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.
Summary [RuCl(NO)2(dppbp)]BF4 (dppbp=(Ph2PCH2)2–) has been synthesised from [RuCl(NO)2(PPh3)2]BF4 and dppbp and characterised in the solid state by a single crystal x-ray determination. The [RuCl(NO)2(dppbp)]+ cation, has an approximately square-pyramidal co-ordination geometry with the dppbp ligand occupyingtrans-basal sites. The nitrosyl ligand in the apical site is partially bent [Ru–N–O=156.2(7)0] and the nitrosyl ligand in the basal side is essentially linear [Ru–N–O=172.5(6)0]. The1Hn.m.r. spectrum of [RuCl(NO)2(dppbp)]BF4 in solution has provided some insight into the dynamics of the complex in solution.  相似文献   

2.
A structurally diverse range of lipophilic, cationic η6‐arene η5‐cyclopentadienyl (η5‐Cp*) full‐sandwich complexes of ruthenium(II) have been prepared and structurally characterized by Fourier‐transform IR and NMR spectroscopy, electrospray mass spectrometry, and elemental microanalyses. Computational experiments incorporating the Hartree–Fock theory and the second‐order Møller–Plesset perturbation theory predict each complex to possess a uniform δ+ electrostatic potential, with the cationic charge of the [RuCp*]+ moiety completely delocalizing throughout the molecular structure of each metallocene. In vitro cytotoxicity studies demonstrate these delocalized lipophilic cations to be potent growth inhibitors of eleven unique tumorigenic cell lines, while exhibiting significantly lower levels of toxicity towards both a normal human fibroblast and a mouse macrophage cell line. Single‐crystal X‐ray structural determinations are additionally reported for five complexes, [Ru(η6‐C6H5(CH2)2CH3)(η5‐C5(CH3)5)]BPh4, [Ru(η6‐C6H5CO2CH2CH3)(η5‐C5(CH3)5)]BF4, [Ru(η6‐C10H8)(η5‐C5(CH3)5)]BPh4, [Ru(η6‐C14H10)(η5‐C5(CH3)5)]BPh4, and [Ru(η6‐C16H10)(η5‐C5(CH3)5)]BPh4.  相似文献   

3.
The synthesis and characterization of heteroleptic complexes with the formulations [(η6-arene)RuCl(fcdpm)] (η6-arene = C6H6, C10H14) and [(η5-C5Me5)MCl(fcdpm)] (M = Rh, Ir; fcdpm = 5-ferrocenyldipyrromethene) have been reported. All the complexes have been characterized by elemental analyses, IR, 1H NMR and electronic spectral studies. Structures of [(η6-C6H6)RuCl(fcdpm)] and [(η6-C10H14)RuCl(fcdpm)] have been determined crystallographically. Chelating monoanionic linkage of fcdpm to the respective metal centres has been supported by spectral and structural studies. Further, reactivity of the representative complex [(η6-C10H14)RuCl(fcdpm)] with ammonium thiocyanate (NH4SCN) and triphenylphosphine (PPh3) have been examined.  相似文献   

4.
Formal [2 + 2 + 2] addition reactions of [Cp*Ru(H2O)(NBD)]BF4 (NBD = norbornadiene) with PhC?CR (R = H, COOEt) give [Cp*Ru(η6‐C6H5? C9H8R)] BF4 (1a, R = H; 2a, R = COOEt). Treatment of [Cp*Ru(H2O)(NBD)]BF4 with PhC?C? C?CPh does not give [2 + 2 + 2] addition product, but [Cp*Ru(η6‐C6H5? C?C? C?CPh)] BF4(3a). Treatment of 1a, 2a, 3a with NaBPh4 affords [Cp*Ru(η6‐C6H5? C9H8R)] BPh4 (1b, R = H; 2b, R = COOEt) and [Cp*Ru(η6‐C6H5? C?C? C?CPh)] BPh4(3b). The structures of 1b, 2b and 3b were determined by X‐ray crystallography. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

5.
The 1,5-bis(3,5-dimethyl-1-pyrazolyl)-3-thiapentane ligand (bdtp) reacts with [Rh(COD)(THF)2][BF4] to give [Rh(COD)(bdtp)][BF4] ([1][BF4]), which is fluxional in solution on the NMR time scale. Its further treatment with carbon monoxide leads to a displacement of the 1,5-cyclooctadiene ligand, generating a mixture of two complexes, namely, [Rh(CO)2(bdtp)][BF4] ([2][BF4]) and [Rh(CO)(bdtp3N,N,S)][BF4] ([3][BF4]). In solution, [2][BF4] exists as a mixture of two isomers, [Rh(CO)2(bdtp2N,N)]+ ([2a]+) and [Rh(CO)2(bdtp3N,N,S)]+ ([2b]+; major isomer) rapidly interconverting on the NMR time scale. At room temperature, [2][BF4] easily loses one molecule of carbon monoxide to give [3][BF4]. The latter is prone to react with carbon monoxide to partially regenerate [2][BF4]. The ligands 1,2-bis[3-(3,5-dimethyl-1-pyrazolyl)-2-thiapropyl]benzene (bddf) and 1,8-bis(3,5-dimethyl-1-pyrazolyl)-3,6-dithiaoctane (bddo) are seen to react with two equivalents of [Rh(COD)(THF)2][BF4] to give the dinuclear complexes [Rh2(bddf)(COD)2][BF4]2 ([4][BF4]2) and [Rh2(bddo)(COD)2][BF4]2 ([5][BF4]2), respectively. In such complexes, the ligand acts as a double pincer holding two rhodium atoms through a chelation involving S and N donor atoms. Bubbling carbon monoxide into a solution of [4][BF4]2 results in loss of the COD ligand and carbonylation to give [Rh2(bddf)(CO)4][BF4]2 ([6][BF4]2). The single-crystal X-ray structures of [3][CF3SO3], [5][BF4]2 and [6][BF4]2 are reported.  相似文献   

6.
Hydro(solvo)thermal reactions of Cd(NO3)2, N-(pyridin-3-ylmethyl)-4-(pyridin-4-yl)-1,8-naphthalimide (NI-mbpy-34), and 5-bromobenzene-1,3-dicarboxylic acid (Br-1,3-H2bdc) afforded a luminescent coordination polymer, {[Cd(Br-1,3-bdc)(NI-mbpy-34)(H2O)]∙2H2O}n (1). Single-crystal X-ray diffraction analysis showed that 1 features a two-dimensional (2-D) gridlike sql layer with the point symbol of (44·62), where the Cd(II) center adopts a {CdO5N2} pentagonal bipyramidal geometry. Thermogravimetric (TG) analysis confirmed the thermal stability of 1 up to about 340 °C, whereas XRPD patterns proved the maintenance of crystallinity and framework integrity of 1 in CH2Cl2, H2O, CH3OH, and toluene. Photoluminescence studies indicated that 1 displayed intense blue fluorescence emissions in both solid-state and H2O suspension-phase. Owing to the good fluorescent properties, 1 could serve as an excellent turn-off fluorescence sensor for selective and sensitive Cr(VI) detection in water, with LOD = 15.15 μM for CrO42 and 14.91 μM for Cr2O72, through energy competition absorption mechanism. In addition, 1 could also sensitively detect Cr3+, Fe3+, and Al3+ ions in aqueous medium via fluorescence-enhancement responses, with LOD = 2.81 μM for Cr3+, 3.82 μM for Fe3+, and 3.37 μM for Al3+, mainly through an absorbance-caused enhancement (ACE) mechanism.  相似文献   

7.
The coordination ability of the [(ppy)Au(IPr)]2+ fragment [ppy = 2-phenylpyridine, IPr = 1,3-bis(2,6-di-isopropylphenyl)-imidazol-2-ylidene] towards different anionic and neutral X ligands (X = Cl, BF4, OTf, H2O, 2-butyne, 3-hexyne) commonly involved in the crucial pre-equilibrium step of the alkyne hydration reaction is computationally investigated to shed light on unexpected experimental observations on its catalytic activity. Experiment reveals that BF4 and OTf have very similar coordination ability towards [(ppy)Au(IPr)]2+ and slightly less than water, whereas the alkyne complex could not be observed in solution at least at the NMR sensitivity. Due to the steric hindrance/dispersion interaction balance between X and IPr, the [(ppy)Au(IPr)]2+ fragment is computationally found to be much less selective than a model [(ppy)Au(NHC)]2+ (NHC = 1,3-dimethylimidazol-2-ylidene) fragment towards the different ligands, in particular OTf and BF4, in agreement with experiment. Effect of the ancillary ligand substitution demonstrates that the coordination ability of Au(III) is quantitatively strongly affected by the nature of the ligands (even more than the net charge of the complex) and that all the investigated gold fragments coordinate to alkynes more strongly than H2O. Remarkably, a stabilization of the water-coordinating species with respect to the alkyne-coordinating one can only be achieved within a microsolvation model, which reconciles theory with experiment. All the results reported here suggest that both the Au(III) fragment coordination ability and its proper computational modelling in the experimental conditions are fundamental issues for the design of efficient catalysts.  相似文献   

8.
The oxidation of pyruvic acid by the title silver(III) complex in aqueous acidic (pH, 1.1–4.5) media is described. The reaction products are MeCO2H and CO2, together with a colourless solution of the Ag+ ion. The free ligand, ethylenebis(biguanide) is released in near-quantitative yield upon completion of the reduction. The parent complex, [Ag(H2L)]3+ and one of its conjugate bases, [Ag(HL)]2+, participate in the reaction with both pyruvic acid (HPy) and the pyruvate anion (Py) as the reactive reducing species. Ag+ was found to be catalytically inactive. At 25.0°C, I=1.0moldm–3, rate constants for the reactions [Ag(H2L)]3++HPy (k 1), [Ag(H2L)]3++Py (k 2), [Ag(HL)]2++HPy (k 3) and [Ag(HL)]2++Py (k 4) arek 1=(94±6)×10–5dm3mol–1s–1, (k 2 K a+k 3 K a1)= (1.3±0.1)×10–5s–1 and k 4=(58±4)×10–5dm3mol–1s–1, respectively, where K a1is the first acid dissociation constant of the [Ag(H2L)]3+ and K a is for pyruvic acid. A comparison between the k 1 and k 4 values is indicative of the judgement that k 2k 3. A one-electron inner-sphere redox mechanism seems more justified than an outer-sphere electron-transfer between the redox partners.  相似文献   

9.
The reaction of 2,6-diethyl-4,8-dimethyl-s-indacenyl-dilithium (Li2Ic′) with [Cp*RuCl]4 gives the organometallic binuclear bis-pentamethylcyclopentadienyl-ruthenium-s-indacene complex, [{Cp*Ru}2Ic′] (1, Ic′ = 2,4-diethyl-4,8-dimethyl-s-indacene), in high yields. The subsequent oxidation of 1 with a ferricinium salt ([Fc]+[BF4]) gives the mixed valence compound [{Cp*Ru}2Ic′]+[BF4] (1+). Compound 1 was structurally characterized by X-ray crystallography, finding that both {Cp*Ru} fragments are coordinated to opposite sites of the Ic′ ligand. The structural and electronic features of 1 and 1+ have been rationalized by Density Functional Theory (DFT) calculations, which suggest that both metallic centers get closer to the Ic′ and subtle electronic reorganizations occurs when chemical oxidation takes place. Cyclic voltammetry and ESR experiments suggest a high electronic interaction between the metallic centers mediated by the Ic′ bridging ligand. Time dependent DFT (TD-DFT) calculations were carried out to understand and assign the intervalence band present in the mixed-valent specie (1+). The main achievement of this article is to feature the relationship of the experimental data with the computational results obtained with the Amsterdam Density Functional package (ADF). Both experimental and theoretical facts demonstrate that the mixed valence system (1+) is a delocalized one, and it can be classified as a Class III system according to the Robin & Day classification.  相似文献   

10.
The oxidative addition of CH3I to planar rhodium(I) complex [Rh(TFA)(PPh3)2] in acetonitrile (TFA is trifluoroacetylacetonate) leads to the formation of cationic, cis-[Rh(TFA)(PPh3)2(CH3)(CH3CN)][BPh4] (1), or neutral, cis-[Rh(TFA)(PPh3)2(CH3)(I)] (4), rhodium(III) methyl complexes depending on the reaction conditions. 1 reacts readily with NH3 and pyridine to form cationic complexes, cis-[Rh(TFA)(PPh3)2(CH3)(NH3)][BPh4] (2) and cis-[Rh(TFA)(PPh3)2(CH3)(Py)][BPh4] (3), respectively. Acetylacetonate methyl complex of rhodium(III), cis-[Rh(Acac)(PPh3)2(CH3)(I)] (5), was obtained by the action of NaI on cis-[Rh(Acac)(PPh3)2(CH3)(CH3CN)][BPh4] in acetone at −15 °C. Complexes 1-5 were characterized by elemental analysis, 31P{1H}, 1H and 19F NMR. For complexes 2, 3, 4 conductivity data in acetone solutions are reported. The crystal structures of 2 and 3 were determined. NMR parameters of 1-5 and related complexes are discussed from the viewpoint of their isomerism.  相似文献   

11.
Four new compounds of formulas [Cu(hfac)2(L)] (1), [Ni(hfac)2(L)] (2), [{Cu(hfac)2}2(µ-L)]·2CH3OH (3) and [{Ni(hfac)2}2(µ-L)]·2CH3CN (4) [Hhfac = hexafluoroacetylacetone and L = 3,6-bis(picolylamino)-1,2,4,5-tetrazine] have been prepared and their structures determined by X-ray diffraction on single crystals. Compounds 1 and 2 are isostructural mononuclear complexes where the metal ions [copper(II) (1) and nickel(II) (2)] are six-coordinated in distorted octahedral MN2O4 surroundings which are built by two bidentate hfac ligands plus another bidentate L molecule. This last ligand coordinates to the metal ions through the nitrogen atoms of the picolylamine fragment. Compounds 3 and 4 are centrosymmetric homodinuclear compounds where two bidentate hfac units are the bidentate capping ligands at each metal center and a bis-bidentate L molecule acts as a bridge. The values of the intramolecular metal···metal separation are 7.97 (3) and 7.82 Å (4). Static (dc) magnetic susceptibility measurements were carried out for polycrystalline samples 1–4 in the temperature range 1.9–300 K. Curie law behaviors were observed for 1 and 2, the downturn of χMT in the low temperature region for 2 being due to the zero-field splitting of the nickel(II) ion. Very weak [J = −0.247(2) cm−1] and relatively weak intramolecular antiferromagnetic interactions [J = −4.86(2) cm−1] occurred in 3 and 4, respectively (the spin Hamiltonian being defined as H = −JS1·S2). Simple symmetry considerations about the overlap between the magnetic orbitals across the extended bis-bidentate L bridge in 3 and 4 account for their magnetic properties.  相似文献   

12.
The differences between the molecular structures of the PCP-pincer complex [RuCl{C6H3(CH2P(C6H5)2)2-2,6}(PPh3)] ([RuCl(PCPH)(PPh3)], 1) and its tetrakis-pentafluorophenyl substituted analogue [RuCl{C6H3(CH2P(C6F5)2)2-2,6}(PPh3)] ([RuCl(PCPF20)(PPh3)], 2) have been rationalised by performing calculations on the cations [Ru(PCPH)(PPh3)]+ (1cat) and [Ru(PCPF20)(PPh3)]+ (2cat). The molecular interactions between the chloride ligand and the axial rings, as found in 1 and 2, respectively, have been studied computationally in the model systems [(C6X5PH2)2Cl] (X = H, F). The calculations on 2cat show that in 2 it is most likely the attractive electrostatic interaction between the chloride ligand and the fluorinated phenyl rings that forces the Cipso atom to occupy an axial position rather than an equatorial one in the observed (X-ray of 2) square pyramidal arrangement. In 1, however, repulsive steric hindrance forces the PPh3 ligand to take the apical position. The applicability of the TD-DFT method for the calculation of the electronic spectra of the PCP-pincer compounds 1 and 2 has been tested. The results indicate that the excitation energies calculated for both complexes are in a reasonable agreement with the experimental absorption maxima. However, for 1, all the calculated transition energies are underestimated.  相似文献   

13.
The oxidation of transition metals such as manganese and copper by dioxygen (O2) is of great interest to chemists and biochemists for fundamental and practical reasons. In this report, the O2 reactivities of 1:1 and 1:2 mixtures of [(TPP)MnII] (1; TPP: Tetraphenylporphyrin) and [(tmpa)CuI(MeCN)]+ (2; TMPA: Tris(2-pyridylmethyl)amine) in 2-methyltetrahydrofuran (MeTHF) are described. Variable-temperature (−110 °C to room temperature) absorption spectroscopic measurements support that, at low temperature, oxygenation of the (TPP)Mn/Cu mixtures leads to rapid formation of a cupric superoxo intermediate, [(tmpa)CuII(O2•–)]+ (3), independent of the presence of the manganese porphyrin complex (1). Complex 3 subsequently reacts with 1 to form a heterobinuclear μ-peroxo species, [(tmpa)CuII–(O22–)–MnIII(TPP)]+ (4; λmax = 443 nm), which thermally converts to a μ-oxo complex, [(tmpa)CuII–O–MnIII(TPP)]+ (5; λmax = 434 and 466 nm), confirmed by electrospray ionization mass spectrometry and nuclear magnetic resonance spectroscopy. In the 1:2 (TPP)Mn/Cu mixture, 4 is subsequently attacked by a second equivalent of 3, giving a bis-μ-peroxo species, i.e., [(tmpa)CuII−(O22−)−MnIV(TPP)−(O22−)−CuII(tmpa)]2+ (7; λmax = 420 nm and δpyrrolic = −44.90 ppm). The final decomposition product of the (TPP)Mn/Cu/O2 chemistry in MeTHF is [(TPP)MnIII(MeTHF)2]+ (6), whose X-ray structure is also presented and compared to literature analogs.  相似文献   

14.
The reactivity of [Pt2(μ-S)2(PPh3)4] towards [RuCl26-arene)]2 (arene=C6H6, C6Me6, p-MeC6H4Pri=p-cymene), [OsCl26-p-cymene)]2 and [MCl25-C5Me5)]2 (M=Rh, Ir) have been probed using electrospray ionisation mass spectrometry. In all cases, dicationic products of the type [Pt2(μ-S)2(PPh3)4ML]2+ (L=π-hydrocarbon ligand) are observed, and a number of complexes have been prepared on the synthetic scale, isolated as their BPh4 or PF6 salts, and fully characterised. A single-crystal X-ray structure determination on the Ru p-cymene derivative confirms the presence of a pseudo-five-coordinate Ru centre. This resists addition of small donor ligands such as CO and pyridine. The reaction of [Pt2(μ-S)2(PPh3)4] with RuClCp(PPh3)2 (Cp=η5-C5H5) gives [Pt2(μ-S)2(PPh3)4RuCp]+. In addition, the reaction of [Pt2(μ-S)2(PPh3)4] with the related carbonyl complex [RuCl2(CO)3]2, monitored by electrospray mass spectrometry, gives [Pt2(μ-S)2(PPh3)4Ru(CO)3Cl]+.  相似文献   

15.
The new cationic mononuclear complexes [(η6-arene)Ru(Ph-BIAN)Cl]BF46-arene = benzene (1), p-cymene (2)], [(η5-C5H5)Ru(Ph-BIAN)PPh3]BF4 (3) and [(η5-C5Me5)M(Ph-BIAN)Cl]BF4 [M = Rh (4), Ir (5)] incorporating 1,2-bis(phenylimino)acenaphthene (Ph-BIAN) are reported. The complexes have been fully characterized by analytical and spectral (IR, NMR, FAB-MS, electronic and emission) studies. The molecular structure of the representative iridium complex [(η5-C5Me5)Ir(Ph-BIAN)Cl]BF4 has been determined crystallographically. Complexes 15 effectively catalyze the reduction of terephthaldehyde in the presence of HCOOH/CH3COONa in water under aerobic conditions and, among these complexes the rhodium complex [(η5-C5Me5)Rh(Ph-BIAN)Cl]BF4 (4) displays the most effective catalytic activity.  相似文献   

16.
The diastereoselective κ2-P,N-coordination of a chiral tricyclic β-iminophosphine ligand to the half-sandwich ruthenium(II) fragments [RuCl(η6-arene)]+ (arene = C6H6, p-cymene, 1,3,5-C6H3Me3, C6Me6), [Ru(η6-p-cymene)(NCMe)]2+ and [Ru(η5-C5H5)(NCMe)]+ is described. The structures of the resulting mono- and dicationic cymene derivatives have been confirmed by X-ray crystallography. Studies on the catalytic activity of these Ru(II) compounds in Diels–Alder cycloaddition processes are also reported.  相似文献   

17.
The self-assembly of 2,6-diformyl-4-methylphenol (DFMP) and 1-amino-2-propanol (AP)/2-amino-1,3-propanediol (APD) in the presence of copper(II) ions results in the formation of six new supramolecular architectures containing two versatile double Schiff base ligands (H3L and H5L1) with one-, two-, or three-dimensional structures involving diverse nuclearities: tetranuclear [Cu4(HL2−)2(N3)4]·4CH3OH·56H2O (1) and [Cu4(L3−)2(OH)2(H2O)2] (2), dinuclear [Cu2(H3L12−)(N3)(H2O)(NO3)] (3), polynuclear {[Cu2(H3L12−)(H2O)(BF4)(N3)]·H2O}n (4), heptanuclear [Cu7(H3L12−)2(O)2(C6H5CO2)6]·6CH3OH·44H2O (5), and decanuclear [Cu10(H3L12−)4(O)2(OH)2(C6H5CO2)4] (C6H5CO2)2·20H2O (6). X-ray studies have revealed that the basic building block in 1, 3, and 4 is comprised of two copper centers bridged through one μ-phenolate oxygen atom from HL2− or H3L12−, and one μ-1,1-azido (N3) ion and in 2, 5, and 6 by μ-phenoxide oxygen of L3− or H3L12− and μ-O2− or μ3-O2− ions. H-bonding involving coordinated/uncoordinated hydroxy groups of the ligands generates fascinating supramolecular architectures with 1D-single chains (1 and 6), 2D-sheets (3), and 3D-structures (4). In 5, benzoate ions display four different coordination modes, which, in our opinion, is unprecedented and constitutes a new discovery. In 1, 3, and 5, Cu(II) ions in [Cu2] units are antiferromagnetically coupled, with J ranging from −177 to −278 cm−1.  相似文献   

18.
Summary The pentadentate macrocycle 1,4,7,10,13-penta-azacyclo-hexadecane [16]aneN5=(3)=L} has been prepared and a variety of copper(II), nickel(II) and cobalt(III) complexes of the ligand characterised. The copper complex [CuL](ClO4)2, on the basis of its d-d spectrum, appears to be square pyramidal, while [NiL(H2O)](ClO4)2 is octahedral. The copper(II) and nickel(II) complexes dissociate readily in acidic solution and these reactions have been studied kinetically. For the copper(II) complex, rate=kH[complex][H+]2 with kH =4.8 dm6 mol–2s–1 at 25 °C and I=1.0 mol dm–3 (NaClO4) with H=43 kJ mol–1 and S 298 =–89 JK–1 mol–1. Dissociation rates of the copper(II) complexes increase with ring size in the order: [15]aneN5 < [16]aneN5 < [17]aneN5. For the dissociation of the nickel(II) complex, rate=kH[Complex][H+] with kH=9.4×10–3 dm3mol–1 s–1 at 25 °C and I =1.0 mol dm–3 (NaClO4) with H=71 kJ mol–1 and S 298 =–47 JK–1mol–1.The cobalt(III) complexes, [CoLCl](ClO4)2, [CoL(H2O)]-(ClO4)3, [CoL(NO2)](ClO4)2, [CoL(DMF)](ClO4)3 (DMF=dimethylformamide) and [CoL(O2CH)](ClO4)2 have been characterised. The chloropentamine [CoCl([16]aneN5)]2+ undergoes rapid base hydrolysis with kOH=1.1× 105dm3 mol–1s–1 at 25°C and I=0.1 mol dm–3 (H=73 kJ mol–1 and S 298 =98 JK–1 mol–1). Rapid base hydrolysis of [CoL(NO2)]2+ is also observed and the origins of these effects are considered in detail.  相似文献   

19.
Halide abstraction from [Pd(μ-Cl)(Fmes)(NCMe)]2 (Fmes = 2,4,6-tris(trifluoromethyl)phenyl or nonafluoromesityl) with TlBF4 in CH2Cl2/MeCN gives [Pd(Fmes)(NCMe)3]BF4, which reacts with monodentate ligands to give the monosubstituted products trans-[Pd(Fmes)L(NCMe)2]BF4 (L = PPh3, P(o-Tol)3, 3,5-lut, 2,4-lut, 2,6-lut; lut = dimethylpyridine), the disubstituted products trans-[Pd(Fmes)(NCMe)(PPh3)2]BF4, cis-[Pd(Fmes)(3,5-lut)2(NCMe)]BF4, or the trisubstituted products [Pd(Fmes)L3]BF4 (L = CNtBu, PHPh2, 3,5-lut, 2,4-lut). Similar reactions using bidentate chelating ligands give [Pd(Fmes)(L-L)(NCMe)]BF4 (L-L = bipy, tmeda, dppe, OPPhPy2-N,N′, (OH)(CH3)CPy2-N,N′). The complexes trans-[Pd(Fmes)L2(NCMe)]BF4 (L = PPh3, tht) (tht = tetrahydrothiophene) and [Pd(Fmes)(L-L)(NCMe)]BF4 (L-L = bipy, tmeda) were obtained by halide extraction with TlBF4 in CH2Cl2/MeCN from the corresponding neutral halogeno complexes trans-[Pd(Fmes)ClL2] or [Pd(Fmes)Cl(L-L)]. The aqua complex trans-[Pd(Fmes)(OH2)(tht)2]BF4 was isolated from the corresponding acetonitrile complex. Overall, the experimental results on these substitution reactions involving bulky ligands suggest that thermodynamic and kinetic steric effects can prevail affording products or intermediates different from those expected on purely electronic considerations. Thus,water, whether added on purpose or adventitious in the solvent, frequently replaces in part other better donor ligands, suggesting that the smaller congestion with water compensates for the smaller M-OH2 bond energy.  相似文献   

20.
Homogeneous olefin polymerization catalysts are activated in situ with a co-catalyst ([PhN(Me)2-H]+[B(C6F5)4] or [Ph3C]+[B(C6F5)4]) in bulk polymerization media. These co-catalysts are insoluble in hydrocarbon solvents, requiring excess co-catalyst (>3 eq.). Feeding the activated species as a solution in an aliphatic hydrocarbon solvent may be advantageous over the in situ activation method. In this study, highly pure and soluble ammonium tetrakis(pentafluorophenyl)borates ([Me(C18H37)2N-H]+[B(C6F5)4] and [(C18H37)2NH2]+[B(C6F5)4]) containing neither water nor Cl salt impurities were prepared easily via the acid–base reaction of [PhN(Me)2-H]+[B(C6F5)4] and the corresponding amine. Using the prepared ammonium salts, the activation reactions of commercial-process-relevant metallocene (rac-[ethylenebis(tetrahydroindenyl)]Zr(Me)2 (1-ZrMe2), [Ph2C(Cp)(3,6-tBu2Flu)]Hf(Me)2 (3-HfMe2), [Ph2C(Cp)(2,7-tBu2Flu)]Hf(Me)2 (4-HfMe2)) and half-metallocene complexes ([(η5-Me4C5)Si(Me)2(κ-NtBu)]Ti(Me)2 (5-TiMe2), [(η5-Me4C5)(C9H9(κ-N))]Ti(Me)2 (6-TiMe2), and [(η5-Me3C7H1S)(C10H11(κ-N))]Ti(Me)2 (7-TiMe2)) were monitored in C6D12 with 1H NMR spectroscopy. Stable [L-M(Me)(NMe(C18H37)2)]+[B(C6F5)4] species were cleanly generated from 1-ZrMe2, 3-HfMe2, and 4-HfMe2, while the species types generated from 5-TiMe2, 6-TiMe2, and 7-TiMe2 were unstable for subsequent transformation to other species (presumably, [L-Ti(CH2N(C18H37)2)]+[B(C6F5)4]-type species). [L-TiCl(N(H)(C18H37)2)]+[B(C6F5)4]-type species were also prepared from 5-TiCl(Me) and 6-TiCl(Me), which were newly prepared in this study. The prepared [L-M(Me)(NMe(C18H37)2)]+[B(C6F5)4]-, [L-Ti(CH2N(C18H37)2)]+[B(C6F5)4]-, and [L-TiCl(N(H)(C18H37)2)]+[B(C6F5)4]-type species, which are soluble and stable in aliphatic hydrocarbon solvents, were highly active in ethylene/1-octene copolymerization performed in aliphatic hydrocarbon solvents.  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