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 31 毫秒
1.
This report summarizes our work on UV-laser induced desorption of small molecules and atoms from transition metal oxides. The systems presented serve as examples for a simple photochemical reaction, the fission of the molecule surface bond. State resolved detection methods were used to record the final state distributions of the desorbing neutral molecules. Detailed results on the systems NO/NiO(1 1 1) and CO/Cr2O3 (0 0 0 1) are presented. The experiments include investigations on stereodynamic aspects like the angular distributions of the desorbing molecules and, in the case of CO desorption, the rotational alignment with respect to the surface normal. Large desorption cross sections of (6 ± 1) · 10–17 cm2 for NO and (3.5 ± 1) · 10–17 cm2 for CO have been found for the desorption at 6.4 eV. The wavelength dependence indicates that the primary excitation step is substrate induced. The final state distributions show a high degree of translational, rotational and vibrational excitation and are clearly nonthermal of origin. The results are consistent with the formation of a negative ion intermediate state of the adsorbate. This observation is supported from a comparison to former results on NO/NiO(1 0 0) for which extensive ab initio calculations including electronically excited states exist. A spin state dependence of the vibrational excitation of NO could only be observed for NO/NiO(1 1 1) and is absent for NO/NiO(1 0 0). We attribute this observation to a spin state dependent coupling of the desorbing molecule to the surface in case the spin lattice orientation of the surface shows a preferential orientation. In the (111) plane the spin orientation is parallel within neighbour nickel ions while it is alternating in the (1 0 0) plane. For both systems studied the velocity component parallel to the surface is constant leading to a strong peaking along the surface normal for the fast molecules. The change from a preferred helicopter rotation (angular momentum vector aligned parallel to the surface normal) to a cartwheel motion (angular momentum vector aligned perpendicular to the surface normal) with increasing rotational excitation for desorption of the flat lying CO is consistent with a change of bonding geometry during the desorption process.  相似文献   

2.
The excitation mechanism in the CO-NiO(100) system induced by a uv-laser pulse has been investigated from first principles. For the laser-driven process, the relevant electronically excited states are identified, and it is shown that a transition within the CO molecule is the crucial excitation step rather than substrate mediated processes. A new mechanism is proposed, in which the formation of a genuine C-Ni bond in the excited state is the driving force for photodesorption rather than electrostatic interactions, as has been found in similar systems. This results in very high velocities of CO molecules desorbing from the NiO(100) surface after electronic relaxation.  相似文献   

3.
The adsorption of ethylene has been studied on stoichiometric NiO(100) and on surfaces reduced to 40% of the stoichiometric oxygen content. The adsorption process was followed with XPS, Auger spectroscopy and LEED at substrate temperatures of 200 to 500 K and at ethylene pressure of 5 × 10?7 Torr. At 200 K, two distinct ethylene species are observed on stoichiometric NiO(100). The first species saturates at 0.02 ML after 200 L and is adsorbed molecularly, interacting with both nickel and oxygen sites. A condensed species then forms which does not saturate for exposures up to 2100 L. Both adsorb reversibly with all traces of carbon absent by 270 K. At 200–300 K, reduced NiO(100) also adsorbs two molecular ethylene species, although with a preference for nickel sites. However, the uptake of ethylene increases only slightly with surface reduction. Adsorption is no longer reversible for the reduced surface and increasing the substrate temperature causes fragmentation of the adsorbed ethylene with a concomitant reduction in lattice oxygen content.  相似文献   

4.
We have studied electronic excitations at the surfaces of NiO (100), Cr2O3(111), and Al2O3(111) thin films with Electron Energy Loss Spectroscopy (EELS). On NiO (100) we observe surface electronic excitations in the energy range of the band gap which shift upon adsorption of NO. Ab initio cluster calculations show that these excitations occur within the Ni ions at the oxide surface. The (111) surface of Cr2O3 is characterized by distinct excitations which are also strongly influenced by the interaction with adsorbates. Temperature-dependent measurements show that two different states of the surface exist which are separated by an activation energy of about 10 meV. For Al2O3(111) we present data for a CO adsorbate. The oxide is quite inert with respect to CO adsorption as indicated by desorption temperatures between 38 K and 67 K. Due to the weak interaction with the substrate the a3II valence excitation of CO shows a clearly detectable vibrational splitting which has not been observed previously for a CO adsorbate in the (sub)monolayer coverage range. For several different adsorption state the lifetimes of the a3II state could be estimated from the halfwidths of the loss peaks, yielding values between 10–15 s for the most strongly bound species and 10–14 s for the CO multilayer.  相似文献   

5.
The formation and thermal stability of NiO on Ni(100) have been investigated using high-resolution electron energy loss spectroscopy (EELS) and low-energy electron diffraction (LEED). Our results indicate that the saturated NiO/Ni(100) layer prepared at 300 K is rather poorly ordered and is thermally unstable at higher temperatures. Heating this NiO/Ni(100) layer to 800 K produces a surface with mixtures of crystalline NiO(100) clusters and c(2 × 2)−O chemisorbed local structures. The long range order of the NiO(100) clusters could be improved by repeated cycles of oxygen adsorption at 300 K followed by heating to 800 K. The NiO(100) clusters obtained after 9 cycles of such dosing-annealing exhibit bulk-like properties, as suggested both by the off-specular EELS measurements and by the experimental observation that the intensities of the multiple loss features follow the expected Poisson distribution. The Ni---O bond strength of the NiO(100) clusters, estimated from the overtone spectra, is 3.6 eV. In addition, the reduction of NiO(100) clusters by H2 at 800 K has also been investigated. The NiO(100) clusters are reduced preferentially with respect to the c(2 × 2)−O overlayer, resulting in a reduction sequence of NiO(100) → c(2 × 2)−O → p(2 × 2)−O → Ni(100).  相似文献   

6.
《Applied Surface Science》1986,26(3):335-356
A surface study is made of CO adsorption and the Fischer-Tropsch (CO hydrogenation) reaction on Co foils with K precoverages of from 0 to 0.8 monolayers (ML, 1 ML = saturation K coverage ≈ 5 × 1014 K/cm2). X-ray photoemission (XPS) and Auger spectroscopy (AES) are used to characterize the adsorbed surface species. The adsorption and reaction occur in an atmospheric-pressure microreactor coupled to the surface analysis vacuum chamber by a rapid, valveless sample-transfer mechanism. After CO exposure (548 K, 100 kPa, 10 s) core levels are seen corresponding to molecular and dissociated CO. Molecular CO 1s levels exhibit a K-dependent binding energy decrease attributed to a surface-to-molecule charge transfer enhanced by the alkali. Potassium enhances the amount of CO adsorbed and thermally stabilizes the adsorbed molecule. Surface analysis after Fischer-Tropsch reaction (523–548 K, 100 kPa total reactant pressure) characterizes the amount and nature of the C deposited. The results are compared with earlier measurements on Fe surfaces. For identical reaction conditions, clean Co shows much less deposited C than clean Fe. It also forms surface carbidic C deposits much less readily than Fe, the C deposited being predominantly graphitic. Potassium predosing of the surfaces enhances the amount of C deposited, attributable to its enhancement of the CO adsorption step. It also changes the nature of the C, inducing carbidic C. When the carbidic C is induced (by a K precoverage of ∼ 0.7 ML), the amount of C deposited by reaction actually decreases, compared to that produced by reactions on surfaces with less K, an effect not previously observed during K-promoted reactions on Fe or Ni. The reactivity of carbidic C to hydrogen on both Co and Fe is reduced by K, helping to explain the lower methanation rate that has been observed after K promotion of Fischer-Tropsch catalysts.  相似文献   

7.
Adsorption of CO on Ni(100) has been investigated using secondary ion mass spectrometry (SIMS) and Auger electron spectroscopy at 175 and 295 K. Interaction with polycrystalline nickel was examined at 295, 325 and 365 K. All the secondary ions, Ni+, Ni2+, NiCO+ and Ni2CO+ show large increases in intensity as CO is adsorbed but there is no simple correlation of the secondary ion species with the sequence of linear and bridge-bonded CO species expected from electron energy loss spectroscopy. Adsorption of CO at 175 K on a hydrogen saturated Ni(100) surface, which is thought to permit only bridge-bonded adsorbed CO, does not result in any enhancement of Ni2CO+. The extent of increases in secondary ion yields after CO adsorption on the nickel surfaces are primarily related to the variations in the heat of adsorption as a function of surface coverage. The presence of more weakly-held species is important in enhancing secondary ion yields.  相似文献   

8.
The adsorption of Xe and CO on Au(100) has been studied by LEED, Auger electron spectroscopy, electron energy loss spectroscopy (EELS) and surface potential measurements. The physical adsorption of xenon showed successive stages preceding the completion of a monolayer. The heat of adsorption was 22 (±2) kJ mol?1 and the maximum surface potential was 0.45 V. Carbon monoxide gave a surface potential of 0.85 V at the highest coverage reached. The heat of adsorption showed a continuous fall from an initial value of 58 (±3) kJ mol?1 as the coverage increased. Ordered adsorption structures were not observed in LEED for either Xe or CO. The EEL spectrum of clean Au(100) agreed well with spectra of polycrystalline gold. New loss features observed with adsorbed Xe and CO are discussed.  相似文献   

9.
Carbon monoxide adsorption has been studied on a series of presulfided Ni(100) surfaces using vibrational spectroscopy. The sulfided Ni(100) surfaces were characterized using Auger electron spectroscopy and low energy electron diffraction, binding states were isolated by heating CO-dosed surfaces to prescribed temperatures, corresponding to the desorption temperatures of the CO. Adsorption of CO on Ni(100) with a p(2 × 2) array of sulfur lead to CO stretching frequencies of 1740 and 1930 cm?1 corresponding to desorption temperatures of 370 and 290 K, respectively. Adsorption of CO into the c(2 × 2)S structure resulted in a CO stretching frequency of 2115 cm?1 and a desorption peak near 140 K. The binding sites on the p(2 × 2)S structure were interpreted as metal four-fold hollows and bridging sites. The high frequency state was interpreted as weak bonding into the four-fold hollow with back donation into the π1 orbital on CO restricted by stearic hindrance due to adsorbed sulfur. Both the thermal desorption and vibrational results indicated that local CO-sulfur interactions are dominant on the presulfided Ni(100) surface in the coverage range studied.  相似文献   

10.
The DOS structures of NiO (0 0 1;1 1 1) surfaces and CO adsorption on these surfaces have been studied with spin-unrestricted and periodic DFT (B3LYP) methods. On the basis of the analysis of orbital interaction on DOSs, the bonding properties of surface atomic orbitals have also been interpreted. It is found that CO adsorption on (0 0 1) and (1 1 1) surfaces have different mechanisms and adsorption energies. A four-electron σ orbital interaction is produced when CO is adsorbed on NiO (1 1 1), CO adsorbption on NiO (1 1 1) surface is obviously stronger than that on surface (0 0 1). It is easy for the clean NiO (1 1 1) surface to reconstruct to (2 × 2) structure, but the surface covered by CO does not undergo such a reconstruction.  相似文献   

11.
The reduction of single crystal NiO(100) under hydrogen has been followed by AES, XPS and LEED for the pressure range of 1.0 × 10?7 to 1.3 × 10?6 Torr and for substrate temperatures of 150–350°C. The kinetics of reduction are controlled both by the rate of removal of lattice oxide at the surface and by the diffusion of subsurface oxygen to the oxygen-depleted surface. The rate of oxygen removal is first-order in surface oxide concentration and in hydrogen pressure. An induction period precedes the reduction reaction and its length is postulated to be controlled by surface defect concentration. The stoichiometric and reduced lattice oxygen species appear to be chemically identical and give a single symmetric XPS peak at 529.4 eV. Nickel spectra indicate a shift in XPS binding energies from those expected of the oxide to those of nickel metal early in the reduction process, although LEED indicates the NiO(100) surface lattice to remain the stable structure for surface reduced to approximately 20% of the stoichiometric oxygen concentration. Ni(100) island formation is observed, with Ni 〈010〉 and 〈001〉 directions along the NiO 〈010〉 and 〈001〉, respectively, but only after the NiO surface is severely depleted in oxygen.  相似文献   

12.
采用密度泛函理论(DFT)研究了CO分子在Pu (100)面上的吸附. 计算结果表明:CO在Pu (100)表面的C端吸附比O端吸附更为有利,属于强化学吸附. CO吸附态的稳定性为穴位倾斜>穴位垂直>桥位>顶位. CO分子与表面Pu原子的相互作用主要源于CO分子的杂化轨道和Pu原子的杂化轨道的贡献. 穴位倾斜吸附的CO分子的离解能垒较小(0.280eV),表明在较低温度下,CO分子在Pu (100)表面会发生离解吸附,离解的C,O原子将占据能量最低的穴位.  相似文献   

13.
CO在Pu(100)表面吸附的研究   总被引:1,自引:0,他引:1       下载免费PDF全文
采用密度泛函理论(DFT)研究了CO分子在Pu (100)面上的吸附. 计算结果表明:CO在Pu (100)表面的C端吸附比O端吸附更为有利,属于强化学吸附. CO吸附态的稳定性为穴位倾斜>穴位垂直>桥位>顶位. CO分子与表面Pu原子的相互作用主要源于CO分子的杂化轨道和Pu原子的杂化轨道的贡献. 穴位倾斜吸附的CO分子的离解能垒较小(0.280eV),表明在较低温度下,CO分子在Pu (100)表面会发生离解吸附,离解的C,O原子将占据能量最低的穴位. 关键词: 密度泛函理论 Pu (100) CO 分子和离解吸附  相似文献   

14.
Several kinds of NiO-based nanostructured films were prepared by pulsed-laser deposition (PLD) and sol–gel method, and CO sensing properties (1%, balanced by N2) of these films were studied. The sensitivity, defined as a difference of optical transmittance by gas atmospheric change (T=T(CO)-T(air)), increased with increasing NiO content for the sol–gel prepared films, and increased with the film thickness for the laser deposited NiO films. Sol–gel films exhibited shorter response time than NiO films prepared by PLD under low Ar pressure of 6.7×10-2 Pa indicating a better gas permeability. A shorter response time was also obtained upon raising argon pressure from 6.7×10-2 Pa to 8.0 Pa during laser ablation due to the morphological change. Covering a NiO film even with a very thin (0.8 nm) layer of SiO2 by sputtering drastically reduced the CO sensitivity. The multilayered NiO/SiO2 films were substantially less sensitive to the CO gas than NiO films due to the same reason. Sensing mechanism of the NiO films is due to catalytic CO oxidation that reduces the concentration of adsorbed O2 species and results in optical transmittance increase upon change in the environment from air to CO. PACS 81.15.Fg; 81.20.Fw; 83.85.Gk  相似文献   

15.
For CO adsorption on Fe(100) different adsorption species are detected with high resolution EELS (electron energy loss spectroscopy) which sequentially fill in with increasing coverage. Up to ~ 350 K and low CO exposure (≦1 L), a predominant molecular species with an unusually low stretching frequency, 1180–1245 cm?1, is detected. This unusual CO bond weakening is consistent with a “lying down” binding configuration of CO. For higher CO coverages at 110 K, further CO adsorption states with vibrational frequencies of 1900–2055 cm?1 are populated which are due to CO bound with the molecular axis perpendicular to the surface.  相似文献   

16.
The chemisorption of CO on the (100) surface of Ni has been studied using an Ni14 cluster and generalized valence bond (GVB) methods. CO is found to bond perpendicular to the Ni surface with optimized NiC and CO bond distances of 1.94 and 1.15 Å, respectively. The calculated NiCO bond strength is 29.7 kcal (experimental values 30–32 kcal). Vibrational frequencies are calculated to be 401 cm?1 for NiC stretch, 327 cm?1 for NiCO bend, and 2129 cm?1 for CO stretch. This decrease of the CO frequency by 71 cm?1 from the free molecule value is consistent with experiment based on self-consistent calculations of the positive ion states. We propose a new explanation for the loss of one PES peak upon chemisorption.  相似文献   

17.
The adsorption of carbon monoxide on the LaB6(1 0 0) and LaB6(1 1 1) surfaces was studied experimentally with the techniques of reflection absorption infrared spectroscopy and X-ray photoelectron spectroscopy. The interaction of CO with the two surfaces was also studied with density functional theory. Both surfaces adsorb CO molecularly at low temperatures but in markedly different forms. On the LaB6(1 1 1) surface CO initially adsorbs at 90 K in a form that yields a CO stretching mode at 1502-1512 cm−1. With gentle annealing to 120 K, the CO switches to a bonding environment characterized by multiple CO stretch values from 1980 to 2080 cm−1, assigned to one, two, or three CO molecules terminally bonded to the B atoms of a triangular B3 unit at the (1 1 1) surface. In contrast, on the LaB6(1 0 0) surface only a single CO stretch is observed at 2094 cm−1, which is assigned to an atop CO molecule bonded to a La atom. The maximum intensity of the CO stretch vibration on the (1 0 0) surface is higher than on the (1 1 1) surface by a factor of 5. This difference is related to the different orientations of the CO molecules on the two surfaces and to reduced screening of the CO dynamic dipole moment on the (1 0 0) surface, where the bonding occurs further from the surface plane. On LaB6(1 0 0), XPS measurements indicate that CO dissociates on the surface at temperatures above 400 K.  相似文献   

18.
CO oxidation reactivity of bare and TiO2-coated nanoparticles consisting of both NiO and Ni(OH)2 surfaces was studied. For the deposition of TiO2, atomic layer deposition was used, and formation of three-dimensional domains of TiO2 on NiO-Ni(OH)2 could be identified. Based on the data of X-ray Photoelectron Spectroscopy, we suggest that upon TiO2 deposition only Ni(OH)2 was remained on the surface, whereas NiO surface disappeared. Both CO adsorption and CO oxidation took place on NiO-Ni(OH)2 surfaces under our experimental conditions. CO adsorption was almost completely suppressed after TiO2 deposition, whereas CO oxidation activity was maintained to large extent. It is proposed that bare NiO cannot be active for CO oxidation, and can only uptake CO under our experimental condition, whereas hydroxylated surface of NiO can be active for CO oxidation.  相似文献   

19.
Chemisorption of CO on the Ni(100)p(2 × 2)O and c(2 × 2)O surfaces has been investigated by high-resolution electron energy loss spectroscopy (EELS) and low-energy electron diffraction (LEED). At 175 K CO adsorption on Ni(100)p(2 × 2)O saturates at about 1 L exposure in a structure interpreted to be Ni(100)p(2 × 2)O—p(2 × 2)CO. The CO layer is stable at 175 K but desorbs readily around 300 K. The EEL spectrum for p(2 × 2)CO shows vibrational losses at 46 meV and 245 meV interpreted to be due to excitations of the Ni-C and C-O stretching vibrations of CO molecules bridge bonded to two nearest neighbour Ni atoms. This interpretation is also supported by the LEED observations. For the preceeding dilute CO layer the vibrational loss spectrum reveals CO adsorption both to Ni bridge sites and hollow sites. At 175 K CO does only adsorb stationary on p(2 × 2)O defects in the Ni(100)c(2 × 2)O surface and not at all on epitaxially grown NiO(111) and (100) surfaces.  相似文献   

20.
Reflection absorption infrared spectroscopy (RAIRS) and high resolution electron energy loss spectroscopy (HREELS) have been used to study the adsorption of oxygen on the (100) and (111) surfaces of lanthanum hexaboride. Exposure of the surface at temperatures of 95 K and above to O2 produces atomic oxygen on the surface and yields vibrational peaks in good agreement with those observed in previous HREELS studies. On the La-terminated (100) surface, RAIRS peaks correspond to vibrations of the boron lattice that gain intensity due to a decrease in screening of surface dipoles that accompanies oxygen adsorption. A sharp peak at ~ 734 cm?1 in the HREEL spectrum shows isotopic splitting with RAIRS into two components at 717 and 740 cm?1 with full widths at half maxima of only 12 cm?1. The sharpness of this mode is consistent with its interpretation as a surface phonon that is well separated from both the bulk phonons and other surface phonons of LaB6. On the boron-terminated LaB6(111) surface, broad and weak features are assigned to both vibrations of the boron lattice and of boron oxide. On the (100) surface, oxygen blocks the adsorption sites for CO, and adsorbed CO prevents the dissociative adsorption of O2.  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