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 15 毫秒
1.
Density functional theory calculations were carried out to investigate Cu12TM (TM = Co, Rh, Ir, Ni, Pd, Pt, Ag, Au) bimetallic metal catalysts for the mechanism of reverse water–gas shift (RWGS) reaction. The three possible reaction pathways relevant to the RWGS reaction are explored, including the CO2 dissociation, carboxyl, and formate mechanisms. Our results indicate that the RWGS reaction prefers to follow the CO2 dissociation mechanism on Cu12TM surfaces. A detailed potential energy diagram of the kinetically favored mechanism is presented that shows that the RDS of reaction are the formation of H2O and carboxyl (HOCO), formate (HCOO) dissociation, respectively. And, Cu12TM (TM = Co, Pt) are lower than other catalysts from the energy barrier of elementary step. Moreover, the catalytic behavior of a Cu12TM cluster is changed significantly due to the modifiers, via the electron transfer from TM to Cu-based cluster, and the activation barrier decreases with doped TM. The turnover frequency of the Cu12Co is the highest value, which thus is more efficiency catalyst to RWGS reaction. To gain insights into the synergistic effect in catalytic activity of the Cu12TM bimetallic cluster, a projected density of states analysis has been performed. Our works will be important for predicting the energetic trends and designing a better catalyst of RWGS reaction.  相似文献   

2.
The reaction of IrRu3(CO)13(μ-H), 1 with HSnPh3 in hexane solvent at reflux has provided the new mixed metal cluster compounds Ir2Ru2(CO)11(SnPh3)(μ-H)3, 2 and IrRu3(CO)11(SnPh3)3(μ-H)4, 3 containing SnPh3 ligands. Compound 2 which was obtained in low yield (3%) contains one SnPh3, two iridium atoms and two ruthenium atoms. The increase in the number of iridium atoms must have resulted from a metal–metal exchange process. The major product 3 (19% yield) contains an open cluster of one iridium and three ruthenium atoms with three SnPh3 ligands and four hydride ligands. Both compounds were characterized structurally by single crystal X-ray diffraction analysis.  相似文献   

3.
Three new platinum–ruthenium complexes: Pt3Ru3(PBut 3)3(CO)12, 8, Pt5Ru3(PBut 3)3(CO)12, 9 and PtRu3(PBut 3)2(CO)83-PBut)(μ-H)2, 10 were obtained from the reaction of Ru3(CO)12 with Pt(PBut 3)2. Compound 8 was obtained from this reaction when conducted at 25 °C. Compounds 9 and 10 were obtained when the reaction was conducted at 68 °C. The structure of 8 consists of a central triangular cluster of three ruthenium atoms with one Pt(PBut 3) group bridging each of the three Ru–Ru bonds. The structure of 9 consists of a capped pentagonal bipyramidal cluster of eight metal atoms that is formed formally by the addition of two platinum atoms to 8. The structure of 10 contains a triangular cluster of three ruthenium atoms with a Pt(PBut 3) group bridging one of the Ru–Ru bonds. A t-butyl phosphido ligand formed by degradation of a molecule of PBut 3 bridges the three ruthenium atoms. This report is dedicated to the memory of Professor F. A. Cotton for his many pioneering contributions to inorganic and metal cluster chemistry.  相似文献   

4.
A complex of Erbium perchloric acid coordinated with l-aspartic acid and imidazole, Er2(Asp)2(Im)8(ClO4)6·10H2O was synthesized for the first time. It was characterized by IR and elements analysis. The heat capacity and thermodynamic properties of the complex were studied with an adiabatic calorimeter (AC) from 80 to 390 K and differential scanning calorimetry (DSC) from 100 to 300 K. Glass transition and phase transition were discovered at 220.45 and 246.15 K, respectively. The glass transition was interpreted as a freezing-in phenomenon of the reorientational motion of ClO4− ions and the phase transition was attributed to the orientational order/disorder process of ClO4− ions. The thermodynamic functions [H T  − H 298.15] and [S T  − S 298.15] were derived in the temperature range from 80 to 390 K with temperature interval of 5 K. Thermal decomposition behavior of the complex in nitrogen atmosphere was studied by thermogravimetric (TG) analysis and differential scanning calorimetry (DSC).  相似文献   

5.
A solvatothermal reaction of the octahedral cluster molybdenum complex (H3O)2[Mo63-Cl)8Cl6] · 6H2O with CaCl2 · 6H2O and OPPh3 in acetonitrile gave the known polymeric complex trans-[{Ca(OPPh3)4}{Mo63-Cl)8Cl6}]. However, a closer examination revealed that this system also produces a novel cluster complex, [Ca(OPPh3)5][Mo63-Cl)8Cl6] · OPPh3, which was isolated and characterized by X-ray diffraction.  相似文献   

6.
Sublimation of europium pivalate binuclear complexes Eu2(Piv)6 and [Eu2(Piv)6 · (Phen)2] (Piv = (CH3)3CCOO, Phen = C12H8N2) in the temperature range of 383–660 K is studied by the Knudsen effusion method with mass-spectrometric analysis of the gas phase. The vaporization of Eu2(Piv)6 is shown to be accompanied by polymerization and the formation of Eu2(Piv)6 and Eu4(Piv)12 molecules. The saturated vapor over the mixed-ligand complex of europium pivalate with o-phenanthroline consists of Phen, Eu2(Piv)6, and Eu4(Piv)12 molecules. The partial pressures of the gas components, as well as the standard enthalpies of sublimation and dissociation of the reaction proceeding with removal of phenanthroline have been determined.  相似文献   

7.
Mo6Se8(Ph3P)6·2H2O cluster complex has been synthesized and its structure has been defined. The compound is triclinic, space group P1ˉ, with unit cell parameters a = 14.3356(5) Å, b = 15.7882(4) Å, c = 25.3949(8) Å, = 95.9750(10)°, β = 91.1030(10)°, γ= 112.2570(10)°, V = 5279.8(3) Å3, Z = 2, ρcalc = 1.772 g/cm3. The complex has a molecular structure. The molybdenum atoms of the {Mo6Se8} cluster nucleus are coordinated by the phosphorus atoms of triphenylphosphine molecules. Original Russian Text Copyright ? 2007 by Yu. V. Mironov, Zh. S. Kozhomuratova, D. Yu. Naumov, and V. E. Fedorov __________ Translated from Zhurnal Strukturnoi Khimii, Vol. 48, No. 2, pp. 389–393, March–April, 2007.  相似文献   

8.
A new coordination polymer, [Cd(NH3)4]2{Cd[Re3Mo3S8(CN)6]}·1.5H2O (I), was prepared by the reaction between solutions of Cd(CH3COO)2 · 2H2O in aqueous ammonia and CaK4[Re3Mo3S8(CN)6] · 8H2O in water. The crystals are cubic, space group Fm3m (Prussian blue structural type); a = 15.0268(4) Å (CIF file CSD no. 431555). According to ESR data, compound I is paramagnetic, g-factor is 2.298. Thermal stability investigation by TGA and powder X-ray diffraction showed that elimination of coordinated NH3 molecules is accompanied by sample amorphization.  相似文献   

9.
In the reaction of Na2Se with [Fe(CO)5] in isopropanol with subsequent acidification with HCl, which is used to synthesize [(μ-H)2Fe33-Se)(CO)9] (II), the cluster [(μ-H)2Fe53-Se)2(CO)14] (I) was detected. In assumption that compound I could serve as a suitable synthon for preparing the bulky heterometallic clusters, its reactions with the Rh-containing complexes were studied. The reaction of I with [Rh(CO)2Cp*] (Cp* is pentamethylcyclopentadienyl) was found to give a mixture of the products. The main reaction products were isolated and their structures were determined: [Fe2Rh(μ3-Se)2(CO)6Cp*], [Fe2Rh(μ3-Se)(μ3-CO)(CO)6Cp*], [FeRh23-Se)(μ-CO)(CO)3Cp 2 * ], [Fe2Rh24-Se)(μ-CO)4(CO)2Cp 2 * ]. Potassium hydride treatment of II with subsequent addition of [Cp*Rh(CH3CN)3](CF3SO3)2 leads to the well-known cluster complex [Fe3Rh(μ4-Se)(CO)9Cp*]. A set of the reaction products indicates that the {Fe5Se2} core cannot be used as one-piece “building block” in the synthesis of heterometallic clusters.  相似文献   

10.
Significant progress has been made in understanding the nature of the transition state andthe paths for electron transfer, especially the influences of environmental factors and themolecular properties on the electron transfer rate. These classical and semi-classical, aswell as quantum-mechanical theory, have been very successful in rationalizing severalstructure-reactivity relationships and in predicting novel features of reactivity. Thesemodels established some links between the electron tra…  相似文献   

11.
Systematic investigations on lowest energy CO adsorbed neutral and ionic Rhn (n = 2–8) clusters in the gas phase are performed with all electron relativistic method using density functional theory within the generalized gradient approximation. Geometrical and electronic parameters are evaluated to understand the bonding nature as well as the binding interaction of CO on stable neutral and ionic rhodium clusters. Anionic adducts exhibit higher adsorption energy along with smaller Rh–C and larger C–O bond distances in comparison to neutral and cationic RhnCO (n = 2–8) clusters. Synergic bond formation is noticed between rhodium and carbon atom of CO molecule due to back-donation of electron from metal d-orbitals to π* orbital of CO in the case of anionic and some neutral clusters. Angular and Mülliken charge analysis along with electron density distribution suggest that anionic rhodium clusters form strong bond with carbon atom of CO than the neutral and cationic clusters.  相似文献   

12.
13.
The reaction of the trinuclear oxo-centered mixed-valence complex [Mn3O(O2CPh)6(Py)2(H2O)] with 2,2′-bipyridyl (Bipy) and another potential tripodal ligand affords the title compound [Mn3(PhCO2)6(Bipy)2] · H2O in good yield. The X-ray crystallographic diffraction study reveals that three mangenese ions are arranged in a linear mode with Mncenter-Mnterminal and Mnterminal-Mnterminal diatances of 3.588 and 7.176 Å, respectively. Molar magnetic susceptibility of the compound gradually decreases from 12.23 (300 K) to 4.45 cm3 K mol?1 (2 K). Taking into account the structure of this compound, the data in the 2.0–300 K range were fit to the appropriate theoretical expression to give J = ?2.73 cm?1, ρ = 2.07%, N a = ?0.0004 cm3 mol?1, g = 1.992, and R 2 = 0.99996. The magnetization versus external magnetic field measurements at 2 K shows that the ground state is S T = 5/2.  相似文献   

14.
A new dichromium(III) cobalt(II) diphosphate(V) of the formula CoCr2(P2O7)2 was detected in the Co3Cr4(PO4)6–Cr(PO3)3 system. The new compound was obtained as a result of high-temperature solid-state reactions between CoCO3, Cr2O3 and (NH4)2HPO4 as well as between Cr(PO3)3 and Co3Cr4(PO4)6. CoCr2(P2O7)2 was characterized using XRD, DTA and IR methods. Results demonstrated that CoCr2(P2O7)2 crystallizes in the triclinic system and its unit cell parameters were calculated. Its infrared spectrum was presented. CoCr2(P2O7)2 melts incongruently at 1270±10 °C with a formation of solid α-CrPO4. The compound Co3Cr4(PO4)6, component of the system under study, was obtained for the first time as a pure phase. Its thermal stability was also investigated. Co3Cr4(PO4)6 is stable in air up to 1410 ± 20 °C.  相似文献   

15.
The structure of tris-complexes Y(MDA)3, Y(HFA)3 and their dimeric forms of Y2(MDA)6 and Y2(HFA)6 (MDA = C3O2H3; HFA = C5O2F6H) are investigated by the non-empirical Hartree-Fock method and also taking into account the correlation of electrons in the context of density functional theory (DFT/B3LYP) using effective pseudopotentials to describe the atomic core and double-exponential valence basis sets supplemented with polarization functions. For the first time the structure of Y(HFA)3, Y2(MDA)6, and Y2(HFA)6 molecules has been studied in this work. The equilibrium configuration of Y(HFA)3 monomer is a structure with a coordination polyhedron of [YO6] that has the form of a distorted octahedron with planar chelate fragments and CF3 substituents in which the C-F bond screens the C-C bond in the chelate fragment. The barrier of the internal rotation of one CF3 group is ~3.0 kJ/mole (DFT/B3LYP). It is shown that rotations of CF3 groups in Y(HFA)3 molecule could be considered as mutually independent. The structures of C 1 and C 2 symmetry are the equilibrium configurations of Y2(MDA)6 and Y2(HFA)6 dimers respectively (DFT/B3LYP). The structure of the coordination polyhedron of [Y2O12] in C 1 and C 2 symmetry configurations can be represented as a combination of eight-and seven-coordinated polyhedra or two seven-coordinated polyhedra respectively. The energy stability of dimeric complexes is studied. According to DFT/B3LYP data, the dimerization energy equals ?80.8 kJ/mole and ?62.3 kJ/mole for Y2(MDA)6 and Y2(HFA)6 respectively. The force fields and frequencies of normal vibrations are calculated for the equilibrium configurations. The trends of changes in the vibrational spectra in Y(MDA)3→Y(HFA)3 and monomer→dimer transitions are discussed. The comparison with available experimental data is made.  相似文献   

16.
The formation of Pd–Ag nanoparticles deposited from the heterobimetallic acetate complex PdAg2(OAc)4(HOAc)4 on α-Al2O3, γ-Al2O3, and MgAl2O4 has been investigated by high-resolution trans-mission electron microscopy, temperature-programmed reduction, and IR spectroscopy of adsorbed CO. The reduction of PdAg2(OAc)4(HOAc)4 supported on γ-Al2O3 and MgAl2O4 takes place in two steps (at 15–245 and 290–550°C) and yields Pd–Ag particles whose average size is 6–7 nm. The reduction of the Pd–Ag catalyst supported on α-Al2O3 occurs in a much narrower temperature range (15–200°C) and yields larger nanoparticles (~10–20 nm). The formation of Pd–Ag alloy nanoparticles in all of the samples is demonstrated by IR spectroscopy of adsorbed CO, which indicates a marked weakening of the absorption band of the bridged form of adsorbed carbon monoxide and a >30-cm–1 bathochromic shift of the linear adsorbed CO band. IR spectroscopic data for PdAg2/α-Al2O3 suggest that Pd in this sample occurs as isolated atoms on the surface of bimetallic nanoparticles, as is indicated by the almost complete absence of bridged adsorbed CO bands and by a significant weakening of the Pd–CO bond relative to the same bond in the bimetallic samples based on γ-Al2O3 and MgAl2O4 and in the monometallic reference sample Pd/γ-Al2O3.  相似文献   

17.
The structure of the salt Cs[Gd(H2O)4Re6Te8(CN)6]·4H2O (space group P-1, a = 9.436(5) Å, b = 12.365(7) Å, c = 15.187(8)Å, α = 89.104(10)°, β = 86.996(10)°, γ = 82.304(9)°) has been established by single crystal XRD. The structure of the compound features layers involving Gd3+ cations bound to cluster anions [Re6Te8(CN)6]4? through cyanide groups. The interlayer space contains cesium cations and crystallization water molecules.  相似文献   

18.
Incorporation of a 5d transition metal into the face-centered cubic metal-cyanide cluster geometry is accomplished for the first time with the isolation of a series of compounds featuring [(Me3tacn)8M8Pt6(CN)24]12+ (M = Cr, Mo) clusters. Reaction of [(Me3tacn)Cr(CN)3] and K2[PtCl4] in a boiling aqueous solution generates [(Me3tacn)8Cr8Pt6(CN)24]Cl12 · 27H2O (1), wherein PtII centers reside at the face-centering sites and the cyanide ligands have reoriented to give PtII–C≡N–CrIII linkages. The cyclic voltammogram obtained for a solution of 1 in DMSO exhibits a quasireversible reduction event centered at E 1/2 = ?1.59 V versus Cp2Fe0/1+. Reaction of 1 with K2[Pt(CN)4] in aqueous solution affords [(Me3tacn)8Cr8Pt6(CN)24][Pt(CN)4]6 · 6H2O (2), in which each face of the cubic cluster is capped by a staggered tetracyanoplatinate anion with a Pt–Pt separation of 3.1552(7) Å. Attempts to perform analogous cluster-forming reactions with [(Me3tacn)Mo(CN)3] revealed a tendency toward cluster decomposition to give mixtures of insoluble products, including [(Me3tacn)8Mo8Pt6(CN)24][Pt(CN)4]6 · 46H2O (3) and [(Me3tacn)8Mo8Pt6(CN)24][Pt(CN)4]2.5[Pt(CN)3Br]2Br3 · 6H2O (4). Crystallographic analyses revealed these compounds to contain the anticipated [(Me3tacn)8Mo8Pt6(CN)24]12+ cluster in fully- and partially-capped forms, respectively. Unfortunately, the insolubility of these molybdenum-containing products precluded characterization of the cluster by cyclic voltammetry.  相似文献   

19.
Single crystals of the Na4[Na2Cr2(C2O4)6] · 10H2O complex were synthesized for the first time. The structure of the complex was determined by X-ray diffraction analysis. The compound crystallizes in the monoclinic crystal system with the unit cell parameters a = 17.290(4) Å, b = 12.521(3) Å, c = 15.149(3) Å, β = 100.45(3)°, Z = 4, space group Cc. Anionic layers [NaCr(C2O4)3] 2n 4n? can be distinguished in the crystal structure of the complex. The Na+ cations and water molecules, involved in the formation of a hydrogen bond network, are located between the anionic layers.  相似文献   

20.
The reactions of [Ni16(C2)2(CO)23]4? and [Ni38C6(CO)42]6? with CuCl afforded mixtures of the previously reported [HNi42C8(CO)44(CuCl)]7? bimetallic octa-carbide cluster and the new [HNi43C8(CO)45]7? and [HNi44C8(CO)46]7? homo-metallic octa-carbides. The three species have very similar properties resulting always in co-crystals such as [NMe4]7[HNi42+2xC8(CO)44+2x(CuCl)1?x]·6.5MeCN (x = 0.14) (86% [HNi42C8(CO)44(CuCl)]7?, 14%[HNi43C8(CO)45]7?/[HNi44C8(CO)46]7?) and [NMe4]7[HNi42+2xC8(CO)44+2x(CuCl)1?x]·5.5MeCN (x = 0.30) (70% [HNi42C8(CO)44(CuCl)]7?, 30% [HNi43C8(CO)45]7?/[HNi44C8(CO)46]7?). The new homo-metallic octa-carbides can be obtained free from the Ni–Cu octa-carbido cluster by reacting [Ni10(C2)(CO)16]2? in thf with a stoichiometric amount of CuCl, and crystals of [NMe4]6[H2Ni43+xC8(CO)45+x]·6MeCN (x = 0.72), which contain [H2Ni44C8(CO)46]6? (72%) and [H2Ni43C8(CO)45]6? (28%), have been obtained. Despite the different charges and compositions, these anions display almost identical structures, which are also closely related to those previously reported for the bimetallic Ni–Cd octa-carbido clusters [Ni42+xC8(CO)44+x(CdCl)]7? and [HNi42+xC8(CO)44+x(CdBr)]6?. Indeed, all these clusters are based on the same Ni42C8 cage decorated by miscellaneous [CdX]+ (X = Cl, Br), [CuCl] and [Ni(CO)] fragments.  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