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 812 毫秒
1.
CpMoCo2(CO)8CCO2CHMe2 in propionic anhydride reacts with HPF6 to yield the mixed metal acylium cation CpMoCo2(CO)8CCO+ which reacts with a variety of nucleophiles, e.g., ROH, R2NH, yielding cluster-bound esters and amides, respectively. The cation also reacts with PhNMe2 in a Friedel-Crafts process. In the presence of water, Co3(CO)9CCO+ can also decompose to yield the neutral dimer [(Co3(CO)9C]2.  相似文献   

2.
The formation of a new compound, the most characteristic IR absorption bands of which appear at 2007 cm-1 and 1956 cm-1, has been in the reaction between Co2(CO)8 and Rh4(CO)12 under carbon monoxide pressure in a hydrocarbon medium. The same compound is also formed either by the reaction of Co2(CO)8 with [Rh(CO)2Cl]2 or by the reaction of Co3Rh(CO)12 with carbon monoxide. The new complex has not been isolated in a pure state, but the formula CoRh(CO)7 is proposed on the basis of the stoichiometry of its formation and its physico-chemical properties. Equilibrium constants and thermo-dynamic parameters for the reaction 2 Co2(CO)8 + Rh4(CO)12  4 CoRh(CO)7 have been estimated. Possible structures for the new complex are discussed on the basis of its IR spectrum.  相似文献   

3.
The dehydrotropylium–Co2(CO)6 ion was generated by the action of HBF4 or BF3 ? OEt2 on the corresponding cycloheptadienynol complex, which in turn has been prepared in four steps from a known diacetoxycycloheptenyne complex. The reaction of the cycloheptadienynol complex via the dehydrotropylium–Co2(CO)6 ion with several nucleophiles results in substitution reactions with reactive nucleophiles (N>1) under normal conditions, and a radical dimerisation reaction in the presence of less reactive nucleophiles. Competitive reactions of the cycloheptadienynol complex with an acyclic trienynol complex show no preference for generation of the dehydrotropylium–Co2(CO)6 ion over an acyclic cation. DFT studies on the dehydrotropylium–Co2(CO)6 ion, specifically evaluation of its harmonic oscillator model of aromaticity (HOMA) value (+0.95), its homodesmotic‐reaction‐based stabilisation energy (≈2.8 kcal mol?1) and its NICS(1) value (?2.9), taken together with the experimental studies suggest that the dehydrotropylium–Co2(CO)6 ion is weakly aromatic.  相似文献   

4.
The reaction of equimolar amounts of [Co(CO)3(NO)] and [PPN]CN, PPN+ = (PPh3)2N+, in THF at room temperature resulted in ligand substitution of a carbonyl towards the cyanido ligand presumably affording the complex salt PPN[Co(CO)2(NO)(CN)] as a reactive intermediate species which could not be isolated. Applying the synthetic protocol using the nitrosyl carbonyl in excess, the title reaction afforded unexpectedly the novel complex salt PPN[Co2(μ-CN)(CO)4(NO)2] ( 1 ) in high yield. Because of many disorder phenomena in crystals of 1 the corresponding NBu4+ salt of 1 has been prepared and the molecular structure of the dinuclear metal core in NnBu4[Co2(μ-CN)(CO)4(NO)2] ( 2 ) was determined by X-ray crystal diffraction in a more satisfactory manner. In contrast to the former result, the reaction of [PPN]SCN with [Co(CO)3(NO)] yielded the mononuclear complex salt PPN[Co(CO)2(NO)(SCN-κN)] ( 3 ) in good yield whose molecular structure in the solid was even determined and its composition additionally confirmed by spectroscopic means.  相似文献   

5.
Two new linked alkyne-bridging tetrahedral carbonyl clusters containing Co2C2 Co2(CO)6(μ-HCCCH2OOC(CH2)3COOCH2CCH-μ)Co2(CO)6, 1, and Co2(CO)6(μ-HCCCH2OOC(CH2)8COOCH2CCH-μ)Co2(CO)6, 2, have been prepared by reactions of two dipropargyl esters (HC≡CCH2OOC)2R (R = (CH2)3, (CH2)8) with Co2(CO)8. Expansion reactions of 1 and Co2(CO)6(μ-HCCCH2OOCCOOCH2CCH-μ)Co2(CO)6, 3, with Fe3(CO)12 give two new mixed-metal linked clusters Co2(CO)6(μ-HCCCH2OOC(CH2)3COOCH2CCH-μ,η4)Co2Fe2(CO)12, 4, and Co2(CO)6(μ-HCCCH2OOCCOOCH2CCH-μ,η4)Co2Fe2(CO)12, 5. The new clusters were characterized by elemental analysis, IR, 1H-NMR and ESI-MS analysis.  相似文献   

6.
The reaction of Re2(CO)8(μ-C6H5)(μ-H), 1 with corannulene (C20H10) yielded the product Re2(CO)8(μ-H)(μ-η2-1,2-C20H9), 2 (65 % yield) containing a Re2 metalated corannulene ligand formed by loss of benzene from 1 and the activation of one of the CH bonds of the nonplanar corannulene molecule by an oxidative-addition to 1 . The corannulenyl ligand has adopted a bridging η2-σ+π coordination to the Re2(CO)8 grouping. Compound 2 reacts with a second equivalent of 1 to yield three isomeric doubly metalated corannulene products: Re2(CO)8(μ-H)(μ-η2-1,2-μ-η2-10,11-C20H8)Re2(CO)8(μ-H), 3 (35 % yield), Re2(CO)8(μ-H)(μ-η2-2,1-μ-η2-10,11-C20H8)Re2(CO)8(μ-H), 4 (12 % yield), and Re2(CO)8(μ-H)(μ-η2-1,2-μ-η2-11,10-C20H8)Re2(CO)8(μ-H), 5 (12 % yield), by a second CH activation on a second rim double bond on the corannulene molecule. The isomers differ by the relative orientations of the coordinated Re2(CO)8(μ-H) groupings. All new products were characterized structurally by single crystal X-ray diffraction analysis.  相似文献   

7.
An ESR study has been made of the high nuclearity paramagnetic metal cluster anions [Rh12(CO)132-CO)10(C)2]3-, [Co13(CO)122-CO)12(C)2]4- and [Co6(CO)82-CO)6C]-. The assignment of the HOMO is based on a mixed valence model which relates the g tensor components of cluster systems to those of an appropriate conventional paramagnetic center. With this model the HOMOs of [Rh12(CO)132-CO)10(C)2]3- and of [Co13(CO)122-CO)12(C)2]4- are found to be mainly comprised of metal dz2 atomic orbitals, while for [Co6(CO)82-CO)6C]- a large overlap between d atomic orbitals and ligand orbitals is suggested. The occupation of the valence molecular orbitals deduced from the ESR data is consistent with the variations in MM bond distance observed by X-ray analysis.  相似文献   

8.
ZnO/Co3O4 porous nanocomposites were successfully fabricated by the thermal decomposition of Prussian Blue analogue (PBA) Zn3[Co(CN)6]2 nanospheres obtained at room temperature. Interestingly, ZnO/Co3O4 porous nanocomposites exhibit room‐temperature ferromagnetism. Moreover, the ZnO/Co3O4 porous nanocomposites show good catalytic activity for CO oxidation, and the CO conversion rate reaches 100 % at 250 °C. It is suggested that the synergistic effect of each component, relative high surface area (32 m2 g?1) and porous structure lead to the promising catalytic properties.  相似文献   

9.
The equilibrium between Co2(CO)8 and Co4(CO)12 has been investigated in hexane solution in the temperature range 105–145°C under carbon monoxide pressure (6–14 bar). the data obtained by infrared analytical monitoring of the system in a high-pressure cell alow a reasonably precise extension of the calculated equilibrium concentration between —20 and 300°C. The thermo- dynamic parameters obtained for this system are: ΔH0 29.5 ± 0.5 kcal mol-1, ΔS0 135 ± 3 cal mol-1 K-1. The stability regions of Co2(CO)8 and Co4(CO)12 in terms of p(CO) and temperature are discussed.  相似文献   

10.
The reaction mechanism of CO oxidation on the Co3O4 (110) and Co3O4 (111) surfaces is investigated by means of spin‐polarized density functional theory (DFT) within the GGA+U framework. Adsorption situation and complete reaction cycles for CO oxidation are clarified. The results indicate that 1) the U value can affect the calculated energetic result significantly, not only the absolute adsorption energy but also the trend in adsorption energy; 2) CO can directly react with surface lattice oxygen atoms (O2f/O3f) to form CO2 via the Mars–van Krevelen reaction mechanism on both (110)‐B and (111)‐B; 3) pre‐adsorbed molecular O2 can enhance CO oxidation through the channel in which it directly reacts with molecular CO to form CO2 [O2(a)+CO(g)→CO2(g)+O(a)] on (110)‐A/(111)‐A; 4) CO oxidation is a structure‐sensitive reaction, and the activation energy of CO oxidation follows the order of Co3O4 (111)‐A(0.78 eV)>Co3O4 (111)‐B (0.68 eV)>Co3O4 (110)‐A (0.51 eV)>Co3O4 (110)‐B (0.41 eV), that is, the (110) surface shows higher reactivity for CO oxidation than the (111) surface; 5) in addition to the O2f, it was also found that Co3+ is more active than Co2+, so both O2f and Co3+ control the catalytic activity of CO oxidation on Co3O4, as opposed to a previous DFT study which concluded that either Co3+ or O2f is the active site.  相似文献   

11.
The decomposition of Co2(CO)8 in hydrocarbon solutions, to Co4(CO)12 and further to metallic Co and carbon monoxide under an inert atmosphere, has been extensively studied and is well documented in the literature. We report here a study of the solid-state decomposition of Co2(CO)8 in a polymeric matrix under various conditions. Co2(CO)8 was incorporated into polystyrene by film casting from CH2Cl2 solutions under CO at room temperature. These films were decomposed at 90°C under N2. The decomposition rate of the Co2(CO)8 in polystyrene films was two orders of magnitude slower than for hydrocarbon solution decompositions. Conversely, the oxidation of Co2(CO)8 to CoO or Co2O3 in air at room temperature was observed to be two orders of magnitude faster in the polystyrene matrix as compared with solutions. When protective layers of polystyrene were cast on both sides of the Co2(CO)8-polystyrene films, oxidation became slower as a function of the thickness of the protective polymer layers. These observations, supported by infrared spectra, TEM micrographs, and electron diffraction patterns, are discussed to compare the solution and solid-state chemistry of Co2(CO)8.  相似文献   

12.
The reaction of Co(CO)3DMPP? with Co2(CO)6(DMPP)2 (DMPP = dimethylphenylphosphine) yields CoCO4? and Co2(CO)5(DMPP)3. The DMPP ligand of Co(CO)3DMPP? can be replaced by CO or triphenylphosphite.  相似文献   

13.
对于羰基混合金属簇的合成,利用配体的交换反应,制备含有不同配体的羰基混合金属簇。配体取代后的羰基金属簇的性质发生了变化,如可逆氧化还原性质[1],催化活性与选择性[2]等。由于过渡金属原子的性质各不相同,配位取代反应也有很大差异,所以研究配体取代反应,制备含有不同配体的羰基过渡金属簇成为金属簇化学的重要组成部分。  相似文献   

14.
The room-temperature reaction of NaCo(CO)4 with halogermanes, or Co2(CO)8 with GeH4, gives GeCo4(CO)14 which is assigned a Ge[Co2(CO)7]2 structure on infrared evidence. This new species eliminates one CO at 50°C to give (CO)4CoGeCO3(CO)9 and adds further CO(CO)4- to give anionic [GeCo6(CO)n]2-.  相似文献   

15.
Reaction of [(CpV)2(B2H6)2], 1 (Cp = η5-C5H5) with four equivalents of [Co2(CO)8] or [Co4(CO)12] in hexane at 70 °C leads to the isolation of the tetranuclear carbonyl cluster, [(η6-C6H5OCo)Co3(CO)9], 2 in modest yield. The geometry of 2 is similar to that of [Co4(CO)12] where all the four Co atoms are arranged in a tetrahedral geometry. The apical cobalt atom in 2 is coordinated to C6H5O ring in a η6-fashion and the other three cobalt atoms are each coordinated to three carbonyl ligands. Compound 2 has been characterized in solution by IR, 1H, 13C NMR and mass spectrometry and the structural types were unequivocally established by crystallographic analysis.  相似文献   

16.
GeMe2H2 reacts under mild conditions with [{Co2(CO)7}2Ge] to replace one bridging CO and give [Co4(CO)13Ge(GeMe2)]. GeH4 similarly yields a trace of [Co6(CO)20Ge2], which may be made in high yield from [Co2(CO)8] and Ge2H6 or Me2Si(GeH3)2. Spectroscopic evidence suggests structures of linked GeCo2 triangles.  相似文献   

17.
The bimetallic NiSn2 complex Ni(SnBu3t)2(CO)3, 1, was obtained from the reaction of Ni(COD)2 and Bu3tSnH and CO. The reaction of Co2(CO)8 and Bu3tSnH afforded the bimetallic Co–Sn complex Co(SnBu3t)(CO)4, 3. Compound 3 was also obtained from the reaction of Co4(CO)12 and Bu3tSnH but in a lower yield. Both compounds 1 and 3 were characterized by single crystal X-ray diffraction, and possess trigonal bipyramidal geometries around the transition metal centre with two and one stannyl ligands, respectively.  相似文献   

18.
Syntheses of Co2Rh2(CO)8(PF3)4 and Co2Rh2(CO)10(PF3)2 are described and their structures are discussed. Evidence is presented for an intermolecular ligand exchange between several tetranuclear cluster complexes. 19F and 31P NMR and mass spectroscopic data are presented and discussed. The complexes Rh4(CO)4(PF3)8 and Co2Ir2(CO)8(PF3)4 have been identified by their mass spectra.  相似文献   

19.
The reaction of Cr(CO)3(NH3)3 with diphenylacetylene affords as a main product the complex with Cr(CO)3 moiety bound to a phenyl ring of diphenylacetylene; Cr(CO)36-PhC2Ph) (I). Complex I readily reacts with Co2(CO)8 yielding the mixed metal complex Cr(CO)362-PhC2Ph)Co2(CO)6 (II). The reaction proceeds with retention of the Cr(CO)36-arene) structural unit, the Co2(CO)6 fragment being bound to the triple bond of diphenylacetylene in μ22-mode. The structure of II was determined by single crystal X-ray analysis. The complex crystallizes in space group P21/c with unit cell parameters a 8.666(3) Å, b 18.046(3) Å, c 15.155(6) Å. β 97.57(3)°, V 2349(2) Å3, Z = 4, Dx = 1.70 g/cm3. The structure was solved by direct methods and refined by full-matrix least-squares technique to R and Rw values of 0.032 and 0.034, respectively, for 3655 observed reflections. The data obtained show that two structural units in II, Cr(CO)36-Ph-) and Co2(CO)622-CC), are distorted due to steric repulsion between these metal carbonyl moieties. The Cr(CO)3 fragment is shifted from the centre of the phenyl ring and slightly tilted with respect to the phenyl ring plane. The Co2C2 tetrahedron in the Co2(CO)622-CC) moiety is distorted in such a way that two of the four CoiCj bonds are elongated.  相似文献   

20.
Co2(CO)8 catalyzes the ring‐opening copolymerization of propylene oxide with CO to afford the polyester in the presence of various amine cocatalysts. The 1H and 13C{1H} NMR spectra of the polyester, obtained by the Co2(CO)8–3‐hydroxypyridine catalyst, show the following structure ? [CH2? CH(CH3)? O? CO]n? . The Co2(CO)8–phenol catalyst gives the polyester, which contains the partial structural unit formed through the ring‐opening copolymerization of tetrahydrofuran with CO. The bidentate amines, such as bipyridine and N,N,N′,N′‐tetramethylethylenediamine, enhance the Co complex‐catalyzed copolymerization, which produces the polyester with a regulated structure. Acylcobalt complexes, (RCO)Co(CO)n (R = Me or CH2Ph), prepared in situ, do not catalyze the copolymerization even in the presence of pyridine. This suggests that the chain growth involves the intermolecular nucleophilic addition of the OH group of the intermediate complex to the acyl–cobalt bond, forming an ester bond rather than the insertion of propylene oxide into the acyl–cobalt bond. Co2(CO)8? Ru3(CO)12 mixtures also bring about the copolymerization of propylene oxide with CO. The molar ratio of Ru to Co affects the yield, molecular weight, and structure of the produced copolymer. The catalysis is ascribed to the Ru? Co mixed‐metal cluster formed in the reaction mixture. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 4530–4537, 2002  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