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.
The adsorption of SO2 on CaO (100) at 300 K has been studied using X-ray photoelectron spectroscopy. Under ultrahigh-vacuum conditions, the surface was exposed to 0–500 Langmuirs of SO2. The resulting adsorption yields a single SO surface species with an S 2p peak at 168.2 eV and an O 1ssol12 peak at 531.7 eV. Subsequent heating of the exposed surface to 673 K indicated no desorption or changes in the binding energies of the S 2p and O 1s12 peaks. On the basis of these data and binding-energy data for standard compounds, the adsorbed species is identified as SO42?. The surface coverage due to the SO42? species was also measured as a function of SO2 exposure. From these data, the initial adsorption is found to be first-order in surface coverage, and the initial sticking probability is found to have a value of 0.4.  相似文献   

2.
Surfaces of mineral cuprite prepared by fracture under UHV have been characterised by synchrotron XPS and near-edge X-ray absorption spectroscopy before and after exposure to ambient air. Before exposure of the cuprite, the Cu 2p photoelectron and Cu L2,3-edge absorption spectra were consistent with CuI with very little d9 character. Surface-enhanced O 1s spectra from the unexposed mineral revealed a surface species, with binding energy 0.95 ± 0.05 eV below the principal cuprous oxide peak, assigned to under-coordinated oxygen. A second surface species, with binding energy about 1 eV higher than the principal peak, was assigned to either hydroxyl derived from chemisorbed water vapour or surface oxygen dimers produced by restructuring of the cuprite fracture surface. The width of the principal O 1s peak was 0.66 ± 0.02 eV. The observed Cu L3- and O K-edge absorption spectra were in good agreement with those simulated for the cuprite structure. After exposure of the fracture surface to ambient air, the low binding energy O 1s surface species was barely discernible, the original high binding energy O 1s surface species remained of comparable intensity, new intensity appeared at an even higher (∼1.9 eV) binding energy, and the Cu L2,3-edge spectrum indicated the presence of CuII, consistent with the formation of a thin surface layer of Cu(OH)2.  相似文献   

3.
Using first-principles calculations, we systematically study the adsorption behavior of a single molecular H2O on the Be(0001) surface. We find that the favored molecular adsorption site is the top site with an adsorption energy of about 0.3 eV, together with the detailed electronic structure analysis, suggesting a weak binding strength of the H2O/Be(0001) surface. The adsorption interaction is mainly contributed by the overlapping between the s and pz states of the top-layer Be atom and the molecular orbitals 1b1 and 3a1 of H2O. The activation energy for H2O diffusion on the surface is about 0.3 eV. Meanwhile, our study indicates that no dissociation state exists for the H2O/Be(0001) surface.  相似文献   

4.
The electronic structure of (111) surface of β-crystobalite is investigat ed using the empirical tight binding method. Our calculations identify surface states in the conduction band, band gap and valence band. The surface state formed from silicon-s and pz orbitals, which is believed to account for the structure in the O K excitation spectra, lies in the band gap. It is seen that oxygen adsorption on the surface removes surface states and gives rise to a sharp peak at about — 3.8 eV below the valence band edge.  相似文献   

5.
The yield and energy distributions of lithium atoms upon electron-stimulated desorption from lithium layers adsorbed on the molybdenum surface coated with an oxygen monolayer have been measured as functions of the impact electron energy and lithium coverage. The measurements are performed using the time-of-flight technique and a surface ionization detector. The threshold of the electron-stimulated desorption of lithium atoms is equal to 25 eV, which is close to the ionization energy of the O 2s level. Above a threshold of 25 eV, the yield of lithium atoms linearly increases with an increase in the lithium coverage. In the coverage range from 0 to 0.45, an additional threshold is observed at an energy of 55 eV. This threshold can be associated with the ionization energy of the Li 1s level. At the electron energies above a threshold of 55 eV, as the coverage increases, the yield of lithium atoms passes through a maximum at a coverage of about 0.1. Additional thresholds for the electron-stimulated desorption of the lithium atoms are observed at electron energies of 40 and 70 eV for the coverages larger than 0.6 and 0.75, respectively. These thresholds correlate with the ionization energies of the Mo 4s and Mo 4p levels. Relatively broad peaks in the range of these thresholds indicate the resonance excitation of the bond and can be explained by the excitation of electrons toward the band of free states above the Fermi level. The mean kinetic energy of the lithium atoms is equal to several tenths of an electronvolt. At electron energies less than 55 eV, the energy distributions of lithium atoms involve one peak with a maximum at about 0.18 eV. For the lithium coverages less than 0.45 and electron energies higher than 55 eV, the second peak with a maximum at 0.25 eV appears in the energy distributions of the lithium atoms. The results obtained can be interpreted in the framework of the Auger-stimulated desorption model, in which the adsorbed lithium ions are neutralized after filling holes inside inner shells of the substrate and lithium atoms.  相似文献   

6.
The interaction of D2 O with a polycrystalline cerium surface, successfully cleaned by heavy Ar+ bombardment and annealing, was studied at 120 were observed at BE 530.3 (Ce2O3) and 532.7eV (adsorbed OD). When clean Ce at 120 K was exposed to D2 O, the O(1s) spectra were initial eV (adsorbed D2 O). For exposures greater than 10 Langmuir (L), a multilayer of ice grows and the O(1s) spectra become dominated by a peak at 5The results of interaction with D2 O are compared with oxidation by O2. The significant differences are: (1) the absence of Ce(IV) when oxidati relatively small extent of oxidation that occurs when Ce is exposed to D2 O at 120 K, and (3) the larger chemical-shift of the Ce(III)-derived specThe XPS studies of the interaction of D2 O with Ce reported here may be summarized as follows:(1) Exposure at 300 K gives rise to O(1s) features characteristic of oxide and hydroxide, while the Ce(3d) spectra indicate Ce(III), but no CE((2) Exposure at 120 K gives O(1s) features characteristic of adsorbed OD, chemisorbed D2 O, a multilayer of ice, and a small amount of oxide. T are characteristic of clean Ce except for slight broadening.(3) Exposure at 120 K followed by warming to 240 and 300 K gives spectra characteristic of hydroxide and oxide surface-species. Between 240 and 300 K, O(1s) intensity.(4) At 300 K, a relatively thick layer of oxide forms, and after an exposure of 50 L the features characteristic of metallic Ce are no longer observabl(5) As compared to the case for O2, exposure to D2 O gives rise to different satellite-splittings in the Ce(3d) spectra, suggesting that di formed in the two cases.(6) The spectra observed for Ce exposed to D2 O are in excellent accord with those found for the heavier lanthanides [4].  相似文献   

7.
The adsorption of N2, NH3, NO, and N2O onto clean polycrystalline dysprosium at 300 and 115 K and the changes undergone by the adsorbed species upon heating from 115 K have been investigated using X-ray photoelectron spectroscopy (XPS) and Auger electron spectroscopy (AES). At 115 K, N2 adsorbs dissociatively, vielding two peaks in the N 1s region at 396.2 and 398.2 eV corresponding respectively to a nitride and to chemisorbed nitrogen N(a). No peaks corresponding to molecularly adsorbed N2 (BE 400.2 eV [10]) were observed. Upon heating the sample the N(a) is converted into the nitride species, as evidenced by a decrease in the 398.2 eV peak and a corresponding increase in the 396.2 eV peak. At a warm-up temperature of 300 K, the N(a) species accounts for only ~10% of the total nitrogen on the surface. Ammonia adsorbed at 115 K shows three distinct peaks, at 401.7, 399.3 and 396.2 eV, corresponding to molecular, partly dissociated, and completely dissociated (nitride) ammonia. Upon heating multilayer ammonia to 175 K, it desorbs to leave predominantly the peak corresponding to the partly dissociated species. Upon further heating the molecular and partly dissociated ammonia is converted into the nitride species. At 400 K only nitride-type nitrogen remains on the surface. The adsorption of NO and N2O at 115 K is predominantly dissociative. NO has N 1s peaks at 403.1 and 396.3 eV corresponding possibly to molecularly adsorbed NO, and to nitride species. After N2O adsorption there is very little nitrogen on the surface. Adsorption of N2 and NO at 300 K yields only the peak at 396.2 eV, whereas NH3 yields, in addition to this peak, a small intensity (~20%) of the peak at 398.2 eV (partly dissociated ammonia).  相似文献   

8.
Electron energy-loss Spectroscopy (EELS) at impact energies of 2.5–3 keV has been used to obtain the electron excitation spectra for the N 1s (K-shell), F 1s (K-shell) and valence shell regions of NF3. The inner shell spectra were recorded using small angle scattering (?1° ) while the valence shell spectrum was obtained at zero degree scattering angle. The inner shell excitation spectra show a strongly enhanced 1s→ δ* type transition and continuum features which are typical for molecules with highly electronegative ligands. One of the peaks in an earlier published photoabsorption study of the N 1s region has been shown to be due to a N2 impurity. The valence shell electron energy-loss spectrum shows a number of transitions which are considered to be mainly due to valence-valence type transitions, with also some evidence of Rydberg structure.The X-ray photoelectron spectra (XPS) of the N 1s and F 1s electrons along with their associated satellite structures have also been recorded using Al Kα (1486.58 eV) radiation. The vertical ionization potentials for the N 1s and F 1s electrons were found to be 414.36 (10) eV and 693.24 (10) eV, respectively. Both spectra exhibit a rich and different satellite structure. These “shake-up” features in the satellite XPS spectra are compared with continuum features of the inner shell electron energy-loss spectra and also with the valence shell spectrum.  相似文献   

9.
The electron-stimulated desorption of Li+ ions from lithium layers adsorbed on the tantalum surface coated with a silicon film has been investigated. The measurements are performed using a static magnetic mass spectrometer equipped with an electric field-retarding energy analyzer. The threshold of the electron-stimulated desorption of lithium ions is close to the ionization energy of the Li 1s level. The secondary thresholds are observed at energies of about 130 and 150 eV. The threshold at an energy of 130 eV is approximately 30 eV higher than the ionization energy of the Si 2p level and can be associated with the double ionization. The threshold at 150 eV can be caused by the ionization of the Si 2s level. It is demonstrated that the yield of Li+ ions does not correlate with the silicon amount in near-the-surface region of the tantalum ribbon and drastically increases at high annealing temperatures. The dependence of the current of Li+ desorption on the lithium concentration upon annealing of the tantalum ribbon at T>1800 K exhibits two maxima. The ions desorbed by electrons with energies higher than 130 and 150 eV make the largest contribution to the current of lithium ions after the annealing. The yield of lithium ions upon ionization of the Li 1s level at an energy of 55 eV is considerably lesser, but it is observed at higher concentrations of deposited lithium. The results obtained can be interpreted in the framework of the Auger-stimulated desorption model with allowance made for relaxation of the local surface field.  相似文献   

10.
UPS-spectra of the cleaved (0001) Zn and (0001) O surfaces of ZnO are taken at hv = 16.8, 21.2, 26.9, and 40.8 eV. Two maxima in the spectra at constant final energy are ascribed to high densities of conduction band states. Using the hv-dependence of the valence band emission, the partial s- and p-densities of states are separated. They yield similar excitation probabilities for Zn-4s-, 3d-, and O-2p-electrons.  相似文献   

11.
Carbon 1s and silicon 2p X-ray photoelectron spectra of phenylsilane plasma polymer films prepared at substrate temperatures,Ts, between 50 and 450°C were recorded. The binding energies, lineshapes, and shake-up satellite intensities are in accordance with a structure consisting of a silicon network with pendant phenyl groups, and the minor dependence on Ts is consistent with the main effect of increasing preparation temperature being the loss of hydrogen and some pendant phenyl, with a concurrent increase in the interconnectivity of the network. The observed C 1s binding energy and linewidth specifically rule out the presence of any significant amount of carbon in silicon carbide form. A simultaneous shift of about 0.6 eV in the binding energies of both the C 1s and Si 2p lines is tentatively interpreted as a shift in the Fermi level with respect to the valence band edge.  相似文献   

12.
The valence band density of states for PbI2 is determined from X-ray and u.v. induced photoelectron spectra. It is shown that the band derived from Pb 6s states is at 8 eV binding energy and not at the top of the valence bands as suggested by band structure and charge density calculations. A rigid shift in the predominantly iodine 5p derived bands to lower binding energy brings the band structure calculations into essential agreement with experiment. Pb 5d core level binding energies determined here are used to derive core level exciton energies of 0.7 eV from published reflectivity spectra.  相似文献   

13.
The soft X-ray emission and photoelectron emission spectra of H2-, Mg- and Pt- phthalocyanine (PC) obtained using synchrotron radiation are reported and compared. In this way, an overall view of the pattern of valence bands is obtained and the electronic structure determined in terms of the component partial densities of states. In particular, from the valence p → 1s carbon and nitrogen K-emission spectra we determine for all three compounds the C and N 2p-like valence-band density of states with strong maxima located at binding energies of 8, 11 and 13.5 eV (carbon 2p) and 8 eV (nitrogen 2p) below the vacuum level. For PtPC the partial density of d-like valence states is determined from photoelectron emission difference-spectra and compared to previous XPS results. The sharp (1.2 eV FWHM) maximum of the Pt-derived partial density of states, observed at 6.9 eV binding energy, is assigned to the 4F term of a 5d86s final-state configuration. A second, broader maximum at around 9.5 eV binding energy contains contributions from other terms of this 5d8 configuration, as well as from a 5d7 satellite (shake-up multiplet).  相似文献   

14.
The adsorption and decomposition of hydrazine and ammonia by clean polycrystalline aluminium surfaces have been studied by X-ray photoelectron spectroscopy. At 85 K both ammonia and hydrazine are adsorbed molecularly, with N(1s) peaks at 400.5 eV. At 290 K hydrazine is initially adsorbed to give an N(1s) peak at about 399 eV, but with time (and further exposure) the position of the peak maximum drift to lower N(1s) values, finally approaching 397 eV after heating the ad-layer to 390 K. These observations are interpreted in terms of a slow dissociative chemisorption process: N2H4(a) → NH2(a) → NH(a) → N(a). There is no doubt that the NN bond in hydrazine is broken and that hydrogen ad-atoms formed inhibit the subsequent adsorption of N2H4 at 290 K. Ammonia dissociates more slowly than hydrazine to give mainly amine (NH2) species at 290 K.  相似文献   

15.
The electron distribution in the valence band from single crystals of titanium carbide has been studied by photoelectron spectroscopy with photon energies h?ω = 16.8, 21.2, 40.8 and 1486.6 eV. The most conspicious feature of the electron distribution curves for TiC is a hybridization between the titanium 3d and carbon 2p states at ca. 3–4-eV binding energy, and a single carbon 2s band at ca. 10 eV. By taking into account the strong symmetry and energy dependence of the photoionization crosssections, as well as the surface sensitivity, we have identified strong emission from a carbon 2p band at ? 2.9-eV energy. Our results are compared with several recent energy band structure calculations and other experimental data. Results from pure titanium, which have been used for reference purposes, are also presented.The valence band from single crystals of titanium carbide have been studied by means of photoelectron spectroscopy, with photon energies ranging from 16.8 to 1486.6 eV.By taking into account effects such as the symmetry and energy dependence of the photoionization cross-sections and surface sensitivity, we have found the valence band of titanium carbide to consist of two peaks. The upper part of the valence band at 3–4 eV below the Fermi level consists of a hybridization between Ti 3d and C 2p states. The C 2p states observed in our spectra were mainly excited from a band about 2.9 eV below the Fermi level. The APW5–9, MAPW10 and EPM11 band structure calculations predict a flat band of p-character between the symmetry points X4 and K3, most likely responsible for the majority of C 2p excitations observed. The C 2s states, on the other hand, form a single band centered around ?10.4 eV.The results obtained are consistent with several recent energy band structure calculations5–11, 13 that predict a combined bonding of covalent, ionic and metallic nature.  相似文献   

16.
The energy band structure of mechanically free and compressed LiRbSO4 single crystals is investigated. It is established that the top of the valence band is located at the D point of the Brillouin zone [k = (0.5, 0.5, 0)], the bottom of the conduction band lies at the Γ point, and the minimum direct band gap E g is equal to 5.20 eV. The bottom of the conduction band is predominantly formed by the Li s, Li p, Rb s, and Rb p states hybridized with the S p and O p antibonding states. The pressure coefficients corresponding to the energies of the valence and conduction band states and the band gap E g are determined, and the pressure dependences of the refractive indices n i are analyzed.  相似文献   

17.
A clean, polycrystalline Ce-metal surface has been successfully produced by using a sputtering and annealing technique on Ce foil. The chemical species conditions at 300 and 120 K have been characterized by XPS. Dissociative chemisorption and reaction occurs to produce a thick layer of Ce2O3. N over this Ce2O3. Two peaks occur in the O(1s) spectra, at BE 529.6 and 530.3 eV, which are assigned to CeO2 and Ce2O3, respecti eV becomes clearly apparent when CE(IV) forms. The principal species formed are oxides, and no positive evidence was found for the existence of physiso  相似文献   

18.
Binding-energy spectra obtained using the dipole (e, 2e) electron impact coincidence method have been used to derive the 3s/3p cross-section ratios for the photoionization of argon up to 75 eV. The 3s and 3p photoionization branching ratios have been obtained by making use of recently determined double photoionization yields. The partial photoionization cross-section (oscillator strength) for 3s ionization, obtained using the branching ratio and the known total photoionization cross-section, shows the deep minimum ca. 10 eV above threshold which has been predicted by those theoretical calculations which include electron correlation effects. Below 50 eV the cross-section is in excellent agreement with the SRPAE calculation. The results are in close agreement with recent measurements made using synchrotron radiation but are consistently smaller below the minimum and larger at the higher energies.  相似文献   

19.
Hydrogen cyanide (HCN) is an important intermediate during the conversion of fuel nitrogen to NOx. The mechanism of HCN oxidation to NO, N2, and N2O on the CaO (100) surface model was investigated using density functional theory calculations to elucidate the effect of in-furnace SOx removal on HCN oxidation in circulating fluidized bed boilers. HCN adsorption on the CaO (100) surface releases as high as 1.396 eV and the HC bond is strongly activated. The CaO (100) surface could catalyze the oxidation of CN radical to NCO with the energy barrier decreasing from 1.560 eV for the homogeneous case to 0.766 eV on the CaO (100) surface. The succeeding oxidation of NCO by O2 forming NO is catalyzed by the CaO (100) surface with the energy barrier decreasing from 0.349 eV (homogeneous process) to 0.026 eV on the CaO (100) surface, while the reaction between NCO and NO forming either NO or N2 is prohibited in comparison with corresponding homogeneous routes. The rate constants of these reactions under fluidized bed combustion temperature range are provided, and the calculation results lead to the conclusion that CaO (100) surface catalyzes the HCN conversion and improves the NO selectivity during HCN oxidation in the HCN/O2/NO atmosphere, which could well explain previous experimental observations. Kinetic parameters of HCN oxidation on the CaO (100) surface are provided in the Arrhenius form for future kinetic model development.  相似文献   

20.
Elastic and direct-inelastic scattering as well as dissociative adsorption and associative desorption of H2 and D2 on Ni(110) and Ni(111) surfaces were studied by molecular beam techniques. Inelastic scattering at the molecular potential is dominated by phonon interactions. With Ni(110), dissociative adsorption occurs with nearly unity sticking probability s0, irrespective of surface temperature Ts and mean kinetic energy normal to the surface 〈 E 〉. The desorbing molecules exhibit a cos θe angular distribution indicating full thermal accommodation of their translation energy. With Ni(111), on the other hand, s0 is only about 0.05 if both the gas and the surface are at room temperature. s0 is again independent of Ts, but increases continuously with 〈 E⊥ 〉 up to a value of ~0.4 forE⊥ 〉 = 0.12 eV. The cos5θe angular distribution of desorbing molecules indicates that in this case they carry off excess translational energy. The results are qualitatively rationalized in terms of a two-dimensional potential diagram with an activation barrier in the entrance channel. While the height of this barrier seems to be negligible for Ni(110), it is about 0.1 eV for Ni(111) and can be overcome through high enough translational energy by direct collision. The results show no evidence for intermediate trapping in a molecular “precursor” state on the clean surfaces, but this effect may play a role at finite coverages.  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