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.
Absolute yields of the metastable excited atoms desorbed from the surfaces of solid Ne and Ar by the creation of surface and bulk excitons have been measured using monochromated synchrotron radiation as a selective excitation source. We have obtained the absolute yields of (2.3 ± 0.7) × 10−3, (1.4 ± 0.4) × 10−3, and (7.8 ± 2.3) × 10−4 atoms/photon at the excitation of S1, B1 and S′ exciton for Ne, respectively, and 1 × 10−5 atoms/photon at S1 excitation for Ar. The probability for metastable atom desorption is found to be about 2 to 10% at the excitation of S1 exciton on the surface of solid Ne.  相似文献   

2.
We present the results of an AES study of the Si(100) electron-stimulated nitridation at RT by ammonia gas. The influence of the gas pressure and electron beam density on the nitridation rate have been determined within the ranges 10−6–10−9 Torr and 5 × 10−3–5 × 10−2 A/cm2, respectively. The silicon nitride growth rate has been found to be proportional to the electron flux and is enhanced with increased ammonia pressure in the range 10−9–10−7 Torr. Beyond 10−7 Torr the Si nitride growth rate is constant and independent of ammonia pressure. A phenomenological model of electron-stimulated nitridation process is suggested, which is in good agreement with the experimental data. The rate of electron-stimulated nitridation has been deduced.  相似文献   

3.
The adsorption of CO2 on the NaCl(100) surface was studied with a high-resolution LEED-system. Measurements without charging up at low electron energies and without damage by the e-beam could be performed by using ultrathin epitaxial films on a conducting Ge(100) substrate. The adsorption behavior was recorded as a function of time and pressure at constant substrate temperatures of 78 and 83 K and CO2 partial pressures from 4 × 10−8−2 × 10−3 Pa. The adsorption system shows a first-order two-dimensional phase transition to a (2 × 1) superstructure including glide planes (herringbone-like structure) at p = 7.2 × 10−8Pa (T = 78 K). The condensation of the CO2 solid is starting at p = 1.5 × 10−4 Pa (T = 78 K). The LEED-pattern shows in this c(2 × 2) superstructure, which corresponds to the pyrite-like structure of the CO2 solid. Both observed superstructures are commensurable with the NaCl(100) surface. Observation of island growth shows that the domains of the (2 × 1) superstructures have already at coverage of 5% of a monolayer an average lateral size of at least 200 A.  相似文献   

4.
M. Sotto 《Surface science》1992,260(1-3):235-244
A LEED and AES study on oxygen adsorption on Cu(100) and (h11) faces with 5 h 15 has been performed under various adsorption conditions (220 K T 670 K and 1 × 10−8 P 6 × 10−5 Torr of oxygen). The dependence of adsorption temp on the oxygen surface superstructures is pointed out. At least, three oxygen surface states exist on a Cu(100) face. For low temperature exposures to oxygen, under conditions of slow surface diffusion, on the (100) face, two oxygen surface phases exist: a “four spots” and a c(2 × 2) superstructure, both observed even at saturation coverage; on all the stepped faces, a c(2 × 2) appears and no faceting is observed. For high temperature exposures, on the (100) face, two oxygen superstructures are observed, a “four spots” followed by a (2√2 × √2)R45° at higher coverages; on all the stepped faces, surface diffusion is activated and oxygen induced faceting occurs. The appearance of faceting is associated with the onset of the formation of the (2√2 × √2)R45° structure on the (100) face. The oxygen induced faceting and the oxygen surface meshes are reversible with coverages. At saturation coverage, a non-reversible surface transition between the c(2 × 2) and (2√2 × √2)R45° superstructures is observed at 420 ± 20 K. The importance of impurity traces on the surface meshes is emphasized. Oxygen coverage at saturation is independent of the studied faces and adsorption temperature. Faceting occurs at a critical coverage value, whatever the stepped faces and adsorption temperature are. Models of the oxygen structure on the (h10) stepped faces are discussed.  相似文献   

5.
Y. B. Zhao  R. Gomer 《Surface science》1991,250(1-3):81-89
The electron impact behavior of CO adsorbed on Pd1/W(110) was investigated. The desorption products observed were neutral CO, CO+, and O+. After massive electron impact residual carbon, C/W = 0.15, but not oxygen was also found, suggesting that energetic neutral O, not detected in a mass analyzer must also have been formed. Formation of β-CO, i.e., dissociated CO with C and O on the surface was not seen. The total disappearance cross section varies only slightly with coverage, ranging from 9 × 10 −18 cm2 at low to 5 × 10−18 cm2 at saturation (CO/W = 0.75). The cross section for CO+ formation varies from 4 × 10−22 cm2 at satura to 2 × 10−21 cm2 at low coverage. That for O+ formation is 1.4 × 10−22 cm2 at saturation and 2 × 10−21 cm2 Threshold energies are similar to those found previously [J.C. Lin and R. Gomer, Surf. Sci. 218 (1989) 406] for CO/W(110) and CO/Cu1/W(110) which suggests similar mechanisms for product formation, with the exception of β-CO on clean W(110). It is argued that the absence or presence of β-CO in ESD hinges on its formation or absence in thermal desorption, since electron impact is likely to present the surface with vibrationally and rotationally activated CO in all cases; β-CO formation only occurs on surfaces which can dissociate such CO. It was also found that ESD of CO led to a work function increase of the remaining Pd1/W(110) surface of 500 meV, which could be annealed out only at 900 K. This is attributed to surface roughness, caused by recoil momentum of energetic desorbing entities.  相似文献   

6.
7.
The kinetics of atomic carbon and oxygen buildup on a Ni(100) surface exposed to carbon monoxide at high temperatures have been investigated by Auger electron spectroscopy. The experimental data, taken at different sample temperatures (453 , T 573 K) and at different CO partial pressures (3 ×10−7 , Pco , 3 ×10−1 mbar) allowed the identification of the CO dissociation mechanism. By fitting the experimental data with a set of rate equations describing CO dissociation, CO reduction of surface oxygen, and C and O recombination, we have been able to determine the pre-exponential factors and the activation energies of these processes.  相似文献   

8.
A combination of low-energy electron diffraction and retarding potential measurements was employed to study gaseous adsorption on atomically clean (001) and (110) Mo single crystal surfaces. Adsorption of oxygen on the (001) surface at room temperature occurred with a sticking coefficient close to unity and produced a large increase in work function and appreciable changes in the intensity distributions of the integral order diffraction beams, without the appearance of any new diffraction beams. These results indicate that a surface monolayer of oxygen was formed with a unit mesh having the same dimensions as that of the underlying molybdenum surface. Exposures above 6 × 10−3 Torr-sec produced a uniform decrease in intensities, thus indicating a second monolayer with amorphous structure. On heating, two additional surface structures were observed, characterized by one-half and one-third order beams, respectively. A clean surface was obtained by heating above 1100 °C. An exposure of 1 to 7 × 10−7 Torr-sec of oxygen for the (110) face resulted in two types of patterns characteristic of lattices with one-quarter and one-half the surface density of the (110) Mo face, with an increased work function accompanying the latter pattern. Exposure of the clean surface at 400 to 800 °C produced similar patterns of enhanced intensities with no increase in work function. Possible models are discussed. It is concluded that place exchange models account for these results, as well as the one-half and one-third order structures on the (001) face, in a more satisfactory manner than adsorption above the surface. An exposure to 10−5 Torr-sec produced a monolayer coverage with a unit mesh similar to that of the molybdenum substrate. Additional exposure resulted in further amorphous adsorption. Adsorption of CO produced changes in the intensity distributions, with the appearance of no new maxima, for both (001) and (110) Mo surfaces. Nitrogen, at an exposure of 3 × 10−3 Torr-min did not adsorb on either the (001) or (110) Mo surface, but when dissociated by electron impact it adsorbed on both Mo surfaces with the same dimensions of unit mesh as those of the Mo substrates and with an increase in work function of 1.05 eV for the (001) and 0.05 eV for the (110) surface.  相似文献   

9.
The electronic properties of SrTiO3(100) surfaces after various treatments have been studied by electron energy loss spectroscopy and Auger electron spectroscopy. A stoichiometric surface without contamination can be obtained by annealing at 910 K under oxygen atmosphere of 5 × 10−5 Pa. The surface heated under ultrahigh vacuum (UHV) at 910 K exhibits a new surface state in the band gap region, which comes from oxygen vacancies at the top Ti-O2 layer. This state is also produced by electron irradiation or Ar-ion bombardment.  相似文献   

10.
The reactions of Si(100) and Si(111) surfaces at 700 °C (973 K) with ethylene (C2H4) at a pressure of 1.3×10−4 Pa for various periods of time were studied by using Auger electron spectroscopy (AES) and electron energy loss spectroscopy (ELS). For a C2H4 exposure level, the amount of C on the (111) surface was larger than that on the (100) surface. The formation of β-SiC grain was deduced by comparing the CKLL spectra from the sample subjected to various C2H4 exposure levels, and from β-SiC crystal.  相似文献   

11.
The behavior of zirconium atoms at the W(100) surface associated with oxygen adsorption at different sample temperatures has been studied by Auger electron spectroscopy (AES), ion scattering spectroscopy (ISS), and the relative change of the work function (Δф) measured by the onset of the secondary electron energy distribution. The results have revealed: (i) adsorption of zirconium onto the W(100) surface followed by the elevation of the sample temperature up to 1710 K in an oxygen partial pressure of 2.7 × 10−4 induces complete diffusion of zirconium atoms into the W(100) substrate; (ii) further exposure of oxygen induces co-existence of oxygen and tungsten on the surface at 1710 K, resulting in a work function of 4.37 eV; (iii) keeping the sample temperature at 1710 K, simple evacuation of the system has resulted in surface segregation of zirconium atoms to the surface to form a zirconium atomic layer on the top-most surface, reducing the work function to 2.7 eV. The results have revealed that this specific behavior of zirconium atoms at high temperature assures, with very good reproducibility, the highly stable performance and long service life of Zr---O/W(100)-emitters in practical use, even in a low vacuum of 10−6 Pa.  相似文献   

12.
A C.W. multi-mode dye laser is used to obtain by optical pumping an orientation of the 2p5 3s3P0 (F = 3/2) state of 21Ne. A magnetic resonance experiment leads to the measurement of the g factor g (3P0) = 3.027 (8) × 10−4 to be compared with the theoretical prediction (3.025(6) × 10−4). One obtains also the metastability exchange cross section σ(3P0) = 18.4 ± 4 Å2 for collisions between metastable (3P0) Ne atoms and ground state Ne atoms. This result is compared with other measurements and theoretical evaluation.  相似文献   

13.
The structure and the electronic valence state occupation of ultrathin K, Rb, and Cs films grown on a GaAs(1 0 0)-(4×2) surface have been studied by means of metastable He atom scattering (MHAS), He atom scattering (HAS), and low-energy electron diffraction (LEED) at temperatures ranging from 150 to 400 K. From the survival probability of the scattered He* atoms, detailed information on the coverage-dependent filling of the alkali metal valence states and their emptying upon subsequent exposure to oxygen were derived. These data reveal for K and Rb a nearly linear band filling with increasing coverage starting at about 0.5 ML whereas a more rapid filling is observed for Cs which is almost completed at about 0.7 ML. Subsequent oxygen adsorption causes a demetallization of the metallic alkali metal monolayers. In case of Cs, a distinct minimum of the He* signal appears at an oxygen exposure of about 0.8 L, presumably indicating the onset of subsurface oxidation.  相似文献   

14.
The diffusion process of copper through grain boundaries of 500 nm thick ion-plated Ag-12at%Sn films was studied in the temperature range 100–250°C. The method is based on the determination of the time of first appearance of Cu on the Ag---Sn surface using Auger electron spectroscopy for determining trace amounts of Cu. An activation energy of Ea = 0.53 eV and a diffusivity of Do = 1.3 × 10−7cm2s−1 was obtained. For comparison, diffusion studies of Cu through ion-plated pure Ag layers have been performed. In this case an activation energy of Ea = 0.68 eV and a diffusivity of Do = 2.3 × 10−5cm2s−1 have been found.  相似文献   

15.
The interaction of oxygen with Cu(100) surfaces was investigated from 85 to 800 K by AES, ELS, LEED and work function change. At T300 K three different states of oxygen bonding are observed:

1. Chemisorption of oxygen (dosages up to 102L), indicated by an increase of the work function change Δφ and O(KLL) signal height.

2. Incorporation of oxygen into the Cu sublayer with further oxygen uptake accompanied by a decrease of Δφ, and a shift of the O(KLL) Auger transition to lower energy (102−106L).

3. Growing of Cu(I) oxide, characterized by an increase of Δφ, shifts of the oxygen and copper Auger transitions and significant changes in ELS and AES line shapes (3×106L, 10−3−5×10−1torr). At low temperature (85 K) a second adsorbed oxygen species is detected.  相似文献   


16.
Surface defects created on Ge(001) exposed to low energy Xe ions are characterized by in situ scanning tunneling microscopy (STM). The temperature of the sample during ion bombardment is 165 C and ion energies range from 20 to 240 eV. The ion collisions create defects (vacancies and adatoms) which nucleate and form vacancy and adatom islands. For fixed total vacancy creation, the vacancy island number density increases with increasing ion energy: the vacancy island number density is 1.6 × 10−20 cm−2 for 40 eV ion bombardment and increases to 4.4 × 10−20 cm−2 for 240 eV ion bombardment. The increased nucleation rate for vacancies is attributed to clustering of defects. The sputtering yield of Ge(001) is also measured by STM. The sputtering yield for 20 eV ions is approximately 10−3 per ion but the net yield for surface defects (sum of adatoms and vacancies) is an order of magnitude higher, 10−2, due to adatom-vacancy pair creation.  相似文献   

17.
The intensity of metastable helium (He*) atoms which survive during the scattering from water- and benzene-adsorbed Cu(1 0 0) surfaces was measured. The survival probability (SP) of He* was found to be sensitive to the electronic states at around the Fermi level, which is derived from the adsorbate/metal hybridization and extend toward the vacuum. The SP is likely to depend largely on the kinetic energy of the He* atoms. The kinetic energy dependence can be understood on the basis of the He* decay mechanism. Metastable-atom deexcitation spectroscopy (MDS) and surface electronic structure calculation have been used for discussing the results for the He* SP.  相似文献   

18.
We have, theoretically and experimentally, investigated the dissociation of methane on the terraces and steps of a Ni(1 1 1) surface. Using Density Functional Theory (DFT) total energy calculations combined with Ultra High Vacuum (UHV) experiments, we find that the steps exhibit a higher activity than the terraces. We have, furthermore, investigated how carbon and sulfur present on the surface will deactivate the steps, leaving only the terraces active. We find the intrinsic sticking probabilities of methane on the steps and terraces at 500 K to be 2.8 × 10−7 for the steps and 2.1 × 10−9 for the terraces, in complete agreement with our calculated difference in activation energy of 17 kJ/mol.  相似文献   

19.
Interfaces prepared by vapor deposition of Sn onto Pt(100) surfaces have been examined using the following techniques: Auger electron and X-ray photoelectron spectroscopy (AES and XPS), low-energy electron diffraction (LEED), and low-energy ion surface scattering (LEISS) with Ne+ ions. Tin deposition was conducted at 320 and 600 K, and the surface composition and order was examined as a function of further annealing to 1200 K. The AES uptake plots (signal versus deposition time) indicate that the Sn growth mode can be described by a layer-by-layer process only up to one adayer at 320 K. Some evidence of 3D growth is inferred from LEED and LEISS data for higher Sn coverages. For deposition at 600 K, AES data indicate significant interdiffusion and surface alloy formation. LEED observations (recorded at a substrate temperature of 320 K) show that the characteristic hexagonal Pt(100) reconstruction disappears with Sn exposures of 4.6 × 1014 atoms cm2Sn = 0.35 monolayer (ML)). Further Sn deposition results in a c(2 × 2) LEED pattern starting at a coverage of slightly above 0.5 ML. The c(2 × 2) LEED pattern becomes progressively more diffuse with increasing Sn exposure with eventual loss of all LEED features above 2.2 ML. Annealing experiments with various precoverages of Sn on Pt(100) are also described by AES, LEED, and LEISS results. For specific Sn precoverages and annealing conditions, c(2 × 2), p(3√2 × √2)R45°, and a combination of the two LEED patterns are observed. These ordered LEED patterns are suggested to arise from ordered PtSn surface alloys. In addition, the chemisorption of CO and O2 at the ordered annealed Sn/Pt(100) surfaces was also examined using thermal desorption mass spectroscopy (TDMS), AES, and LEED.  相似文献   

20.
D.-S. Choi  R. Gomer 《Surface science》1990,230(1-3):277-282
The diffusion of W on a (211) plane of a W field emitter has been re-examined by means of the fluctuation autocorrelation method. Diffusion along channels yielded E = 16.8 ± 0.5 kcal, D0 = (3 ± 1) × 10−5 cm2 s−1. For diffusion across channels E =6.6 kcal, D0 = 4 × 10−9cm2 s−1 at T < 752 K, and E = 24 kcal, D0 = 5 × 10−4 cm2 s−1 at T > 752 K. The results for diffusion along channels yield E and D0 values intermediate between recent results by Wang and Ehrlich [Surf. Sci. 206 (1988) 451] using field ion microscopy (E = 19 kcal, D0 = 7.7 × 10−3 cm2 s−1) and Tringides and Gomer [J. Chem. Phys. 84 (1986) 4049], using the same method as the present work but a larger slit (E = 13.3 kcal, D0 = 7 × 10−7 cm2 s−1). The results for cross channel diffus good agreement with those of Tringides and Gomer below 752 K, where these authors stopped. The new high temperature results suggest that the channel wall exchange mechanism postulated by Tringides and Gomer for cross channel diffusion at low T gives way to diffusion by climbing over the channel walls with higher E but also higher D0 above 752 K. Possible reasons for the discrepancies between these three sets of results and the absence of cross channel diffusion in the work of Wang and Ehrlich are briefly discussed.  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