共查询到20条相似文献,搜索用时 15 毫秒
1.
Shiotsuki M White PS Brookhart M Templeton JL 《Journal of the American Chemical Society》2007,129(13):4058-4067
The diimine platinum(II) ethylene hydride complex [(N/\N)Pt(H)(ethylene)][BAr'4] (1, N/\N = [(2,6-Me2C6H3)N=C(An)-C(An)=N(2,6-Me2C6H3)], An = 1,8-naphthalenediyl, Ar' = 3,5-(CF3)2C6H3) was prepared by protonation of the diethyl complex (N/\N)PtEt2 with [H(OEt2)2][BAr'4]. The energy barrier to interchange of the platinum hydride with the olefinic hydrogens in 1 was determined to be 19.2 kcal/mol by spin saturation transfer experiments. Complex 1 initiates ethylene dimerization; the ethyl ethylene complex (N/\N)Pt(Et)(ethylene)+ (2) has been identified as the catalyst resting state. Trapping of 1 by ethylene to yield 2 is a second-order process; kinetic studies suggest this occurs via trapping of a reversibly formed beta-agostic ethyl complex. Complex 2 has been isolated and characterized by X-ray crystallography. The barrier to migratory insertion of 2, the turnover-limiting step in catalysis, was determined to be 29.8 kcal/mol. The 1-butene hydride complex, (N/\N)Pt(H)(1-butene)+ (3), is a key intermediate in the dimerization cycle and has also been isolated and characterized. Surprisingly rapid rates of degenerate associative exchange of free ethylene with bound ethylene in complexes 1 and 2 as well as the rate of degenerate exchange of free nitrile with bound nitrile in (N/\N)Pt(Et)(CH3CN)+ are reported. 相似文献
2.
3.
Lachaize S Sabo-Etienne S Donnadieu B Chaudret B 《Chemical communications (Cambridge, England)》2003,(2):214-215
Silylation of ethylene by chlorosilanes is catalysed by ruthenium complexes. Mechanistic investigations reveal the presence of a complicated network of reactions leading to new sigma-silane, ethylene and silyl complexes. 相似文献
4.
Takaya Matsumoto Roy A Periana Douglas J Taube Hajime Yoshida 《Journal of molecular catalysis. A, Chemical》2002,180(1-2):1-18
The novel, anti-Markovnikov, arylation of olefins with benzene to produce straight-chain alkylbenzenes with higher selectivity than branched alkylbenzenes is catalyzed by [Ir(μ-acac-O,O′,C3)(acac-O,O′)(acac-C3)]2 (acac=acetylacetonato), 1 [J. Am. Chem. Soc. 122 (2000) 7414]. The reaction of benzene with propylene gave n-propylbenzene and cumene in 61 and 39% selectivities, respectively. The reaction of benzene and styrene afforded 1,2-diphenylethane in 98% selectivity. Considering the anti-Markovnikov regioselectivity and lack of inhibition by water, we propose that the reaction does not proceed via a Friedel–Crafts, carbocation, mechanism. Complex 1, amongst the various transition metal complexes examined, is the most efficient for catalyzing the anti-Markovnikov olefin arylation. The crystal structure of complex 1 was solved and is consistent with a binuclear Ir(III) structure with three different types of coordinated acac ligands as reported by earlier solution IR and NMR analyses. [Ir(μ-acac-O,O′,C3)(acac-O,O′)Cl]2, 2, was prepared by the reaction of complex 1 with benzoyl chloride, and the crystal structure was also reported. 相似文献
5.
Leatherman MD Svejda SA Johnson LK Brookhart M 《Journal of the American Chemical Society》2003,125(10):3068-3081
The synthesis of a series of (alpha-diimine)NiR(2) (R = Et, (n)Pr) complexes via Grignard alkylation of the corresponding (alpha-diimine)NiBr(2) precursors is presented. Protonation of these species by the oxonium acid [H(OEt(2))(2)](+)[BAr'(4)](-) at low temperatures yields cationic Ni(II) beta-agostic alkyl complexes which model relevant intermediates present in nickel-catalyzed olefin polymerization reactions. The highly dynamic nature of these agostic alkyl cations is quantitatively addressed using NMR line broadening techniques. Trapping of these complexes with ethylene provides cationic Ni alkyl ethylene species, which are used to determine rates of ethylene insertion into primary and secondary carbon centers. The Ni agostic alkyl cations are also trapped by CH(3)CN and Me(2)S to yield Ni(R)(L)(+) (L = CH(3)CN, Me(2)S) complexes, and the dynamic behavior of these species in the presence of varied [L] is discussed. The kinetic data obtained from these experiments are used to present an overall picture of the ethylene polymerization mechanism for (alpha-diimine)Ni catalysts, including effects of reaction temperature and ethylene pressure on catalyst activity, polyethylene branching, and polymer architecture. Detailed comparisons of these systems to the previously presented analogous palladium catalysts are made. 相似文献
6.
Reaction of bis(phosphinite) PCP iridium pincer complexes (p-XPCP)IrHCl (5a-f) [X = MeO (5a), Me (5b), H (5c), F (5d), C(6)F(5) (5e), Ar(F)(= 3,5-bis(trifluoromethyl)phenyl) (5f)] with NaOtBu in neat cyclooctane (COA) generates 1:1 mixtures of the respective (p-XPCP)IrH(2) complexes 4a-f and the cyclooctene (COE) olefin complexes (p-XPCP)Ir(COE) (6a-f) at 23 degrees C. At higher temperatures, complexes 4 and 6 are equilibrated because of the degenerate transfer dehydrogenation of COA with free COE (6 + COA right harpoon over left harpoon 4 + 2COE), as was shown by temperature-dependent equilibrium constants and spin saturation transfer experiments at 80 degrees C. At this temperature, the COE complexes 6 exchange with free COE on the NMR time scale with the more electron-deficient complexes 6 exchanging COE faster. The exchange is dissociative and zero order in [COE]. Further analysis reveals that the stoichiometric hydrogenation of COE by complex 4f, and thus the separated back reaction 4f + 2COE --> 6f + COA proceeds at temperatures as low as -100 degrees C with the intermediacy of two isomeric complexes (p-Ar(F)PCP)Ir(H)(2)(COE) (8f, 8f'). COE deuteration with the perdeuterated complex 4f-d(38) at -100 degrees C results in hydrogen incorporation into the hydridic sites of complexes 8f,8f'-d(38) but not in the hydridic sites of complex 4f-d(38), thus rendering COE migratory insertion in complexes 8f,8f' reversible and COE coordination by complex 4f rate-determining for the overall COE deuteration. 相似文献
7.
Moorlag C Clot O Wolf MO Patrick BO 《Chemical communications (Cambridge, England)》2002,(24):3028-3029
Switching between P,S- and P,C coordination modes of 3'-phosphinoterthiophene to Ru(II) results in substantial differences in the electronic spectra and cyclic voltammetry of these complexes. 相似文献
8.
Zhang R Sun S Peng X Sun L 《European journal of mass spectrometry (Chichester, England)》2004,10(5):599-603
Electrospray ionization mass spectrometry (ESMS) has been employed to study the formation of fragment ions of a series of rhenium(I) bipyridyl complexes [(4,4'-di-(COOEt)2-bpy) Re(CO)3XPyPF6], where bpy is 2,2'-bipyridine, Py is pyridine, and X is H, 4-methyl, 3-methyl, 4-hydroxyl, 3-hydroxyl, 4-amino, or 3-amino of the pyridine ligand. The effects of substituents (X) on the stabilities of the complexes have been investigated with the increase of fragmentor voltages. For different X, the stabilities of the complexes increase as X become more electron-donating from H to CH3, OH, and NH2. For the same substituent, the p-substituted pyridines have stronger stabilizing effect than the corresponding m-substituted ones. Ligand exchange reaction was found in acetonitrile, where the pyridine ligand has been replaced by the solvent indicated by the formation of [M-PF6- XPy+MeCN]+ in the fragmentation. 相似文献
9.
Photophysical properties for a number ruthenium(II) and osmium(II) bipyridyl complexes are reported in dilute acetonitrile solution. The lifetimes of the excited metal to ligand charge transfer states (MLCT) of the osmium complexes are shorter than for the ruthenium complexes. Rate constants, kq, for quenching of the lowest excited metal to ligand charge transfer states by molecular oxygen are found to be in the range (1.1-7.7) x 10(9) dm3 mol(-1) s(-1). Efficiencies of singlet oxygen production, fDeltaT, following oxygen quenching of the lowest excited states of these ruthenium and osmium complexes are in the range of 0.10-0.72, lower values being associated with those compounds having lower oxidation potentials. The rate constants for quenching of the excited MLCT states, kq, are found to be generally higher for osmium complexes than for ruthenium complexes. Overall quenching rate constants, kq were found to give an inverse correlation with the energy of the excited state being quenched, and also to correlate with the oxidation potentials of the complexes. However, when the contribution of quenching due exclusively to energy transfer to produce singlet oxygen, kq1, is considered, its dependence on the energy of the excited states is more complex. Rate constants for quenching due to energy dissipation of the excited MLCT states without energy transfer, kq3, were found to show a clear correlation with the oxidation potential of the complexes. Factors affecting both the mechanism of oxygen quenching of the excited states and the efficiency of singlet oxygen generation following this quenching are discussed. These factors include the oxidation potential, the energy of the lowest excited state of the complexes and spin-orbit coupling constant of the central metal. 相似文献
10.
Peter Illner Ralph Puchta Nico van Eikema Hommes Rudi van Eldik 《Journal of organometallic chemistry》2005,690(15):3567-3576
The kinetics of the formation of the active species cis-[PtII(PPh3)2Cl(SnCl3)] and cis-[PtII(PPh3)2(SnCl3)2] from the hydroformylation catalyst precursor cis-[PtII(PPh3)2Cl2] in the presence of SnCl2, was studied in two different imidazolium-based ionic liquids. A large range of different chlorostannate melts consisting of 1-butyl-3-methyl-imidazolium cations and [SnxCly](−y + 2x) anions with varying molar fraction of SnCl2, were prepared and characterized by 1H and 119Sn NMR. The observed chemical shifts point to major changes in the composition of the anionic species within the melt. The second ionic liquid employed, viz., 1-butyl-3-methyl-imidazolium-bis(trifluormethylsulfonyl)amide was prepared in a colorless quality that enabled its application in kinetic studies. The concentration and temperature dependence of the substitution of Cl− by [SnCl3]− to yield cis-[PtII(PPh3)2Cl(SnCl3)], could be studied in detail. Theoretical (DFT) calculations were employed to model the reaction progress and to resolve the role of the ionic liquid in the activation of the catalyst. The available results are presented and a plausible mechanism for the formation of the catalytically active species is suggested. 相似文献
11.
The synthesis and characterization of a series of ruthenium bis(bipyridine) complexes where the inorganic moiety acts as a photolabile protecting group is described. Complexes of the type [Ru(bpy)2L2]+ where bpy = 2,2'-bipyridine and L = butylamine, gamma-aminobutyric acid, tyramine, tryptamine, and serotonin were studied by nuclear magnetic resonance, cyclic voltammetry, and electronic absorption spectroscopy. In all cases, ligands are coordinated by the amine group. The complexes are stable in water for several days and deliver one molecule of ligand upon irradiation with visible light (450 nm). These properties make them suitable for their use as biological caged compounds. 相似文献
12.
There have been synthesized Pt(II) stereoisomeric complexes with hydroxy-α-amino acid serine (SerH = NH2CH(CH2OH)COOH is α-amino-β-hydroxypropionic acid): trans-[Pt(S-SerH)2Cl2], trans-[Pt(R-SerH)(S-SerH)Cl2] with monodentately (through NH2 group ) bound SerH and cis-, trans-[Pt(R-Ser)(S-Ser)], trans-[Pt(S-Ser)2] with bidentately bound (through groups NH2 and COO) ligands (R, S is the absolute configuration of asymmetric carbon atom). The successive phases in the synthesis of Pt(II) stereoisomeric complexes with serine were studied by 195Pt NMR spectroscopy. To identificate the compounds synthesized the method of elemental analysis, IR and NMR (195Pt, 13C, 1H) spectroscopy were used. For trans-[Pt(R-Ser)(S-Ser)] the X-ray diffraction data were obtained. 相似文献
13.
Grant GJ 《Dalton transactions (Cambridge, England : 2003)》2012,41(29):8745-8761
Platinum(II) and palladium(II) complexes of the trithiacrown [9]aneS(3) containing a range of Group 15 donors are reviewed. These complexes have the general formula [M([9]aneS(3))(L(2))](n+) where L represents at least one Group 15 donor. Complexes involving pnictogens, with the exception of bismuth, are observed. The complexes generally have an elongated square pyramidal geometry with a long distance interaction to the third sulphur of the [9]aneS(3) which forms the apex of the square pyramid. This axial metal-sulphur distance is quite sensitive to the donor properties of L. Poorer donors such as Sb and As ligands show short axial distances whereas the better N donor ligands show longer distances. Pt(II) complexes of the formula [Pt([9]aneS(3))(EPh(3))(2)](2+) (E = P, As, Sb) show a considerable distortion towards a trigonal bipyramidal geometry due to intramolecular π-π interactions. Over seventy of these types of complexes have been crystallographically characterized and are discussed in this article. Other unique features of the complexes, including NMR spectroscopy, redox chemistry, and electronic spectroscopy, are also discussed. 相似文献
14.
15.
The rate constants and distribution of deuterated products of the H-D exchange in alkanes (C1–C8) have been measured for aqueous solutions of K2PtCl4. The reaction involves the intermediate formation of alkyl, 1,1-(carbene). 1,2-(olefin), 1,3-(cyclopropane) and 1,4-(cyclobutane) type hydrocarbon complexes. The selective attack of Pt(II) on the primary, secondary and tertiary C–H bonds in alkanes (110) is due to the predominance of steric over polar effects.
H-D (C1–C8) K2PtCl4 , , 1,1-(), 1,2-(), 1,3-() 1,4-() . Pt(II) ., ., . C–H (110) .相似文献
16.
Luo Mei Yin Hao Zhang Jia Hai Hu Ke Liang Pang Wen Min 《Research on Chemical Intermediates》2009,35(1):123-136
A novel family of oxazolinyl copper(II) catalysts have been developed and used as Lewis acid catalysts in the asymmetric Henry
reaction of various aldehydes with nitromethane. The corresponding nitroalcohol products were obtained in moderate yields
(40–80%) and with moderate enantioselectivity (10–40% ee). 相似文献
17.
Photophysical properties of Ru(II) bipyridyl complexes containing hemilabile phosphine-ether ligands
Emission and absorbance spectra, along with low-temperature excited-state lifetimes, were obtained for the hemilabile complexes, [Ru(bpy)2L](PF6)2 [L = (2-methoxyphenyl)diphenylphosphine (RuPOMe) (1) and (2-ethoxyphenyl)diphenylphosphine (RuPOEt) (2)] in solid 4:1 ethanol/methanol solution. Spectral data were evaluated with ground-state reduction potentials using Lever parameters. Lifetime data for these complexes were collected from 77 to 160 K, and the rate constant for the combined radiative and nonradiative decay process, k, the thermally activated process prefactor, k'(0), the rate constant for the MLCT --> d-d transition, k', and the activation energy, DeltaE', were calculated from a plot of ln(1/tau) versus 1/T for both (1) and (2). The low-temperature luminescence lifetimes of (1) were observed to decrease with increases in water concentration. The photophysical and kinetic data of (1) and (2) are compared to literature data for [Ru(bpy)3](PF6)2. The emission maxima of (1) and (2) are blue-shifted relative to [Ru(bpy)3](PF6)2 due to the presence of the strong-field phosphine ligand, which enhances pi back-bonding to the bipyridyl ligands. The thermal activation energy, DeltaE', is significantly larger for [Ru(bpy)3](PF6)2 than for (1) and (2) resulting in a faster MLCT --> d-d transition for (1) and (2). These results are discussed in the context of radiationless decay through thermally activated ligand-field states on the metal complex. 相似文献
18.
Qi Wang Lidong Li Zhiqiang Fan 《Journal of polymer science. Part A, Polymer chemistry》2005,43(8):1599-1606
In the presence of tetraethylaluminoxane (TEAO), iron complexes were used to catalyze ethylene polymerizations with extremely high activities and generally produced polyethylene with a bimodal molecular weight distribution (MWD). This bimodal MWD of polyethylene was mainly derived from residual triethylaluminum in TEAO and was produced through a mechanism of chain transfer to aluminum. Ethylaluminoxane and tetraisobutylaluminoxane also were used to polymerize ethylene with high activities in the presence of iron complexes, and only polyethylene with a unimodal MWD was produced. The ratio of the rate constant of chain transfer to aluminum (ktrA) to the rate constant of chain propagation (kp) was determined to be 0.12 for {[ArN?C(Me)]2C5H3N}FeCl2 when Ar was 2,6‐diisopropylphenyl ( 1 ) and 2.48 for {[ArN?C(Me)]2C5H3N}FeCl2 when Ar was 2,6‐dimethylphenyl ( 2 ); these values are far larger than those for metallocene‐based catalysts. This explains why an iron complex usually produces polyethylene with a broader MWD than metallocene‐based catalysts. Additionally, it can be concluded from the great difference between 1 and 2 with respect to ktrA/kp that an iron complex with less congested aryl substituents is subjected to chain transfer to aluminum. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 1599–1606, 2005 相似文献
19.
A series of ruthenium complexes having the general form [Ru(bpy)(3-n)(CN-Me-bpy)(n)](PF(6))(2) (where bpy = 2,2'-bipyridine, CN-Me-bpy = 4,4'-dicyano-5,5'-dimethyl-2,2'-bipyridine, and n = 1-3 for complexes 1-3, respectively) have been synthesized and characterized using a variety of steady-state and nanosecond time-resolved spectroscopies. Electrochemical measurements indicate that the CN-Me-bpy ligand is significantly easier to reduce than the unsubstituted bipyridine (on the order of ~500 mV), implying that the lowest energy (3)MLCT (metal-to-ligand charge transfer) state will be associated with the CN-Me-bpy ligand(s) in all three compounds. Comparison of the Huang-Rhys factors derived from spectral fitting analyses of the steady state emission spectra of complexes 1-3 suggests all three compounds are characterized by excited-state geometries that are less distorted relative to their ground states as compared to [Ru(bpy)(3)](PF(6))(2); the effect of the more nested ground- and excited-state potentials is reflected in the unusually high radiative quantum yields (13% (1), 27% (2), and 40% (3)) and long (3)MLCT-state room-temperature lifetimes (1.6 μs, 2.6 μs, and 3.5 μs, respectively) for these compounds. Coupling of the π* system into the CN groups is confirmed by nanosecond step-scan IR spectra which reveal a ~40 cm(-1) bathochromic shift of the CN stretching frequency, indicative of a weaker CN bond in the (3)MLCT excited state relative to the ground state. The fact that the shift is the same for complexes 1-3 is evidence that, in all three complexes, the long-lived excited state is localized on a single CN-Me-bpy ligand rather than being delocalized over multiple ligands. 相似文献
20.
The conversion of quadricyclane () to norbornadiene () is catalyzed by stannous chloride and stannous chloride-phosphine complexes. A newly synthesized polymer-bound phosphine-stannous chloride complex also proved effective in the catalytic conversion of to . 相似文献