=(iPr2PCH2CH2)2N)). Compounds 3a and 3b have pseudo‐trigonal‐bipyramidal geometries, in which the amido nitrogen atom is in the equatorial plane. At room temperature, compounds 3a and 3b were capable of adding dihydrogen, with heterolytic splitting, thereby forming pairs of isomeric amine‐hydride complexes [Mo(NO)(CO)H(PNHP)] ( 4a(cis) and 4a(trans) ) and [W(NO)(CO)H(PNHP)] ( 4b(cis) and 4b(trans) ; cis and trans correspond to the position of the H and NO groups). H2 approaches the Mo/W?N bond in compounds 3a , 3b from either the CO‐ligand side or from the NO‐ligand side. Compounds 4a(cis) and 4a(trans) were only found to be stable under a H2 atmosphere and could not be isolated. At 140 °C and 60 bar H2, compounds 3a and 3b catalyzed the hydrogenation of imines, thereby showing maximum turnover frequencies (TOFs) of 2912 and 1120 h?1, respectively, for the hydrogenation of N‐(4 ‐ methoxybenzylidene)aniline. A Hammett plot for various para‐substituted imines revealed linear correlations with a negative slope of ?3.69 for para substitution on the benzylidene side and a positive slope of 0.68 for para substitution on the aniline side. Kinetics analysis revealed the initial rate of the hydrogenation reactions to be first order in c(cat.) and zeroth order in c(imine). Deuterium kinetic isotope effect (DKIE) experiments furnished a low kH/kD value (1.28), which supported a Noyori‐type metal–ligand bifunctional mechanism with H2 addition as the rate‐limiting step.  相似文献   

9.
Formation of Cyclic Silylphosphanes. Reaction of Li-Phosphides with R2SiCl2 (R? Me, Et, t-Bu) The reaction of Me2SiCl2 with Li-phosphides (mixture of LiPH2, Li2PH) leads to the formation of Me2Si(PH2)Cl 1 , Me2Si(PH2)2 2 , H2P? SiMe2? PH? SiMe2Cl 3 , (H2P? SiMe2)2PH 4 , (HP? SiMe2)3 6 , 5 , 7 , 8 , 9 , 10 , 40 . Excess of phosphides in Et2O – as well as excess of LiPH2 – favourably forms 10 . Li2PH (virtually free of Li3P and LiPH2) is obtained by reaction of LiPH2 · DME with LiBu; Li3P by reaction of PH3 with LiBu in toluene. Isomerization by Li/H migration determines the course of reaction of the PH-bearing compounds with Li-phosphides. With Me2SiCl2 Li3P mainly generates compound 10 . The reaction of the Li-phosphides with Et2SiCl2 mainly leads to (HP? SiEt2)3 18 and (HP? SiEt2)2 17 as well as to Et2Si(PH2)Cl 11 , Et2Si(PH2)2 12 , (ClEt2Si)2PH 13 , H2P? SiEt2? PH? SiEt2Cl 14 , (H2P? SiEt2)2PH 15 and 16 . In the reaction with LiPH2 · DME the same compounds are obtained and isomerization by Li/H migration (formation of PH3) already begins at ?70°C. In toluene ClEt2Si? P(SiEt2)2P? SiEt2Cl is additionally formed. Derivatives of 9, 10, 40 are not observed. The reaction of (t-Bu)2SiCl2 with LiPH2 leads to HP[Si(t-Bu)2]2PH 20 (yield 76%) and formation of PH3, the reaction with Li2PH to 20 (54%) besides HP[Si(t-Bu)2]2PLi 21 .  相似文献   

10.
Polymerization and polymer properties of diphenylacetylenes with bulky silyl groups (SiMe2i-Pr, SiMe2t-Bu, SiMe2Ph, SiEt3) at para or meta position were studied under comparison with those of the SiMe3 derivatives. The present monomers polymerized in good yields with TaCl5-cocatalysts to form high molecular-weight polymers (M w > 4 × 105). The polymer yields of para-substituted monomers were similar to that of the SiMe3 derivative, while those of meta substituted monomers were lower than that of m-SiMe3 derivative. Most of the polymers were totally soluble in common solvents such as toluene and CHCl3, although the polymers with p-SiMe2t-Bu and p-SiMe2Ph groups were partly insoluble in all solvents. These polymers resembled SiMe3-containing homologues in the UV-visible absorption and thermal stability. The oxygen permeability coefficients of these polymers were in the range of 10?9?10?8 cm3 (STP) cm/(cm2·s cm Hg)—lower than those of the corresponding SiMe3-containing polymers. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
The gas-phase acidities of the six dimethylphenol isomers were determined experimentally, by using the kinetic method, and theoretically, through quantum chemistry calculations. The experimental values, relative to the gas-phase acidity of phenol, are (in kJ mol−1): −1.76 ± 0.76 (2,3-Me2C6H3OH), 1.78 ± 0.29 (2,4-Me2C6H3OH), 0.83 ± 0.58 (2,5-Me2C6H3OH), −4.39 ± 0.89 (2,6-Me2C6H3OH), 5.38 ± 1.08 (3,4-Me2C6H3OH), and 1.88 ± 0.08 (3,5-Me2C6H3OH). This trend was discussed by considering the substituent effects on the thermodynamic stabilities both of the parent phenols and the corresponding phenoxide ions. The above acidity data, the literature values for 2-, 3-, and 4-methylphenol, and the substituent effects analysis allowed to develop a simple empirical method to estimate the acidity of any methyl-substituted phenol.  相似文献   

12.
The behavior of different anilines H2NC6H4R (R = o-Me, p-Me, o-, m- and p? i Pr, p-OMe, p-CO2Et) and 2,6-Me2C6H3NH2 towards trihalophosphoranes was studied. 2,6-Me2C6H3NH2 failed to form the diaminophosphonium salt [Ph2PNH(2,6-Me2C6H3)2]Br, and the aminophosphine oxide Ph2(2,6-Me2C6H3NH)PO was the only isolated product. Both o- and p-toluidine gave the corresponding diaminophosphonium salts; however in the case of o-toluidine, the yield was low and a mixture with the respective aminophosphine oxide was observed. Anilines containing methoxy and ethoxycarbonyl groups in para-position form the diaminophosphonium salts in reasonable yields.  相似文献   

13.
Element-Organic Amine/Imine Compounds, XXXI. - Cyclometallation with N-tert-Butyl-Phosphorus-Nitrogen Iridium Complexes The interaction of R1R2N–PNR3 ( 1 ) (R1  SiMe3, tBu, iC3H7; R2  R3  SiMe3, tBu) with [M(COD)(μ-Cl)]2 ( 2 ), M  Rh, Ir, affords the amino(imino)phosphane complexes 3 , whose PN bond adds methanol with formation of the diamidophosphite complexes 4 . Already below 0°C the iridium compounds of 4 undergo cyclometallation of a tBu methyl group (R2) with formation of the hydrido-iridium metallaheterocycles 5 . The structures of 4b and 5a are elucidated by X-ray analyses.  相似文献   

14.
The syntheses of several dialkyl complexes based on rare‐earth metal were described. Three β‐diimine compounds with varying N‐aryl substituents (HL1=(2‐CH3O(C6H4))N?C(CH3)CH?C(CH3)NH(2‐CH3O(C6H4)), HL2 = (2,4,6‐(CH3)3 (C6H2))N?C(CH3)CH?C(CH3)NH(2,4,6‐(CH3)3(C6H2)), HL3 = PhN?C(CH3)CH(CH3) NHPh) were treated with Ln(CH2SiMe3)3(THF)2 to give dialkyl complexes L1Ln (CH2SiMe3)2 (Ln = Y ( 1a ), Lu ( 1b ), Sc ( 1c )), L2Ln(CH2SiMe3)2(THF) (Ln = Y ( 2a ), Lu ( 2b )), and L3Lu(CH2SiMe3)2(THF) (3). All these complexes were applied to the copolymerization of cyclohexene oxide (CHO) and carbon dioxide as single‐component catalysts. Systematic investigation revealed that the central metal with larger radii and less steric bulkiness were beneficial for the copolymerization of CHO and CO2. Thus, methoxy‐modified β‐diiminato yttrium bis(alkyl) complex 1a , L1Y(CH2SiMe3)2, was identified as the optimal catalyst, which converted CHO and CO2 to polycarbonate with a TOF of 47.4 h?1 in 1,4‐dioxane under a 15 bar of CO2 atmosphere (Tp=130 °C), representing the highest catalytic activity achieved by rare‐earth metal catalyst. The resultant copolymer contained high carbonate linkages (>99%) with molar mass up to 1.9 × 104 as well as narrow molar mass distribution (Mw/Mn = 1.7). © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 6810–6818, 2008  相似文献   

15.
By use of a THF-containing trimethylsilylmethyl scandium catalyst system (C5Me4SiMe3)Sc(CH2SiMe3)2(THF)/[Ph3C][B(C6F5)4], the multi-component copolymerization of 10-bromo-1-decene (BrDC) with ethylene, propylene, and dienes has been achieved to afford a new family of bromine-functionalized polyolefins with controllable composition and high molecular weight. The copolymerization of BrDC with ethylene afforded the well-defined BrDC–ethylene copolymers with high BrDC incorporation (up to 12 mol%) and high molecular weight (Mw > 100 kg mol−1). The terpolymerization of propylene, ethylene with BrDC afforded random ethylene–propylene–BrDC terpolymers with controllable bromine content (2 ~ 11 mol%), high molecular weight (Mw > 100 kg mol−1) and low glass transition temperature (Tg = −51 °C ~ −67 °C). Moreover, the tetrapolymerization of ethylene, propylene, BrDC, and ethylidene norbornene or conjugated dienes such as isoprene and myrcene has been achieved for the first time to afford selectively the bromine-functionalized ethylene–propylene–diene rubbers containing various types of double bonds.  相似文献   

16.
New Aminometalanes of Aluminum and Gallium The reaction of secondary amines R′RNH with trimethyaluminum leads to the formation of dimeric aminoalanes [RR′NAlMe2]2 ( 1 ) (R = 2,6-Me2C6H3, R′ = SiMe2(2,4,6-Me3C6H2)) and 2 (R = Ph, R′ = SiMe3). Using a different stoichiometric ratio, a monomeric aminoalane [RR′N]2AlMe ( 3 ) (R = Ph, R′ = SiPh2Me) is obtained, having an aluminum atom of coordination number three due to the steric demand of the substituents. The synthesis of the corresponding aminogallanes 4 , 5 and 6 is achieved by reaction of lithium amides LiNRR′ (R = Ph, 2,6-iPr2C6H3; R′ = SiMe3, SiMe2iPr) with dimethylgalliumchloride, Me2GaCl, in n-hexane. The formation of the dimeric species is in 1 through carbon while that in 2 and 3 is formed through nitrogen. The X-ray single crystal structure analysis of 1 , 2 , 3 and 4 are reported.  相似文献   

17.
The reactions of PhCH2SiMe3 ( 1 ), PhCH2SiMe2tBu ( 2 ), PhCH2SiMe2Ph ( 3 ), 3,5‐Me2C6H3CH2SiMe3 ( 4 ), and 3,5‐Me2C6H3CH2SiMe2tBu ( 5 ) with nBuLi in tetramethylethylenediamine (tmeda) afford the corresponding lithium complexes [Li(tmeda)][CHRSiMe2R′] (R, R′ = Ph, Me ( 6 ), Ph, tBu ( 7 ), Ph, Ph ( 8 ), 3,5‐Me2C6H3, Me ( 9 ), and 3,5‐Me2C6H3, tBu ( 10 )), respectively. The new compounds 5 , 7 , 8 , 9 and 10 have been characterized by 1H and 13C NMR spectroscopy, compounds 7 , 8 and 9 also by X‐ray structure analysis.  相似文献   

18.
A rational approach to modulating easy-axis magnetic anisotropy by varying the axial donor ligand in heptacoordinated FeII complexes has been explored. In this series of complexes with formulae of [Fe(H4L)(NCS)2] ⋅ 3 DMF ⋅ 0.5 H2O ( 1 ), [Fe(H4L)(NCSe)2] ⋅ 3 DMF ⋅ 0.5 H2O ( 2 ), and [Fe(H4L)(NCNCN)2] ⋅ DMF ⋅ H2O ( 3 ) [H4L=2,2′-{pyridine-2,6-diylbis(ethan-1-yl-1-ylidene)}bis(N-phenylhydrazinecarboxamide)], the axial positions are successively occupied by different nitrogen-based π-donor ligands. Detailed dc and ac magnetic susceptibility measurements reveal the existence of easy-axis magnetic anisotropy for all of the complexes, with 1 [Ueff=21 K, τ0=1.72×10−6 s] and 2 [Ueff=25 K, τ0=2.25×10−6 s] showing field-induced slow magnetic relaxation behavior. However, both experimental studies and theoretical calculations indicate the magnitude of the D value of complex 3 to be larger than those of complexes 1 and 2 due to the axial bond angle being smaller than that for an ideal geometry. Detailed analysis of the field and temperature dependences of relaxation time for 1 and 2 has revealed that multiple relaxation processes (quantum tunneling of magnetization, direct, and Raman) are involved in slow magnetic relaxation for both of these complexes. Magnetic dilution experiments support the role of intermolecular short contacts.  相似文献   

19.
Novel molybdenum(VI/V) POM-based self-constructed frameworks [MoVI12O242-O)12(trz)6(H2O)6] ⋅ 6Hma ⋅ 18H2O ( 1 , Htrz=1H-1,2,3-triazole, ma=methylamine), [MoVI7O142-O)8(trz)5(H2O)] ⋅ 7Hma ⋅ 5H2O ( 2 ), Na3[MoV6O62-O)9(Htrz)3(trz)3] ⋅ 7.5H2O ( 3 ) and [MoV8O82-O)12(Htrz)8] ⋅ 30H2O ( 4 ) have been covalently decorated with tri-coordinated deprotonated/protonated 1,2,3-triazoles. Channels with an inner diameter of 7.5 Å were found in 1 , whereas a tunnel composed of stacking molecules with an inner diameter of 4.1 Å along the b-axis exists in 2 ; it is occupied by free disordered methylamines, showing selective adsorption of O2 and CO2 at 25 °C. Obvious downfield shifts were observed by 13C NMR spectroscopies for methylamines inside the confined channels in 1 and 2 . There are diversified pores in 3 and 4 , which are formed by the molecules themselves and intermolecular accumulations. Adsorption tests indicate that 3 and 4 are fine adsorption materials for CH4 and CO2 under low pressure that rely on the environments built by the POMs. Correspondingly, 1 and 2 display reversible photoresponsive thermochromism that is subtlety influenced by the channels. The polyoxometalate organic frameworks (POMOFs) with multiple functional adsorptions are easy to assemble. Their photo-/thermoresponse properties offer a new pathway for the self-constructions of one-off hybrid materials that possess the good properties of both POMs and MOFs.  相似文献   

20.
A series of ionic (mono-/di-)phosphines ( L2 , L4 , and L6 ) with structural similarity and their corresponding neutral counterparts ( L1 , L3 , and L5 ) were applied to modulate the catalytic performance of RuCl3 ⋅ 3H2O. With the involvement of the ionic diphosphine ( L4 ), in which the two phosphino-fragments were linked by butylene group, RuCl3 ⋅ 3H2O with advantages of low cost, robustness, and good availability was found to be an efficient and recyclable catalyst for the alkoxycarbonylation of aryl halides. The L4 -based RuCl3 ⋅ 3H2O system corresponded to the best conversion of PhI (96 %) along with 99 % selectivity to the target product of methyl benzoate as well as the good generality to alkoxycarbonylation of different aryl halides (ArX, X=I and Br) with alcohols MeOH, EtOH, i-PrOH and n-BuOH. The electronic and steric effects of the applied phosphines, which were analyzed by the 31P NMR for 1J31P-77Se1J measurement and single-crystal X-ray diffraction, were carefully co-related to the performance RuCl3 ⋅ 3H2O catalyst. In addition, the L4 -based RuCl3 ⋅ 3H2O system could be recycled successfully for at least eight runs in the ionic liquid [Bmim]PF6.  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-23

京公网安备 11010802026262号