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.
应用原子和表面簇合物相互作用的5-参数Morse势(简称5-MP)方法系统地研究了氧-铂台阶面体系.理论结果表明:在Pt(s)-[n(111)×(100)]型台阶面上,氧原子吸附在台阶下的四重位,对应稳定吸附态β2;平台上靠近四重位的三重吸附态被湮灭,其它三重位对应吸附态β1;而且平台的长度对四重吸附态有影响.  相似文献   

2.
The [Ni36Pt4(CO)45]6- and [Ni37Pt4(CO)46]6- clusters have been obtained in mixture upon reaction in acetonitrile of [Ni6(CO)12]2- salts with K2PtCl4 in a 2.5:1 molar ratio. The two hexaanions were indistinguishable by spectroscopic techniques. Crystallization of their trimethylbenzylammonium salts led to crystals of composition 0.5[NMe3CH2Ph]6[Ni36Pt4(CO)45]-0.5[NMe3CH2Ph]6[Ni37Pt4(CO)46]·C3H8O, hexagonal,space group P63 (No. 173), a=17.853(9), c=27.127(13) Å, Z=2; final R=0.057. The metal core of the [Ni36Pt4(CO)45]6- anion consists of a Pt4 tetrahedron fully encapsulated in a shell of 36 Ni atoms belonging to a very distorted and incomplete 5 tetrahedron. The [Ni37Pt4(CO)46]6- hexaanion derives from the former by capping the unique triangular face of the metal polyhedron with an additional Ni(CO) fragment. The [Ni36Pt4(CO)45]6--[Ni37Pt4(CO)46]6- mixture is rapidly degraded to the known [Ni9Pt3(CO)21]4- cluster by exposure to carbon monoxide. Its reaction with protic acids initially affords the corresponding [H6-nNi36Pt4(CO)45]n--[H6-nNi37Pt4(CO)46]n- (n=5, 4) derivatives, and eventually leads to rearrangement to the known [H6-n Ni38Pt6(CO)48]n- species. Both [Ni36Pt4(CO)45]6--[Ni37Pt4(CO)46]6- and [HNi36Pt4(CO)45]5--[HNi37Pt4(CO)46]5- mixtures have been chemically and electrochemically reduced to their corresponding [Ni36Pt4(CO)45]n--[Ni37Pt4(CO)46]n- (n=7–9) and [HNi36Pt4(CO)45]n--[HNi37Pt4(CO)46]n- (n=6–8) mixtures.  相似文献   

3.
Adsorption and decomposition of cyclohexanone (C(6)H(10)O) on Pt(111) and on two ordered Pt-Sn surface alloys, (2 × 2)-Sn/Pt(111) and (√3 × √3)R30°-Sn/Pt(111), formed by vapor deposition of Sn on the Pt(111) single crystal surface were studied with TPD, HREELS, AES, LEED, and DFT calculations with vibrational analyses. Saturation coverage of C(6)H(10)O was found to be 0.25 ML, independent of the Sn surface concentration. The Pt(111) surface was reactive toward cyclohexanone, with the adsorption in the monolayer being about 70% irreversible. C(6)H(10)O decomposed to yield CO, H(2)O, H(2), and CH(4). Some C-O bond breaking occurred, yielding H(2)O and leaving some carbon on the surface after TPD. HREELS data showed that cyclohexanone decomposition in the monolayer began by 200 K. Intermediates from cyclohexanone decomposition were also relatively unstable on Pt(111), since coadsorbed CO and H were formed below 250 K. Surface Sn allowed for some cyclohexanone to adsorb reversibly. C(6)H(10)O dissociated on the (2 × 2) surface to form CO and H(2)O at low coverages, and methane and H(2) in smaller amounts than on Pt(111). Adsorption of cyclohexanone on (√3 × √3)R30°-Sn/Pt(111) at 90 K was mostly reversible. DFT calculations suggest that C(6)H(10)O adsorbs on Pt(111) in two configurations: by bonding weakly through oxygen to an atop Pt site and more strongly through simultaneously oxygen and carbon of the carbonyl to a bridged Pt-Pt site. In contrast, on alloy surfaces, C(6)H(10)O bonds preferentially to Sn. The presence of Sn, furthermore, is predicted to make the formation of the strongly bound C(6)H(10)O species bonding through O and C, which is a likely decomposition precursor, thermodynamically unfavorable. Alloying with Sn, thus, is shown to moderate adsorptive and reactive activity of Pt(111).  相似文献   

4.
1 INTRODUCTION Diazoles have both a pyridine-type nitrogen and a pyrrole-type nitrogen (imino group), so they can act as bifunctional groups to coordinate with many metal ions or Lewis acids giving a variety of products[1, 2]. Recently researches on the pyrazole ligand have been focused on the following three sections. (i) The first is that the ligand acts as a bridged ligand and coordinates to metal forming di- and polynuclear complexes. Such system arouses much interest mainly because o…  相似文献   

5.
A complete NMR study involving both 1D and 2D 13C-{103Rh} and 31P-{103Rh} HMQC measurements, on [Rh6C(CO)14(PPh3)]2- are reported and discussed, together with the multiple Rh quantum effects found for resonances associated with edge- and face-bridging CO's. As found in [Rh6C(CO)15]2-, the carbonyl ligands in [Rh6C(CO)14(PPh3)]2- undergo CO-intermolecular exchange with 13CO at different rates; for the edge-bridging CO's, the lower the value of 1J(Rh–CO), the faster the rate of intermolecular exchange with 13CO.  相似文献   

6.
The title complex 1, C28H26CoN4O2, crystallizes in the triclinic system, space group P-1 with a = 9.506(3), b = 9.506(3), c = 14.837(5) A, α = 106.397(5), β = 106.012(5), γ =91.889(6)°, V = 1217.7(7) A3, Z = 2, Dc=1.390 g/cm3, Mr= 509.46, F(000) = 530,μ(MoKα) =0.738 mm-1, the final R = 0.0359 and wR = 0.0951 for 4275 observed reflections with I > 2σ(I).The complex adopts a distorted tetrahedral coordination sphere around the cobalt atom.  相似文献   

7.
8.
《Chemical physics letters》1986,125(2):134-138
H2S decomposition on the clean and (2 × 2)-S covered Pt(111) surfaces has been characterized using high-resolution electron energy loss (HREELS) and temperature-programmed desorption (TPD) spectroscopies. On the Pt(111)-(2 × 2)-S surface, a mixture of molecular H2S and sulfhydryl (SH) species forms following H2S adsorption at 110 K. The molecular H2S desorbs at 140 K leaving a (2 × 2)-S overlayer saturated with SH; the SH is stable up to 190 K. On the clean Pt(111) surface, a mixture of atomic sulfur, SH and chemisorbed molecular H2S is formed following H2S adsorption at 110 K. On the clean surface, adsorbed SH decomposes near 150 K. We report here the first definitive observation of an adsorbed sulfhydryl species on a metal surface.  相似文献   

9.
10.
Heating of an aqueous solution of [Pt(en)Py2Cl2]Cl2 · 2H2O (I) with KBr excess leads to the formation of [Pt(en)Py2Br2]Br2 · H2O (II). The interaction of a solution of II with bromine water results in the precipitation of polybromide ([Pt(en)Py2Br2]Br2 · Br2), which within a few days in the reaction solution partly transforms into oximide platinum(IV) complex, [Pt(HN-C(O)-C(O)-NH)Py2Br2] · H2O (III). Complex [Pt(en)PyBr3]Br · H2O (IV) with an impurity of II was prepared by reacting KBr excess and the product of [Pt(en)Py2]Cl2 oxidation with chlorine in 0.05 N HCl. The action of HNO3 on the solution of IV produced a nitrate derivative ([Pt(en)PyBr3]NO3, V). Complex IV, unlike II, does not react with bromine. The IR spectra of all the obtained compounds were recorded. Complexes II, III, and V were studied by X-ray crystallography. The crystals of II are monoclinic, space group P21/c, a = 15.640(2) Å, b = 9.345(1) Å, c = 14.167(2) Å, β = 102.63(1)°, V = 2020.5(5) Å3, Z = 4, R hkl = 0.033. The crystals of III are triclinic, space group P $\bar 1$ , a = 7.108(1) Å, b = 10.946(1) Å, c = 11.020(2) Å, α = 83.63(1)°, β = 80.31(1)°, γ = 75.02(1)°, V = 814.4(2) Å3, Z = 2, R hkl = 0.033. In the near-planar five-membered chelate ring (torsion angle NCCN is 7°), the C-O distances (1.23(1) Å) correspond to double bonds; the C-C (1.53(1) Å) and C-N (1.31(1) Å), distances correspond to ordinary bonds. The crystals of V are monoclinic, space group P21/c, a = 8.306(2) Å, b = 8.995(2) Å, c = 20.231(4) Å, β = 97.48(2)°, V = 1498.6(6) Å3, Z = 4, R hkl = 0.037.  相似文献   

11.
The surface reaction pathways of isoxazole and oxazole on Si(100)-2 × 1 surface were theoretically investigated. They both form a weakly bound Si-N dative bond adduct on Si(100)-2 × 1 surface. In the case of isoxazole, the barrierlessly formed Si-N adduct is the most important surface product, that cannot be easily converted into other species. On the other hand, a facile concerted [4+2](CC) cycloaddition without involving the initial Si-N dative bond adduct was also found in the case of oxazole adsorption. The existence of Diels-Alder reactions is attributed to the particular arrangement of the two heteroatoms of oxazole in such a way that the two Si-C σ-bonds can be formed in a [4+2] fashion. In short, the unique geometric arrangements and electronegativity of these similar heteroatomic molecules yielded distinctively different surface reaction characteristics.  相似文献   

12.
Nanoparticulate Cd(1-x)Zn(x)O (x = 0, 0.05-0.26, 1) is synthesized in a simple two-step synthesis approach. Vapor-diffusion induced catalytic hydrolysis of two molecular precursors at low temperature induces co-nucleation and polycondensation to produce bimetallic layered hydroxide salts (M = Cd, Zn) as precursor materials which are subsequently converted to Cd(1-x)Zn(x)O at 400 °C. Unlike ternary materials prepared by standard co-precipitation procedures, all products presented here containing < 30 mol% Zn(2+) ions are homogeneous in elemental composition on the micrometre scale. This measured compositional homogeneity within the samples, as determined by energy dispersive spectroscopy and inductively coupled plasma spectroscopy, is a testimony to the kinetic control achieved by employing slow hydrolysis conditions. In agreement with this observation, the optical properties of the materials obey Vegard's Law for a homogeneous solid solution of Cd(1-x)Zn(x)O, where x corresponds to the values determined by inductively coupled plasma analysis, even though powder X-ray diffraction shows phase separation into a cubic mixed metal oxide phase and a hexagonal ZnO phase at all doping levels.  相似文献   

13.
《Solid State Sciences》2000,2(1):143-148
A new hybrid material was synthesized by the microwave route from a mixture of Al2O3/HF/1,6 diaminohexane/EtOH. The structure of the hybrid fluoroaluminate, determined by single crystal X-ray diffraction, reveals a [H3N(CH2)6NH3]·AlF5 formulation and a monoclinic symmetry with the space group P21, with a=7.898(1) Å, b=5.514(1) Å, c=12.672(3) Å, β=103.69(2)°, V=536.2(2) Å3 and Z=2. The unit cell contains infinite inorganic chains of [AlF6] corner-sharing octahedra, linked each other by hexanediammonium cations.  相似文献   

14.
Infrared and Raman spectra are reported for (Me4N)Al2Cl7, (Et4N)Al2Cl7, KAl2Br7 and CsAl2I7 in the solid state at ordinary temperature. An assignment for practically all the vibrational modes is proposed for the three ions Al2C7, Al2Br7 and Al2I7. To elucidate the species occurring in the Friedel-Crafts liquids [the Ar:2AlX3: HX liquid mixture (X = Cl, Br)], the spectra of these solutions were also recorded. The quasi identity of the halogenide part of the spectra with the previous one leads to the unique formulation ArH+2Al2X7 for these solutions, in these cases the Al2X7 ion appears at a higher local symmetry of the AlX3 groups than in the solid state. An approximate valence force field calculation confirms the assignments. It also shows the transferability of force constants for a X-AlX3 model. Spectroscopic particularities are explained in terms of different coupling between bridged and terminal stretching modes.  相似文献   

15.
Cd(H2O)2+6–8 reacts with cis-(R, S)-[Pd(egta)]2– producing equimolar amounts of [Cd(egta)]2– and [Pd2(egta)Cl2]2–. The progress of the reaction and products have been followed by recording 1H- and 13C-n.m.r spectra as a function of time. The PdII released in forming [Cd(egta)]2– is thousands of times more reactive than CdII, and intercepts another [Pd(egta)]2– to form the 2:1 complex [Pd2(egta)Cl2]2–; the 2:1 complex is not attacked by CdII. The role of pendant carboxylates below the PdN2O2 plane of cis-(R, S)-[Pd(egta)]2– in supplying a site for docking of an incoming CdII or PdII centre, and in leading the metal near the lone pair of rupturing Pd–N bond of [Pd(egta)]2–, or simply by increasing the residence time of CdII or PdII nearby to accelerate the number of collisions between the ruptured N-base and external metal ions, is described. Although mixed-metal [Cd(Pd)(egta)] intermediates are required for the reaction, no such species achieves a detectably large enough concentration to be seen by 1H-n.m.r. The observed spectra are the sum of the reactant, [Pd(egta)]2–, and products, [Cd(egta)]2– and [Pd2(egta)Cl2]2–, throughout the time-dependent change.  相似文献   

16.
Second order rate constants and activation parameters H, S, and V have been measured for the oxidation of [Co(en)2(SOCH2CO2)]+ by S2O82– and by IO4– in highly aqueous H2O – t-BuOH mixtures. The changes in solvation on going from the initial to the transition state are discussed on the basis of the transfer functions Gto, Hto and Sto. Whereas Gt changes smoothly as the proportion of t-BuOH increases, the plots of Ht and TSt exhibit mirror behaviour and pass through extrema located around x2(t- BuOH)=0.038. Information on the role of solvation is complemented by the determination of activation volumes. These are discussed in terms of intrinsic and solvational contributions. It is proposed that changes in hydrophobic hydration are of principal importance in determining the response of H, S, and V to changes in solvent composition in H2O – t-BuOH mixtures.  相似文献   

17.
The sodium benzophenone ketyl-induced reaction of [Ru3(CO)12] with bis(diphenylphosphanyl)amine Ph2PN(H)PPh2 (dppa) in THF resulted in the formation of the expected metal cluster [Ru3(CO)10(μ-dppa)] ( 1 ) in high yield. 1 was fully characterized by spectroscopic means and crystals of the compound suitable for X-ray diffraction were obtained from dichloromethane/dioxane. The molecular structure of 1 as its dioxane solvate was determined by X-ray crystallography. The compound crystallized in a new crystal structure of [Ru3(CO)10(μ-dppa)] in the triclinic space group P1 , whereas that compound was described in an earlier report crystallizing from chloroform in the monoclinic space group P21/c.  相似文献   

18.
Noncovalentinteractionsplayadominantroleinmanyforefrontareasofmodemchemistry,frommat6rialsdesigntOmolecularbiology.Recelltly,theK -ninteractionshaveattractedmuchattentionbecauseoftheionseleCtivityinpotaSsiumchannelsl.Inothersystem,Sunneretal.showedthefirsttimebymassspectrometrythatK bindstobenzeneinthegasPhase:.ThisfindingstimulatedthestUdyoftheK -xinteractions.Indeed,theK -ninteractalshavealsobeen,identifiedbyX-raysinglecrystaldistractionanalysisforseveralcompouxtds:Km-be~usor--Phenyl',…  相似文献   

19.
The salt Rb[C6H3(COO)2()] · [C6H3(COOH)3] · 2H2O (I) of trimesic acid was synthesized and its thermal stability and conductivity (10–11 ohm–1 cm–1 at 298 K) were measured. Molecular and crystal structures of I were established by X-ray diffraction analysis. Hydrogen bonding system in complex I was detected by IR and Raman spectroscopies. X-ray diffraction data agree with vibration spectroscopy data.  相似文献   

20.
Differential capacitance curves in the (In-Ga)/[N-methylformamide + mc KCl + (1 ? m)c KClO4] and (In-Ga)/[N-methylformamide + mc KBr + (1 ? m)c KClO4] systems are measured using an ac bridge for the following molar portions m of the surface-active anion: 0, 0.01, 0.02, 0.05, 0.1, 0.2, 0.5, and 1. The Cl? and Br? anions specific adsorption in the systems can be described quantitatively by the Frumkin isotherm. The principal parameters of Cl? and Br? anions adsorption at the (In-Ga)/N-methylformamide interface are determined by different methods. Unlike Ga/N-methylformamide interface, where the adsorption energy increased in the sequence I? ≈ Br? < Cl?, at the (In-Ga)/N-methylformamide interface it increased in the reverse sequence: Cl? < Br? < I?. The adsorption parameters at the charge density q = 0, obtained by three different methods, are close to each other. However, the parameters α1 and α2, which characterize the charge effect on the adsorption energy, when determined by the analyzing of dependences of adsorption potential drop E ads on ln(mc), differ from those determined by two other methods. The error may be caused by the assuming that the adsorption potential drop is proportional to the coverage of dense layer with the specifically adsorbed ions.  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