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 78 毫秒
1.
Binary diffusion coefficients, D12, of the metal acetylacetonates, palladium(II) acetylacetonate and cobalt(III) acetylacetonate, were measured from 308.2 to 343.2 K over the pressure range from 9 to 40 MPa at infinite dilution in supercritical carbon dioxide using the chromatographic impulse response method. The effects of pressure, temperature, density, and viscosity on D12 values were examined. It was observed that the D12 values of palladium(II) acetylacetonate and cobalt(III) acetylacetonate were larger than those of lipids with similar molecular weights, such as arachidonic acid and monoolein, respectively. Furthermore, the measured D12 data of each metal acetylacetonate were well correlated by the hydrodynamic equation D12/T as a function of carbon dioxide viscosity.  相似文献   

2.
Methylaluminoxane prepared from trimethylaluminium on a surface of ice mainly can be described as [Al4O3(CH3)6]4. This is shown through analysis, phase separation experiments with diethyl ether and molecular weight determinations in benzene, 1, 4-dioxane, tetrahydrofuran and trimethylaluminium. It is discussed that the reason for forming this ball-like structure is the saturation of four coordinated Al4O3(CH3)6. This molecule has a cavern which contains a solvent molecule or a molecule of trimethylaluminium. The consequences for the formation of the catalytically active structure together with metallocene are discussed.  相似文献   

3.
A novel one‐pot synthesis of dimethyl(1,5‐cyclooctadiene)platinum(II), i.e. [(COD)Pt(CH3)2] (complex 1), was developed in 92% yield using platinum acetylacetonate, 1,5‐cyclooctadiene and trimethylaluminium. Complex 1 was fully characterized by 1H and 13C NMR, mass spectrometry, cell dimensions and elemental analysis. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

4.
Polymerization of methyl methacrylate was carried out by four initiating systems, namely, cobalt(II) or (III) acetylacetonate–tert-butyl hydroperoxide (t-Bu HPO) or dioxane hydroperoxide (DOX HPO). Dioxane hydroperoxide systems were much more effective for the polymerization of methyl methacrylate than tert-butyl hydroperoxide systems, and cobaltous acetylacetonate was more effective than cobaltic acetylacetonate in both hydroperoxides. The initiating activity order and activation energy for the polymerization were as follows: Co(acac)2–DOX HPO (Ea-9.3 kcal/mole) > Co (acac)3–DOX HPO (Ea = 12.4 kcal/mole) > Co(acac)2t-Bu HPO (Ea = 15.1 kcal/mole) > Co(acac)3t-Bu HPO (Ea-18.5 kcal/mole). The effects of conversion and hydroperoxide concentration on the degree of polymerization were also examined. The kinetic data on the decomposition of hydroperoxides catalyzed by cobalt salts gave a little information for the interpretation of polymerization process.  相似文献   

5.
Trace amount of chromium was first converted quantitatively to volatile chromium chelates such as chromium(III) acetylacetonate, Cr(acac)1, chromium(III) trifluoracetylacetonate, Cr(tfa)3, and chromium(III) hexafluoroacetylacetonate, Cr(hfa)3, and the chelate solution was then analyzed by gas chromatography. It was found that Cr(tfa)3 is the most suitable chemical form for this purpose. The method was profitably applied to the determination of chromium(III) incorporated in the six kinds of commerical products of potassium Chromate.  相似文献   

6.
The electron transfer kinetics of the reaction between the surfactant-cobalt(III) complex ions, cis-[Co(en)2(C12H25NH2)2]3+, cis-α-[Co(trien)(C12H25NH2)2]3+(en:ethylenediamine, trien:triethylenetetramine, C12H25NH2 : dodecylamine) by iron(II) in aqueous solution was studied at 298, 303, 308 K by spectrophotometry method under pseudo-first-order conditions using an excess of the reductant in self-micelles formed by the oxidant, cobalt(III) complex molecules, themselves. The rate constant of the electron transfer reaction depends on the initial concentration of the surfactant cobalt(III) complexes. ΔS# also varies with initial concentration of the surfactant cobalt(III) complexes. By assuming outer-sphere mechanism, the results have been explained based on the presence of aggregated structures containing cobalt(III) complexes at the surface of the self-micelles formed by the surfactant cobalt(III) complexes in the reaction medium. The rate constant of each complex increases with initial concentration of one of the reactants surfactant-cobalt(III) complex, which shows that self micelles formed by surfactant-cobalt(III) complex itself has much influence on these reactions. The electron transfer reaction of the surfactant-cobalt(III) complexes was also carried out in a medium of various concentrations of β-cyclodextrin. β-cyclodextrin retarded the rate of the reaction.  相似文献   

7.
A new apparatus based on the circulation method was developed to measure the solubility of metal complexes in supercritical carbon dioxide (scCO2) at a wide range of temperatures and pressures. A UV–vis spectrometer, which was connected to a small saturation cell through optical fibers, was used to determine solubility. The solubilities of cobalt(III) acetylacetonate (Co(acac)3) and chromium(III) acetylacetonate (Cr(acac)3) in scCO2 were measured to check the validity of both the apparatus and the method and to accumulate new solubility data. The solubility data for Cr(acac)3 obtained in this study were in good agreement with the data reported in the literature.The measured solubilities of Co(acac)3 and Cr(acac)3 were also correlated with the empirical equation including the three adjustable parameters, based on the equation proposed by Chrastil. The parameters were determined by fitting the equation to the experimental data for each metal complex and the calculated results closely replicated the experimental data.  相似文献   

8.
The solubilities of tris(2,2,6,6-tetramethyl-3,5-heptanedionate) cobalt(III) (Co(thd)3) and chromium(III) (Cr(thd)3) in supercritical carbon dioxide (scCO2) were measured at temperatures ranging from 313 to 343 K. The measurements were carried out using a circulation-type apparatus with a UV–vis spectrometer. The solubilities of both Co(thd)3 and Cr(thd)3 increased as both the density of scCO2 and the temperature increased, which has the same tendency as cobalt(III) acetylacetonate (Co(acac)3) and chromium(III) acetylacetonate (Cr(acac)3) had in our previous work. The solubilities of Cr(thd)3 were higher than that of Co(thd)3, and the solubilities of Co(thd)3 and Cr(thd)3 were about 50- and 70-fold higher than those of Co(acac)3 and Cr(acac)3, respectively. The measured solubilities of the metal complexes were correlated with the equation based on Chrastil's equation. The parameters were determined by correlating the experimental data for each metal complex, and the correlated results well reproduced the experimental data, especially Co(thd)3. Moreover, the charge density distributions on the molecular surface of CO2 and the metal complexes used in the measurement were estimated by the quantum chemical calculation and the COSMO-RS to clear the effect of the molecular structure of the metal complexes on the affinity for CO2.  相似文献   

9.
The redox reaction between cobalt(II) and gold(III) chloride in the presence of 1.10-phenanthroline or 2,2'-bipyridine was studied, and a titration of the cobalt(II) complex with a gold(III) chloride solution was developed. A 4-fold amount of 1,10-phenanthroline or 2,2'-bipyridine was necessary for rapid quantitative reaction; the permissible pH range was 1.5–5. The oxidation of the cobalt(II) complex proceeds rapidly at 40–50°C, and a direct potentiometric titration was possible. The following maximum errors were obtained: 3.3% for 0.2–1.0 mg Co, 2.0% for 1–5 mg Co, and 0.70% for 10–40 mg Co. The following ions did not interfere: Ni(II), Zn(II), Pb(II), Cd(II), Mn(II), Fe(II), Cr(III), Al(III), Th(IV), Se(IV), Ti(IV), U(VI), Mo(VI), SO2-4 and PO3-4. Even small quantities of silver(I), copper(II), palladium(II), mercury(II)and iron(III) interfered. The method was applied to the determination of high cobalt contents in high-temperature nickel-base alloys.  相似文献   

10.
A novel cyanate resin was prepared by the reaction of 10-(2,5-Dihydroxyphenyl)-10-H-9-oxa-10-phospha-phenantbrene-10-oxide (DOPO-HQ) and 2, 2-Bis (4-cyanatophenyl) propane (BACY) when trace of cobalt(III) acetylacetonate(CoAt(III)) was added, The curing behavior during the reaction of DOPO-HQ/BACY/CoAt(III) system was analyzed by gelation time (GT), differential scanning calorimetry (DSC) and Fourier transformation infrared spectrometry (FTIR), respectively. Compared with the BACY/CoAt(III) system (GT = 1239?s, Tp (exothermic peak temperature) = 215?°C), the GT of DOPO-HQ/BACY/CoAt(III) system is 110?s at 180?°C, and the Tp is 154?°C when containing 10?wt % DOPO-HQ. The limit oxygen index (LOI) and vertical burning test (UL-94) demonstrate that the flame retardancy of BACY resin is improved by DOPO-HQ. Specifically, the DOPO-HQ/BACY resin prepared by a mass ratio of DOPO-HQ: BACY = 10:90 has a LOI value of 33.1% and a UL-94 of V-0 rating, while the LOI value of BACY resin is 28.5% and UL-94 is no rating. In addition, the DOPO-HQ/BACY resin containing 10?wt % DOPO-HQ has excellent dielectric properties, the dielectric constant (Dk) is 2.69 and dielectric loss (Df) is 0.007. The as-synthesized DOPO-HQ/BACY resin shows promising application prospect as electronic packaging materials.  相似文献   

11.
In aqueous solution ruthenium trichloride reacted with picolinic acid (Hpic) in the presence of a base to afford [Ru(pic)3]. In solution it shows intense ligand-to-metal charge transfer transitions near 310 and 370 nm, together with a low-intensity absorption near 2000 nm. [Ru(pic)3] is one-electron paramagnetic and shows a rhombic ESR spectrum in 1:1 dimethylsulphoxide-methanol solution at 77 K. The distortions from octahedral symmetry have been calculated by ESR data analysis. The axial distortion is larger than the rhombic one. In acetonitrile solution it shows a reversible ruthenium(III)-ruthenium(II) reduction at −0.09 V vs. SCE and a reversible ruthenium(III)-ruthenium(IV) oxidation at 1.52 V vs. SCE. Chemical or electrochemical reduction of [RuIII(pic)3] gives [RuII(pic)3], which in solution shows intense MLCT transitions near 360, 410 and 490 nm, and is converted back to [Ru(pic)3] by exposure to air. Reaction of [Ru(pic)3] with 8-quinolinol (HQ) in dimethylsulphoxide solution affords [RuQ3]. [Ru(bpy)(pic)2] (bpy = 2,2′-bipyridine) has been prepared by the reaction of Hpic with [Ru(bpy)(acac)2]Cl (acac = acetylacetonate ion) in ethyleneglycol. It is diamagnetic and in solution shows intense MLCT transitions near 370, 410 and 530 nm. In acetonitrile solution it shows a reversible ruthenium(II)-ruthernium(III) oxidation at 0.44 V vs. SCE and a reversible one-electron reduction of bpy at − 1.64V vs. SCE.  相似文献   

12.
The exchange of methyl groups between tetramethyltitanium and trimethylaluminium has been studied by the direct observation of a new resonance line in the 2D NMR spectrum of the reaction mixture containing perdeuterated and undeuterated organometallic compounds.The kinetics and mechanism of the exchange reaction are influenced by the ethyl ether concentration in a hexane solution of the tetramethyltitanium compound. At a molar ratio TiMe4/Et20 > 1 rapid exchange occurs between unsolvated metal alkyls. Tetramethyltitanium etherate reacts with trimethylaluminium in two steps, an initial fast complexation to Me3Ti [AlMe4] which accounts for the transfer of a single methyl group from the titanium to the aluminium atom and a subsequent slow reverse process which leads to the transfer of methyl groups from the aluminium to the titanium atom and which results ultimately in a random equilibrium distribution of the labelled methyl groups amongst both metal atoms.Mixing the etherates of both titanium and aluminium methyl compounds led to the appearance of an extra signal downfield (τ 5.6 ppm) which is apparently due to the methyl resonance of the species [TiMe3]+ in the solvent-separated ion-pair form of the complex [TiMe3]+ [AlMe4]?.  相似文献   

13.
Reaction of tert.Butyl-lithium with Aluminiumtribromide; Molecular Structures of [HAl(CMe3)2]3 and [LiHAl(CMe3)3]2 The reaction of tert.butyl-lithium with aluminiumtrihalide at low temperatures was reinvestigated. Four compounds were isolated: trimeric di(tert.butyl)alane(III) 1 , monomeric (solution in benzene)/dimeric (solid state) lithium-tri(tert.butyl)alanate(III) 2 , tri(tert.butyl)alane 3 and lithium-tetra(tert.butyl)alanate 4 . X-ray structure analyses gave a planar sixmembered Al? H heterocycle for 1 and a Li? H-bridged dimer showing intramolecular interactions of lithium with C? H σ-bonds for 2 .  相似文献   

14.
The (alkyl)-bis(dimethylglyoximato)pyridinecobalt attached to polychloromethylstyrene by a cobalt–carbon bond was prepared by the reaction of Co(II)(DH)2Py with polychloromethylstyrene in benzene. The fraction of p-vinylbenzyl·Co(DH)2Py introduced to the polymer was 8.1 and 2.1 mole %. The photodecomposition of the polymer-bonded cobaloxime was investigated by following the change of the visible spectrum. The rate constant kdec of the polymer-bonded cobaloxime was 1.1 × 10?2 sec?1 in benzene; it is one-fourth of that of its monomeric analog, benzyl·Co(DH)2Py. The kdec values of the cobaloximes were also measured in benzene–dimethyl sulfoxide mixed solvents, and the polymer effects were discussed. The dependence of the photodecomposition on energy of the irradiation light was investigated, and it was found that the absorption band near 470 nm is important for the photodecomposition of the cobalt–carbon bond. Spectroscopic measurements of the ligand exchange reaction of polymer-bonded cobaloxime with pyridine in dimethyl sulfoxide gave a larger equilibrium constant (1.2 × 104 liter/mole) than that of benzyl·Co(DH)2Py (9.4 × 102 liter/mole). The kinetic data of the ligand exchange reaction indicated that the larger equilibrium constant for the polymeric system is due to the smaller rate constant of the reverse reaction. The thermodynamic parameters were also obtained.  相似文献   

15.
The surfactant–cobalt(III) complex, cis-[Co(trien)(4AMP)(DA)](ClO4)3, trien = triethylenetetramine, 4AMP = 4-aminopyridine, DA = dodecylamine was synthesized and characterized by various spectroscopic and physico-chemical techniques. The critical micelle concentration (CMC) value of this surfactant–cobalt(III) complex in aqueous solution was found out from conductance measurements. The conductivity data (at 303, 308, 313, 318 and 323 K) were used for the evaluation of the temperature-dependent CMC and the thermodynamics of micellization (ΔG m ° , ΔHm and ΔS m ° ). Also the kinetics of reduction of this surfactant–cobalt(III) complex by hexacyanoferrate(II) ion in micelles, β-cyclodextrin, ionic liquids (ILs) and in liposome vesicles (DPPC) media were studied at different temperature. The rate constant for the electron transfer reaction in micelles was found to increase with increase in the initial concentration of the surfactant–cobalt(III) complex. This peculiar behaviour of dependence of second-order rate constant on the initial concentration of one of the reactants has been attributed to the presence of various concentration of micelles under different initial concentration of the surfactant–cobalt(III) complex in the reaction medium. Inclusion of the long aliphatic chain of the surfactant complex ion into β-cyclodextrin leads to decrease in the rate constant. Below the phase transition temperature of DPPC, the rate decreased with increasing concentration of DPPC, while above the phase transition temperature the rate increased with increasing concentration of DPPC. It is concluded that below the phase transition temperature, there is an accumulation of surfactant–cobalt(III) complex at the interior of the vesicle membrane through hydrophobic effects, and above the phase transition temperature the surfactant–cobalt(III) complex is released from the interior to the exterior surface of the vesicle. In the presence of ionic liquid medium the second order rate constant for this electron transfer reaction for the same complex was found to increase with increasing concentration of ILs has also been studied. An outer-sphere mechanism is proposed for all these reactions and the results have been explained based on the hydrophobicity of the ligand and the reactants with opposite charges.  相似文献   

16.
The novel (E,E)-dioxime 5,6:13,14-dibenzo-9,10-benzo(15-crown-5)-2,3-bis(hydroxyimino)-7,12-dioxo-1,4,8,11-tetraazacyclotetradecane (H2L) has been synthesized by the reaction of 4′,5′-diaminobenzo(15-crown-5) with N,N′-bis(2-carbomethoxyphenyl)diaminoglyoxime (1). Only mononuclear CoIII and RuII complexes with a metal/ligand ratio of 1:2 have been isolated. The cobalt(III) complex bridged with BF2+ is achieved with H-bonded cobalt(III) complex and borontrifluoride ethyl ether complex. The reaction of BF2 bridged cobalt(III) complex with bis(benzonitril)palladium(II) chloride gives a trinuclear complex. The structures of dioxime and its complexes are proposed according to elemental analyses, 1H and 13C-NMR, IR and mass spectral data.  相似文献   

17.
The kinetics and mechanism of reduction of the surfactant complex ions, cis-chloro/bromo(dodecylamine)(triethylenetetramine)cobalt(III) by iron(II) in aqueous solution were studied at 303, 308 and 313 K by spectrophotometry under pseudo-first-order conditions using an excess of the reductant. The second-order rate constant increases with cobalt(III) concentration and the presence of aggregation of the complex itself alters the reaction rate. The reductions are acid-independent in the range [H+] = 0.05–0.25 mol dm−3. Variation of ionic strength (μ) influences the reaction rate. Activation and thermodynamic parameters have been computed. It is suggested that the reaction of Fe2+(aq) with the cobalt(III) complex proceeds by an inner-sphere mechanism. The critical micelle concentration (CMC) values of these surfactant metal complexes in aqueous solution were obtained from conductance measurements. Specific conductivity data (at 303, 308 and 313 K) served for the evaluation of the temperature-dependent CMC and the standard Gibbs energy of micellization (ΔGm0).  相似文献   

18.
The kinetics of the complex formation reaction of the β-hexamethyldiethylcorrole, [H3(β-Me6Et2)Cor], with zinc and cobalt acetates and with zinc acetylacetonate in DMF was studied. The rate of the complexation process between H3(β-Me6Et2)Cor and Zn(OAc)2 was shown to decrease in the following series of the solvents as DMF < AN < benzene and to suggest elevated chemical activity of the NH bonds in a molecule of corrole. The destruction of β-octaalkylcorrole in the basic solvents (DMF and DMSO) was discovered for the first time, and the kinetic parameters of this process were determined. With account for the structural distinctions between corroles H3Cor and porphyrins H2P, the probable mechanism of their reactions with metal salts was discussed and the transition state structure, which noticeably differ from the analogous reactions with porphyrins in the formation of the [H3Cor · Zn(OAc)2(Solv) n?3] intermediate.  相似文献   

19.
The synthesis, X-ray structure and properties of a pentanuclear cobalt(III) coordination cluster [{L(O2CCH3)Co2O(OCH3)2}2Co](ClO4)3 (1) (L? = 2,6-bis((3-aminopropylimino)methyl)-4-methylphenolate) are described. The dinucleating L? is coordinated with two cobalt(III) centers to form the {L(O2CCH3)Co2O(OCH3)2} unit, where each metal center is in a distorted octahedral N2O4 environment. The oxo and the methoxo ligands of these two dinuclear units assemble a distorted octahedral O6 coordination sphere around the central cobalt(III). Elemental analysis and spectroscopic (IR, NMR, UV–vis, and HRMS) features are consistent with the pentanuclear structure of the complex. The diamagnetic complex is a 1?:?3 electrolyte in solution. It is redox-active and displays a metal-centered reduction at E1/2 = ?0.04 V (vs. Ag/AgCl).  相似文献   

20.
An oligomer solution was prepared from methyltriethoxysilane with HNO3 catalyst. Aluninum acetylacetonate was introudced into the solution to eliminate crystalline precipitates (CH3SiO3/2)8. Self-sustained film was attempted by dropping the solution onto an aqueous solution, into which NaCl was added to elevate the solution density. The film formation greatly depended both on the reaction time and on the supporting solution. The structure of the film thus obtained was investigated by FTIR attenuated total reflectance spectroscopy.  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