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 co-adsorption of Cu on O2 and a W{100}surface is studied by Auger electron spectroscopy (AES), thermal desorption (TD), low energy electron diffraction (LEED) and by work function change (δø) measurements. It is shown that the presence of Cu on the surface initially decreases sO, the sticking coefficient of O2. For longer oxygen exposures and for higher adsorption temperatures, θO reaches values larger than those on the clean surface for the same O2 exposure. Except at the highest θO values and temperatures, the sticcking coefficient for copper, sCu, is unity and is independent of the oxygen coverage θO in the range studied (0 ? θO ? 2). Co-adsorption at room temperatures does not produce any long range order while co-adsorption at elevated temperature leads to the ordered structures (1 × 1), p(2 × 1), p(2 × 2) and c(2 × 2). The saturation coverage of the two dimensional co-adsorbate at 800 K is given by the relation θCu + 85 θO = 2. The work function is a complicated function of θO and θCu and is determined predominantly by the temperature at which oxygen is adsorbed. At high temperatures the sequence of adsorption has no influence, in contrast to the room temperature behavior.  相似文献   

2.
An intensity analysis with low-energy electron diffraction is reported for the (3 × 3)30° surface structure obtained by the adsorption and presumed dissociation of H2S on the (111) surface of rhodium. Intensity-versus-energy curves were measured with a video LEED analyser for nine diffracted beams at normal incidence, and comparisons made with the renormalised forward scattering method for four different types of structural models in which the metal atoms remain in their regular bulk positions. The best correspondence between the experimental and calculated intensities occurs with sulphur atoms adsorbed in the “expected” 3-coordinate adsorption sites. The reliability index proposed by Pendry is minimised with S atoms 1.53 Å above the topmost metal layer; this corresponds to nearest-neighbour RhS bond distances equal to 2.18 Å. Comparisons are made with structural data available for related systems, and with the predictions of a model analysis of surface bond lengths given recently by one of the authors.  相似文献   

3.
The binding energies and eigenfunctions of the surface states of positrons outside materials with negative positron affinity are considered. When the material is represented by a continuum hydrodynamic model with finite dispersion the potential takes the Newns-Barton form. The characteristic material parameter is B = ωp/β where ωp is the bulk plasma frequency and β ? √35x the Fermi velocity. For aluminium B ? 0.6a?1 (a is the Bohr radius) and the Newns-Barton potential yields a ground state binding energy E(1) ? 0.46 ev - considerably smaller than the binding energy of the pure image (or non dispersive) potential for which E(1) ? 0.85 ev.  相似文献   

4.
Adsorption of CO on Ni(111) surfaces was studied by means of LEED, UPS and thermal desorption spectroscopy. On an initially clean surface adsorbed CO forms a √3 × √3R30° structure at θ = 0.33 whose unit cell is continuously compressed with increasing coverage leading to a c4 × 2-structure at θ = 0.5. Beyond this coverage a more weakly bound phase characterized by a √72 × √72R19° LEED pattern is formed which is interpreted with a hexagonal close-packed arrangement (θ = 0.57) where all CO molecules are either in “bridge” or in single-site positions with a mutual distance of 3.3 Å. If CO is adsorbed on a surface precovered by oxygen (exhibiting an O 2 × 2 structure) a partially disordered coadsorbate 2 × 2 structure with θo = θco = 0.25 is formed where the CO adsorption energy is lowered by about 4 kcal/mole due to repulsive interactions. In this case the photoemission spectrum exhibits not a simple superposition of the features arising from the single-component adsorbates (i.e. maxima at 5.5 eV below the Fermi level with Oad, and at 7.8 (5σ + 1π) and 10.6 eV (4σ) with COad, respectively), but the peak derived from the CO 4σ level is shifted by about 0.3 eV towards higher ionization energies.  相似文献   

5.
Two different 3 × 3 R30° adatom models for 13 monolayer Ge coverage of the Si(111)?2 × 1 surface were studied with the Pseudopotential method. Total energy calculations indicate that the T4 model (adatom above the second layer) is preferable to the H3 model (adatom above the hollow site)  相似文献   

6.
The interaction of NO with a Ni (111) surface was studied by means of LEED, AES, UPS and flash desorption spectroscopy. NO adsorbs with a high sticking probability and may form two ordered structures (c4 × 2 and hexagonal) from (undissociated) NOad. The mean adsorption energy is about 25 kcalmole. Dissociation of adsorbed NO starts already at ?120°C, but the activation energy for this process increases with increasing coverage (and even by the presence of preadsorbed oxygen) up to the value for the activation energy of NO desorption. The recombination of adsorbed nitrogen atoms and desorption of N2 occurs around 600 °C with an activation energy of about 52 kcalmole. A chemisorbed oxygen layer converts upon further increase of the oxygen concentration into epitaxial NiO. A mixed layer consisting of Nad + Oad (after thermal decomposition of NO) exhibits a complex LEED pattern and can be stripped of adsorbed oxygen by reduction with H2. This yields an Nad overlayer exhibiting a 6 × 2 LEED pattern. A series of new maxima at ≈ ?2, ?8.8 and ?14.6 eV is observed in the UV photoelectron spectra from adsorbed NO which are identified with surface states derived from molecular orbitals of free NO. Nad as well as Oad causes a peak at ?5.6 eV which is derived from the 2p electrons of the adsorbate. The photoelectron spectrum from NiO agrees closely with a recent theoretical evaluation.  相似文献   

7.
Low energy electron diffraction (LEED), Auger electron spectroscopy (AES) and photoemission yield spectroscopy (PYS) measurements have been performed on a set of ultrahigh vacuum cleaved Si(111) surfaces with different bulk dopings as a function of Ga or In coverage θ. The metal layers are obtained by evaporation on the unheated substrate and θ varies from zero to several monolayers (ML). First, the 2×1 reconstruction of the clean substrate is replaced by a 3×3 R30° structure at 13 ML, meanwhile the dangling bond peak at 0.6 eV below the valence band edge Evs is replaced by a peak at 0.1 eV for Ga or 0.3 eV for In, below Evs. At the same time, the ionization energy decreases by 0.4 eV (Ga) or 0.6 eV (In), while the Fermi level pinning position gets closer to the valence band edge by about 0.1eV. Upon increasing θ, new LEED structures develop and the electronic properties keep on changing slightly before metallic islands start to grow beyond θ ~1 ML.  相似文献   

8.
Recently, microwave studies of the rotational spectra of water and its various isotopic species have been reported. These studies provide rotational constants and among others the quartic distortion constants, which depend on the quadratic part of the vibrational potential function. These data are collected and discussed, and the molecular force field and structure of water is considered in light of this recent microwave data. The quartic distortion data gives force constants which are very reasonable considering the difficulties in the distortion analysis of these light molecules, where as many as 22 parameters are being evaluated to fit the observed spectrum. The infrared and microwave data are combined within the theoretical framework of the small oscillations model and the results compare favorably with the true harmonic force field. The infrared and microwave valence bond force constants of H2O are (mdyn/Å):
?r=7.746, ?θ=0.700, ?rr=?0.093, ?=0.379
The results further confirm the usefulness of rotation-vibration data in the determination of force constans, and show that even for water with extremely large anharmonicity effects, a very representative force field can be obtained by combining ground state infrared and microwave data.Various molecular structures have been evaluated, and the average structures in the ground vibrational state for H2O, D2O and T2O are found to be:
〈r〉〈θO-H=0.9724 A?HOH=104.50°O-D=0.9687 A?DOD=104.35°O-T=0.9671 A?TOT=104.26°
A one-dimensional approximation to the anharmonicity effects is applied to determine the equilibrium molecular structure of H2O from the average structure data. The result is as follows:
re=0.9587 A? and θe=103.9°
  相似文献   

9.
Experimental data are presented for the angular dependence of the relative flux, the mean energy and the speed ratio of deuterium molecules desorbing from a Ni(111) crystal surface at a surface temperature of Ts = 1143 K and at sulphur coverages ranging between 30% and less than 2% of a monolayer.The angular flux distribution is sharply peaked in the forward direction (cosdθwith 3 ? d ? 5) and the mean energy 〈E〉 of the desorbate depends strongly on the desorption angle θ. For normal desorption (θ = 0°) 〈E〉2k is about 700 K higher than Ts and for glancing angles (θ = 80°) it decreases to about 400 K below Ts The results obtained on sulphur free and sulphur covered Ni(111) surfaces are compared with our former data on polycrystalline nickel. The main differences in the kinetic features can be ascribed to the surface roughness. Accordingly, the angular distributions of flux, mean energy, and speed ratio, which deviate strongly from the Knudson and Maxwellian law, do not seem to depend considerably on sulphur coverage and surface structure. A qualitative explanation for these deviations is presented using the principle of detailed balancing.  相似文献   

10.
The inclusive spectrum of positive particles has been measured at the ISR energy √s = 52.5 GeV as function of transverse momentum, pT for large pT values using the Split Field Magnet facility. The angular dependence of the distribution is observed and discussed in the range 9° ? θ1 ? 21°. An upper limit for the production of ? meson with pT ? 3 GeV/c is derived from the same sample of data.  相似文献   

11.
12.
Internal conversion of soft virtual bremsstrahlung, calculated in an essentially model independent infrared limit, yields numerous low mass-low energy lepton pairs. The resulting single lepton spectra rise sharply towards pT = 0 for large angles θcm and towards x = 0 for θcm = 0°. The ratio eπ for pT ? 400 MeV/c exhibits roughly the experimentally observed properties.  相似文献   

13.
Angular-resolved photoemission spectra from a (2 × 2)R45° sulphur layer on Pd(100) reveal peaks independent of photon energy, which show strong dispersion in the electron energy range 3…9 eV relative to the vacuum level. The appearance of these levels can be correlated with absolute gaps in the projected bulk band structure of palladium. We interpret these features, which are also observed in secondary electron emission, as adsorbate-induced surface resonances above EF. The sulphur 3p-derived levels below EF show strong dispersion effects in the [110] but not in the [100] crystal azimuth. The projected band structure also shows that sd hybridisation gives rise to an absolute gap at approximately this energy in the [110] (ΓX) direction.  相似文献   

14.
The time evolution of the KLL Auger spectrum of carbon as a function of temperature is used to derive the kinetics of the surface diffusion and bulk-to-surface precipitation of carbon on polycrystalline nickel. The results show that the activation energy for the surface diffusion of carbon atoms on polycrystalline nickel is 6.9 ± 0.6 kcalmole, and the activation energy for bulk-to-surface precipitation is 9.4 ± 0.6 kcalmole. The dependence on the surface diffusion coefficient Ds (cm2s?1), on the absolute temperature T can be represented, over the experimental temperature range, 350–425° C, by: ln Ds = 10.27 ? 3568T.  相似文献   

15.
A detailed LEED study is reported of the surface phases stabilised by hydrogen chemisorption on W {001}, over the temperature range 170 to 400 K, correlated with absolute determinations of surface coverages and sticking probabilities. The saturation coverage at 300 K is 19(± 3) × 1014 atoms cm?2, corresponding to a surface stoichiometry of WH2, and the initial sticking probability for both H2 and D2 is 0.60 ± 0.03, independent of substrate temperature down to 170 K. Over the range 170 to 300 K six coverage-dependent temperature-independent phases are identified, and the transition coverages determined. As with the clean surface (2 × 2)R45° displacive phase, the c(2 × 2)-H phase is inhibited by the presence of steps and impurities over large distances (~20 Å), again strongly indicative of CDW-PLD mechanisms for the formation of the H-stabilised phases. These phases are significantly more temperature stable than the clean (2 × 2)R45°, the most stable being a c(2 × 2)-H split half-order phase which is formed at domain stoichiometries between WH0.3 and WH0.5. LEED symmetry analysis, the dependence of half-order intensity and half-width on coverage, and I-V spectra indicate that the c(2 × 2)-H phase is a different displacive structure from that determined by Debe and King for the clean (2 × 2)R45°. LEED I-V spectra are consistent with an expansion of the surface-bulk interlayer spacing from 1.48 to 1.51 Å as the hydrogen coverage increases to ~4 × 1014 atoms cm?2. The transition from the split half-order to a streaked half-order phase is found to be correlated with changes in a range of other physical properties previously reported for this system. As the surface stoichiometry increases from WH to WH2 a gradual transition occurs between a phase devoid of long-range order to well-ordered (1 × 1)-H. Displacive structures are proposed for the various phases formed, based on the hypothesis that at any coverage the most stable phase is determined by the gain in stability produced by a combination of chemical bonding to form a local surface complex and electron-phonon coupling to produce a periodic lattice distortion. The sequence of commensurate, incommensurate and disordered structures are consistent with the wealth of data now available for this system. Finally, a simple structural model is suggested for the peak-splitting observed in desorption spectra.  相似文献   

16.
Interactions between oxygen under low pressure and a niobium-oxygen solid solution had been studied, in the regime where adsorption is the rate-determining step, from 1000 to 1700 K. It is shown that at saturation of solid solution, there exists a constant limiting value Θl of superficial coverage, comparable to a limiting bulk concentration cl. The ratios θ = Θ/Θl and ? = c/cl are called “relative ratio of occupation” (superficial and bulk). KSV is the equilibrium constant of segregation between adsorbed and dissolved oxygen atoms: (Odiss?v) + σ ? (Ochim?σ) + v (σ and v being respectively surface and bulk sites), KSV = [(1 ? θ)/θ] [?/(1 ? ?)]. The experimentally determined expression: KSV = 5.7 exp[?(22.1 ? 12.1 θ)/ RT] shows that lateral superficial interactions have a large influence on the enthalpy of transfer between the bulk and the surface of the sample. Adsorption is direct and non activated. At the solubility limit, only a fraction of the superficial sites is occupied. We estimate it to be one half. The sticking probability b of oxygen on a niobium oxygen solid solution is given by b = (1 ? θ/2)2, its value at zero coverage being estimated as unity.  相似文献   

17.
The self-consistent pseudopotential method is applied to the Si (111) 7 × 7 reconstructed surface in the vacancy model with a simplified 3 × 3 superlattice structure. Numerical results with and without relaxation of surface atoms are presented. It is concluded that the relaxation, if any, is to be much smaller than the atomic distance to explain the photoemission spectrum of the 7 × 7 surface. The importance of the many-body effect is suggested in the photoemission process associated with the dangling bond surface states of Si.  相似文献   

18.
The external molecular vibrations of deuterated imidazole have been investigated at 12 K for wave vectors along the three principal directions of the Brillouin zone by neutron scattering and by Raman scattering at the zone center. A complete symmetry assignment is obtained. The dominant role of the N-H … N hydrogen bond within the intermolecular interactions is reflected in a relatively simple manner in the shape of the dispersion curves because of the chain-like arrangement of hydrogen bonds. The experiment yields directly a stretching force constant of 0.33 mdyn Å?1 for this bond. This value is compatible with monopole-dipole and dipole-dipole interactions as major contributions to the attractive interactions of the hydrogen bond of this length (2.86 Å). Relatively strong static monopoles and dipoles are confirmed by a CNDO2 calculation and are consequently incorporated into the interaction model.  相似文献   

19.
The adsorption of Xe on a Ni(100) surface has been studied in UHV between 30 and 100 K using LEED, thermal desorption spectroscopy (TDS), work function (Δφ) measurements, and UV photoemission (UPS). At and below 80 K, Xe adsorbs readily with high initial sticking probability and via precursor state adsorption kinetics to form a partially ordered phase. This phase has a binding energy of ~5.2 kcal/mole as determined by isosteric heat measurements. The heat of adsorption is fairly constant up to medium coverages and then drops continuously as the coverage increases, indicating repulsive mutual interactions. The thermal desorption is first order with a preexponential factor of about 1012 s?1, indicative of completely mobile adsorption. Adsorbed Xe lowers the work function of the Ni surface by 376 mV at monolayer coverage. (This coverage is determined from LEED to be 5.65 × 1014 Xe molecules/cm-2.) For not too high coverages, θ, Δφ(θ) can be described by the Topping model, with the initial dipole moment μ0 = 0.29 D and the polarizability α being 3.5 × 10?24 cm3. In photoemission, the Xe 5p32 and 5p12 orbitals show up as intense peaks at 5.56 and 6.83 eV below Ef which do not shift their position as the coverage varies. Multilayer adsorption (i.e. the filling of the second and third layers) can be seen by TDS. The binding energies of these α states can be estimated to range between 4.5 and 3.5 kcal/mole. The results are compared and contrasted with previous findings of Xe adsorption on other transition metal surfaces and are discussed with respect to the nature of the inert-gas-metal adsorptive bond.  相似文献   

20.
Local chemisorption geometries of formate (HCO2) and methoxy (CH3O) groups on Cu(100) were examined by means of surface extended (SEXAFS) and near-edge X-ray absorption fine structure (NEXAFS) measurements above the O K edge. At 300 K the oxygen of the formate group are equivalent and asymmetrically located near the four-fold hollow site yielding two CuO distances between 2.31 and 2.45 Å. These distances are at least 0.3 Å longer than typical CuO distances of surface and bulk compounds due to a Cu-C steric interaction. The CO bonds and the OCO angle of the formate are estimated to be 1.27 ± 0.04 A? and 127 ± 7°, respectively. At 200 K the methoxy group has a CuO distance of 1.97 ± 0.05 A? and the CO axis is tilted with respect to the surface normal. The exact chemisorption site of the methoxy goup could not be determined, but the atop site is ruled-out.  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