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 252 毫秒
1.
We present the results of a density functional calculation on adsorption of O2, CO, and their coadsorption at various sites of neutral, cationic, and anionic Pd4 clusters. For all the clusters, the dissociative adsorption of oxygen sitting on Pd bridge sites is found to be preferable. Both O2 and CO binding energies are found to be higher for the anionic Pd4 cluster followed by cationic and neutral cluster. However, binding energies of O2 or CO in the coadsorption complexes follow the trend: anionic > neutral > cationic. © 2010 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

2.
Cationic and anionic cobalt oxide clusters, generated by laser vaporization, were studied using guided-ion-beam mass spectrometry to obtain insight into their structure and reactivity with carbon monoxide. Anionic clusters having the stoichiometries Co2O3(-), Co2O5(-), Co3O5(-) and Co3O6(-) were found to exhibit dominant products corresponding to the transfer of a single oxygen atom to CO, indicating the formation of CO 2. Cationic clusters, in contrast, displayed products resulting from the adsorption of CO onto the cluster accompanied by the loss of either molecular O 2 or cobalt oxide units. In addition, collision induced dissociation experiments were conducted with N 2 and inert xenon gas for the anionic clusters, and xenon gas for the cationic clusters. It was found that cationic clusters fragment preferentially through the loss of molecular O 2 whereas anionic clusters tend to lose both atomic oxygen and cobalt oxide units. To further analyze how stoichiometry and ionic charge state influence the structure of cobalt oxide clusters and their reactivity with CO, first principles theoretical electronic structure studies within the density functional theory framework were performed. The calculations show that the enhanced reactivity of specific anionic cobalt oxides with CO is due to their relatively low atomic oxygen dissociation energy which makes the oxidation of CO energetically favorable. For cationic cobalt oxide clusters, in contrast, the oxygen dissociation energies are calculated to be even lower than for the anionic species. However, in the cationic clusters, oxygen is calculated to bind preferentially in a less activated molecular O 2 form. Furthermore, the CO adsorption energy is calculated to be larger for cationic clusters than for anionic species. Therefore, the experimentally observed displacement of weakly bound O 2 units through the exothermic adsorption of CO onto positively charged cobalt oxides is energetically favorable. Our joint experimental and theoretical findings indicate that positively charged sites in bulk-phase cobalt oxides may serve to bind CO to the catalyst surface and specific negatively charged sites provide the activated oxygen which leads to the formation of CO 2. These results provide molecular level insight into how size, stoichiometry, and ionic charge state influence the oxidation of CO in the presence of cobalt oxides, an important reaction for environmental pollution abatement.  相似文献   

3.
Small aluminum oxide cluster cations and anions, produced by laser vaporization, were investigated regarding their reactivity toward CO and N2O employing guided-ion-beam mass spectrometry. Clusters with the same stoichiometry as bulk alumina, Al2O3, exhibited atomic oxygen transfer products when reacted with CO, suggesting the formation of CO2. Anionic clusters were less reactive than cations but showed higher selectivity towards the transfer of only a single oxygen atom. Cationic clusters, in contrast, exhibited additional products corresponding to the sequential transfer of two oxygen atoms and the loss of an aluminum atom. To determine if these stoichiometric clusters could be generated from oxygen-deficient species, clusters having a stoichiometry with one less oxygen atom than bulk alumina, Al2O2, were reacted with N2O. Cationic clusters were found to be selectively oxidized to Al2O3(+), while anionic clusters added both one and two oxygen atoms forming Al2O3(-) and Al2O4(-). The oxygen-rich Al2O4(-) cluster exhibited comparable reactivity to Al2O3(-) when reacted with CO.  相似文献   

4.
Two 15-membered octaazamacrocyclic nickel(II) complexes are investigated by theoretical methods to shed light on their affinity forwards binding and reducing CO2. In the first complex 1[NiIIL]0, the octaazamacrocyclic ligand is grossly unsaturated (π-conjugated), while in the second 1[NiIILH]2+ one, the macrocycle is saturated with hydrogens. One and two-electron reductions are described using Mulliken population analysis, quantum theory of atoms in molecules, localized orbitals, and domain averaged fermi holes, including the characterization of the Ni-CCO2 bond and the oxidation state of the central Ni atom. It was found that in the [NiLH] complex, the central atom is reduced to Ni0 and/or NiI and is thus able to bind CO2 via a single σ bond. In addition, the two-electron reduced 3[NiL]2− species also shows an affinity forwards CO2.  相似文献   

5.
Three new copper(II), nickel(II) and cobalt (II) dinuclear complexes with a bis-amide ligand derived from tartaric acid have been prepared and characterized. For this purpose, the ligand (R,R)-(+)-di-N,N′-methylpyridino-tartramide (dmpt) was synthesized via the classical aminolysis of (R,R)-(+)-dimethyltartrate with pyridylmethylamine. The molecular structures of the complexes Na[Cu2(dmptH−3)(CO3)] · 8H2O (1) and [Ni2(dmptH−2)2] · 9.75H2O (2) were elucidated by X-ray diffraction, and the complex [Co2(dmptH−3)(μ-OH)] · NaClO4 · 5H2O (3) by XAS. The crystal structure of (1) shows that the two metallic centres are in a square planar environment. Each copper(II) is bound to pyridyl and deprotonated amidic nitrogen atoms and to the oxygen atoms of hydroxyl and carbonato groups. In complex (2), both nickel atoms are in a distorted octahedral environment with an identical set of donors atoms, N4O2, coming from four nitrogen atoms of two pyridylmethylamido moieties and two oxygen donor atoms of alcohol groups. XAS analysis of complex (3) allows us to propose a CoN2O4 chromophore, with two nitrogen atoms coming from pyridyl and amidic groups and two bridged oxygen atoms from a deprotonated alcohol group and an hydroxyl group; the hexacoordination is achieved by two water molecules. The spectroscopic, electrochemical and magnetic properties of these complexes were investigated.  相似文献   

6.
It is demonstrated by ESR measurements that O 2 (CO + O2) radical anions result from CO + O2 adsorption on the oxidized surface of CeO2. These radical anions are stabilized in the coordination sphere of Ce4+ cations located in isolated and associated anionic vacancies. This reaction shows an activation behavior determined by CO adsorption. The variation of O 2 (CO + O2) concentration with CO adsorption temperature suggests that surface carbonates and carboxylates participate in this reaction. In the (0.5– 10.0)%CeO2/ZrO2 system, O 2 forms on supported CeO2 and is stabilized on Ce4+ and Zr4+ cations. The stability of O 2 -Ce4+ complexes is lower on supported CeO2 than on unsupported CeO2, indicating a strong interaction between the cerium cations and the support.__________Translated from Kinetika i Kataliz, Vol. 46, No. 3, 2005, pp. 423–429.Original Russian Text Copyright © 2005 by Il’ichev, Kuli-zade, Korchak.  相似文献   

7.
The mechanism of the oxygen evolution on RuO2 and Ru0.9Ni0.1O2−δ anodes was studied in 0.1 M HClO4 using 18labeling combined with differential electrochemical mass spectrometry (DEMS). It was shown that the mechanism of the oxygen evolution is potential sensitive. At potentials negative to 1.12 V vs. SCE all the evolved oxygen originates from the electrolyte solution. At higher potentials an additional mechanism involving an exchange of the oxygen between electrolyte and electrocatalyst starts to apply. The extent of this oxygen exchange mechanism reflects the chemical composition of the electrocatalyst and is significantly higher at Ru0.9Ni0.1O2−δ electrodes.  相似文献   

8.
The structural and electronic properties of two series of Group VB transition metal oxide clusters, M4O n ? and M4O n (M = Nb, Ta; n = 8–11), are investigated using density functional theory calculations. Generalized Koopmans’ theorem is applied to predict the vertical detachment energies and simulate the photoelectron spectra. Large highest occupied molecular orbital–lowest unoccupied molecular orbital gaps are observed for these two stoichiometric M4O10 clusters and estimated to be 3.98 and 4.38 eV for M = Nb and Ta, respectively. The M4O 10 ?/0 (M = Nb, Ta) clusters are polyhedral cage structures with high symmetry (T d for the neutral and D 2d for the anion) in which each metal atom joints three bridging and one terminal O atoms. For the Nb oxide species, Nb4O 8 ?/0 and Nb4O 9 ?/0 can be viewed as removing two and one terminal O atoms from Nb4O 10 ?/0 , respectively. The Ta species follow the same rule to the Nb species, except that the anionic Ta4O8 ? is formed by removing one terminal and one bridging O atoms from Ta4O10 ?. The Ta4O9 containing a localized Ta3+ site can readily react with O2 to form the Ta4O11 which can also be viewed as replacing a terminal oxygen atom in Ta4O10 by a peroxo O2 unit, whereas the added oxygen atom is found to be a bridging one in the O-rich clusters Nb4O 11 ?/0 and the anionic Ta4O11 ?. Molecular orbital analyses are performed to analyze the chemical bonding in the tetra-nuclear metal oxide clusters and to elucidate their structural and electronic evolution.  相似文献   

9.
由摩尔比分别为1:2和1:8的NiCl2·6H2O和Na2B4O7·10H2O作为反应物, 合成两种非晶态镍硼酸盐, 同时通过水热法合成β-Ni(OH)2. 化学分析和热重-微商热重法(TG-DTG)分析结果确定两种非晶态镍硼酸盐的分子组成分别为NiO·0.8B2O3·4.5H2O和NiO·B2O3·3H2O. 激光拉曼(Raman)实验结果表明镍硼酸盐样品中主要存在的硼氧阴离子为B3O3(OH)52-和B2O(OH)62-. 同步辐射扩展X射线吸收精细结构(EXAFS)方法对样品进行结构解析, 通过数据拟合给出样品中Ni 原子周围近邻配位原子种类、配位数以及原子间距离. 用不同晶体结构作为标准对两种非晶态镍硼酸盐进行拟合的结果表明, 样品中Ni 原子周围局域结构与Ni3B2O6晶体(ICSD No.31387)中的吻合较好. Ni 原子周围配位原子为O、B和Ni, 对于NiO·0.8B2O3·4.5H2O, 配位数分别为5.7、3.8和3.8, 配位距离分别为0.208、0.263 和0.311 nm; 对于NiO·B2O3·3H2O, 配位数分别为6.0、4.0 和4.0, 配位距离分别为0.207、0.262和0.310 nm.  相似文献   

10.
Co-Mo-based catalysts supported on mixed oxide supports MgO-Al2O3 with different Mg/Al atom ratios for water gas shift reaction were studied by means of TPR, Raman, XPS and ESR. It was found that the octahedral Mo species in oxidized Co-Mo/MgO(x)-Al2O3 catalyst and the contents of Mo5+, Mo4+, S2− and S2−2 species in the functioning catalysts increased with increasing the Mg/Al atom ratio of the support under the studied experimental conditions. This is favorable for the formation of the active Co-Mo-S phase of the catalysts. Catalytic performance testing results showed that the catalysts Co-Mo/MgO-Al2O3 with the Mg/Al atom ratio of the support in the range of 0.475–0.525 exhibited optimal catalytic activity for the reaction.  相似文献   

11.
A fully encapsulated Pt 4 tetrahedron in an incomplete tetrahedron of 36 nickel atoms is present in [Ni36Pt4(CO)45]6− ( 1 ; see picture for the metal framework), which is obtained as an inseparable mixture with [Ni37Pt4(CO)46]6− ( 2 ) by reaction of [Ni6(CO)12]2− with K2[PtCl4]. The trimethylbenzylammonium salts of 1 and 2 cocrystallize in a 1:1 ratio. The additional Ni atom of 2 caps the truncated vertex of 1 .  相似文献   

12.
A three‐dimensional polymeric NiII complex, [Ni(bpp)(NIP)(H2O)]n (bpp = 1,3‐di(4‐pyridyl)propane and NIP = 5‐nitroisophthalate), has been synthesized and characterized. The coordination number of the nickel atom is six (NiN2O4) and the coordination environment around the NiII atom may be described as a distorted octahedron in which two nitrogen atoms of “bpp” ligand occupy the cis positions. The effective magnetic moment for this complex indicate that the interactions between two NiII atoms through the effective exchange media are antiferromagnetic. Self‐assembly of these compounds in the solid state via π–π‐stacking interactions is discussed.  相似文献   

13.
Studies on Oxide Catalysts. XXXIV. Redoxbehaviour of Nickel in Zeolite NiNaY. 1. Reducibility and Reoxidizability of Nickel in Zeolites NiNaY The properties of metallic nickel in reduced (470–870 K) and reoxidized (470, 670 K) samples were studied by chemical analysis (reaction with K2Cr2O7) and spectroscopic methods (FMR, IR after CO adsorption, UV/VIS). The reduction of Ni2+ cations from oxidic clusters proceeds in an onestep reaction. Contrary to this, isolated Ni2+ cations are reduced stepwise to Ni+ cations and subsequently to metallic nickel. The reduction degree depends in characteristic manner on the reduction temperature. Metallic nickel which was reduced at temperatures < 620 K, can be completely reoxidized at 470 K. Higher temperatures result in metallic aggregations which are not completely reoxidized even at 670 K.  相似文献   

14.
Density‐functional with generalized gradient approximation (GGA) for the exchange‐correlation potential has been used to calculate the energetically global‐minimum geometries and electronic states of NinAl (n = 2–8) neutral clusters. Our calculations predict the existence of a number of previously unknown isomers. All structures may be derived from a substitution of a Ni atom at marginal positions by an Al atom in the Nin+1 cluster. Aluminum atom remains on the surface of the geometrical configurations. Moreover, these species prefer to adopt three‐dimensional (3D) spacial forms at the smaller number of nickel atoms compared with the pure Nin+1 (n ≥ 3) configuration. Atomization energies per atom for NinAl (n = 2–8) have the same trend as the binding energies per atom for Nin (n = 3–9). The stabilization energies reveal that Ni5Al is the relatively most stable in this series. In comparison with the magnetic moment of pure metal nickel (0.6 μB), the average magnetic moment of Ni atom increases in Ni Al clusters except the Ni3Al. Moreover, except the case of Ni5Al, Ni average magnetic moment decreases when alloyed with Al atoms than that in pure Ni clusters, which originate the effective charge transferring from Al to Ni atoms. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

15.
The radiolysis of water without a protecting agent is found to form a low steady-state concentration of molecular hydrogen. The addition of bromide anion leads to a linear response for molecular hydrogen production with doses up to 300 kGy. Bromide concentrations are found to remain constant over this dose range due to recycling of the oxidized species containing the bromine atom by hydrated electrons, H atoms or HO2 (O2). This process appears to occur many times with little change in total bromide anion concentration. Efficient electron scavengers are found to have no effect on bromide anion concentration except possibly at extremely high concentrations. Nitrous oxide saturated solutions show a significant depletion of bromide anion concentration with a concurrent formation of BrO3 and a suppressed yield of molecular hydrogen.  相似文献   

16.
Infrared spectra of CO-treated platinum hydrosols subsequently treated with acetylene, hydrogen, and oxygen reveal that v(CO)ads decreases from 2070 cm−1 with increasing gas-treatment time. This has been attributed to a reduction in the coverage of adsorbed CO. In Pt sol/CO/C2H2 systems, v(CO)ads decreases to a limiting value of ca. 2060 cm−1 after exposure to acetylene. In the Pt sol/CO/H2 systems, v(CO)ads decreases to ca. 2050 cm−1 after exposure to hydrogen gas. The lower frequency in the Pt sol/CO/H2 system has been attributed to CO adsorption on more active metal sites formed from the reduction of surface platinum oxides. Exposure of the CO-treated platinum hydrosols to O2 gas was found to cause the eventual disappearance of the v(CO)ads band in infrared spectra, which was attributed to oxidation of adsorbed CO to CO2 by weakly bound surface layers of platinum oxides formed by the oxygen treatment.  相似文献   

17.
In this study, three novel tetranuclear nickel(II) cubane-type clusters with the general formula [Ni4(L)43-CH3O)4(CH3OH)4] [L: the anion of 5-methyl-2-hydroxybenzaldehyde (1), 2-hydroxypropiophenone (2), and 2-hydroxybenzophenone (3)] were synthesized and characterized by single-crystal X-ray diffraction analysis. The crystal structure of each compound contains a tetranuclear cubane core [Ni4O4] based on an approximately cubic array of altering nickel and oxygen atoms with intracluster metal–metal separations of 3.04–3.14 Å. Each Ni(II) atom is surrounded by two oxygen atoms from the ligand (L) and by the μ3-CH3O oxygen atom that bridges three Ni atoms of the cubane core. The coordination sphere of Ni is completed with one methanol molecule and making six-coordinate with a distorted octahedral geometry. These complexes were characterized also by spectroscopy (IR and UV–Vis). Simultaneous TG/DTG–DTA techniques were used to analyze their thermal behavior under inert atmosphere, with particular attention to determine their thermal degradation pathways, which was found to be a multi-step decomposition accompanied by the release of the ligand molecules. Finally, the kinetic analysis of the decomposition processes was performed for the first step of complex (3), since only this verifies the requirement of applying an isoconversional method like Kissinger–Akahira–Sunose (KAS). For this step, we found the average value E a = 107.8 ± 4.5 kJ mol?1.  相似文献   

18.
The effects of CO complexation on highly exothermic vanadium oxidation reactions is evaluated. We study the chemiluminescent (CL) reaction products formed when vanadium vapor entrained in Ar or CO is oxidized by O3 or NO2. The multiple collision V+Ar+O3→VO*(C 4Σ, 4Φ, 2X)+Ar+O2 reactive encounter yields two previously unreported VO excited states, whereas the V+Ar+NO2→VO*+Ar+NO reactive encounter populates states up to and including VO* C 4Σ. The multiple collision V+nCO+O3 reactive encounter would appear to form a VOCO excited state complex, emitting in the region 420–560 nm, via the formation and oxidation of V(CO)2 viz. V(CO)2+O3→VOCO*+CO+O2 and a relaxed VO excited state emitter via V+nCO+O3→VO*+nCO+O2 where the VO excited state excitation is mediated by V–CO complexation. In complement, the much less exothermic V–NO2 encounter displays an emission which, in concert with previous studies of CO complexation, suggests the formation of a VO(CO)2 excited state complex viz. V(CO)2+NO2→VO(CO)2*+NO. The experiments characterizing CL are complemented by comparative laser-induced fluorescence studies of the VO X 4Σ–CO and Ar interactions and their influence on the VO C 4Σ–X 4Σ laser-induced excitation spectrum. These studies, in conjunction with further attempts to excite LIF in the 420–560 nm region, suggest that the observed complex emissions result primarily from VO excited state interactions. Complementary time-of-flight mass spectroscopy of vanadium and vanadium-oxide–carbonyl complex formation demonstrates the formation of V(CO), V(CO)2, V2(CO), and VOCO, the latter three of which demonstrate clear metastable-ion dissociation peaks for the processes VOCO+→V++CO2, V(CO)2+→V++2CO, and V2(CO)+→V2++CO, suggesting that these vanadium complexes when formed in a reaction-based environment may be photodissociated with light in the visible and ultraviolet regions.  相似文献   

19.
The first stages of Co–Ni and Co–Ni–Mo deposition in sulphate–citrate medium at pH 4.0 were analysed. In both cases, the formation of non-hydrogenated nickel on the electrode before alloy deposition was detected by linear sweep voltammetry and inductively coupled plasma mass spectrometry. Co–Ni electrodeposition was anomalous since the Co/Ni ratio in the alloy was higher than the corresponding [Co(II)]/[Ni(II)] ratio in solution. The adsorption of Co(II) over the initial nickel could explain the anomalous codeposition, which persisted with the addition of molybdate to the Co–Ni bath. However, the formation of intermediate molybdenum oxides also took place. A mechanism has been proposed to describe the sequence of steps for Co–Ni–Mo electrodeposition. Under our conditions, the alloy is formed mainly from free Co2+ and Ni2+ cations, whereas molybdate is reduced firstly to molybdenum oxide from MoO4(H3Cit)2− and, secondly, NiCit catalyses the subsequent reduction to molybdenum metal of the intermediate [MoO2–NiCit]ads species.  相似文献   

20.
Two 2-terephthalate (tp) bridged complexes, [Cu2(tp)(pren)4](ClO4)2 (pren = 1,3-diaminopropane) (1) and [Ni2(tp)(pren)4(Him)2](ClO4)2 (Him = imidazole) (2), have been synthesized and characterized by X-ray single-crystal structural analysis. In the discrete dinuclear [Cu2(tp)(pren)4]2+ cation of complex (1), each CuII atom has a square-pyramidal geometry, being coordinated by four nitrogen atoms (avg. 2.031 Å) from two pren ligands at the basal plane and one oxygen atom [2.259(3) Å] from a bis-monodentate tp group at the axial position. In the discrete dinuclear [Ni2(tp)(pren)4(Him)2]2+ cation of complex (2), each NiII center is coordinated by five nitrogen atoms [Ni—N 2.069(3)–2.109(2) Å] from one Him group and two pren groups, and completed by one oxygen atom [Ni—O 2.138(3) Å] from a bis-monodentate tp group to furnish a distorted octahedron. Magnetic susceptibility studies show that the pair of metal atoms, although being separated by >11.5 Å, exhibit weak intramolecular antiferromagnetic interactions in complexes (1) (g = 2.07 and J = –3.4 cm–1) and (2) (g = 2.10 and J = –0.7 cm–1). The electrochemical behaviors of the complexes have also been studied by cyclic voltammogram processes.  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