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.
Oxygen adsorption on the LaB6(100), (110) and (111) clean surfaces has been studied by means of UPS, XPS and LEED. The results on oxygen adsorption will be discussed on the basis of the structurs and the electronic states on the LaB6(100), (110) and (111) clean surfaces. The surface states on LaB6(110) disappear at the oxygen exposure of 0.4 L where a c(2 × 2) LEED pattern disappears and a (1 × 1) LEED pattern appears. The work function on LaB6(110) is increased to ~3.8 eV by an oxygen exposure of ~2 L. The surface states on LaB6(111) disappear at an oxygen exposure of ~2 L where the work function has a maximum value of ~4.4 eV. Oxygen is adsorbed on the surface boron atoms of LaB6(111) until an exposure of ~2 L. Above this exposure, oxygen is adsorbed on another site to lower the work function from ~4.4 to ~3.8 eV until an oxygen exposure of ~100L. The initial sticking coefficient on LaB6(110) has the highest value of ~1 among the (100), (110) and (111) surfaces. The (100) surface is most stable to oxygen among these surfaces. It is suggested that the dangling bonds of boron atoms play an important role in oxygen adsorption on the LaB6 surfaces.  相似文献   

2.
Thermionic emission properties of the single crystal hexaborides LaB6, CeB6, PrB6, NdB6, SmB6, EuB6, (La, Sr)B6, (La, Ba)B6, (La, Ce)B6, (La, Pr)B6, (La, Sm)B6, and (La, Dy)B6 are measured in the temperature range between 1250 and 1700°C. Of these, LaB6 is shown to have the highest emission current density in the temperature range investigated. The LaB6-based mixed hexaborides, (La, M)B6, show current densities similar to LaB6, but a little lower. Analyses by Auger electron spectroscopy indicate that the surface composition of (La, M)B6 approaches that of LaB6 at elevated temperatures and that the thickness of the surface layer whose composition is different from that of the bulk is typically several atomic layers. The formation of the surface layer is considered to be caused by a relatively slow evaporation rate of La compared to that of the other metal.  相似文献   

3.
The adsorption of oxygen on Ag(111) has been studied by ellipsometry in conjunction with AES and LEED. The oxygen pressure varied between 10?5 and 10?3 Torr and the crystal temperature between room temperature and 250° C. Changes in the Auger spectrum and the LEED pattern upon oxygen adsorption are very small. Oxygen coverages were derived from the changes in the ellipsometric parameter Δ. At room temperature a maximum coverage is reached within a few minutes. Its value increases with the damage produced by the preceding argon ion bombardment. The sticking coefficient derived from the initial rate of Δ-change amounts to 3 × 10?5 for well-annealed surfaces and 2.5 ? 5 × 10?4 for damaged surfaces. After evacuation no desorption takes place. Other types of adsorption, associated with much larger changes in Δ, were observed upon bombardment with oxygen ions and with oxygen activated by a hot filament. The reaction of CO with adsorbed oxygen was studied ellipsometrically at room temperature in the CO pressure range 10?7–10?6 Torr. The initial reaction rate is proportional to the CO pressure. The reaction probability (number of oxygen atoms removed per incident CO molecule) is 0.36.  相似文献   

4.
The system Cu(210)-O2 has been examined using LEED and AES, combined with optical simulation of diffraction patterns to investigate the detailed structure of the adsorbed layer. Exposure at 300 K and 5 × 10?9 Torr resulted in LEED patterns showing pronounced streaks. The corresponding structures are believed to require an adsorption mechanism in which O2 dissociation can occur only at a limited number of surface sites and in which O atoms after dissociation diffuse over quite large distances (?10 nm) before becoming chemisorbed. Heating these structures to 500–600 K produced a sharp (2 × 1) pattern; this step is thought to involve equilibration of the adsorbed layer. Further combinations of exposure (?1 × 10?6Torr) and heating (up to 500 K) resulted in a series of (2 × 1) and (3 × 1) patterns, while heating to 800 K at any stage of the oxygen interaction regenerated the clean surface.  相似文献   

5.
《Surface science》1986,171(1):L425-L431
The surface oxidation of epitaxial and polycrystalline Gd samples grown in ultrahigh vacuum on W(110) substrates has been investigated using Auger-electron spectroscopy (AES) and low energy electron diffraction (LEED). The surface crystallography of clean epitaxial films monitored by LEED is hcp(0001) and remains unchanged even after 300 L oxygen exposure at room temperature. The LEED pattern of bulk Gd2O3 in Mn2O3 structure is observed only when oxygen is exposed at an elevated substrate temperature of about 500°C. AES clearly reveals various stages of oxidation as a function of the oxygen exposure for epitaxial as well as polycrystalline films. It is found that the oxidation does not proceed beyond one monolayer of the Gd surface.  相似文献   

6.
Electron beam assisted adsorption and desorption of oxygen was studied by Auger electron spectroscopy (AES). Beam assisted adsorption was observed on clean as well as on oxidized surfaces. After an oxygen exposure of 1000 × 10?7 Torr min and continuous irradiation with beam voltage of 1.5 kV and beam current density 2 microA mm?2, the oxygen 510 eV signal amplitude from the point of beam impact was 2.5 times greater than the signal from the non-irradiated region. The Ge 89 eV signal showed a corresponding decrease. Enhanced adsorption occurred at beam energies as low as 16.5 eV. After irradiation, the oxidized surface was not carbon contaminated. Following an oxygen exposure of 30 min at 0.1 Torr and 550°C and subsequent additional beam assisted exposure of 1000 × 10?7 Torr min, the maximum oxide thickness was about 18 Å. Beam assisted desorption did not occur from thin oxygen layers (0–510 eV signal strength less than 5 units, calculated oxide thickness about 6 Å), but occurred from thick oxides and stopped after the signal amplitude had decreased to 5 units. Based on these results, a model for the structure of the oxygen layer covering the Ge(111) surface is proposed. Mechanisms for adsorption and desorption are discussed. The implications of beam assisted adsorption and desorption on electron beam operated surface measurements (LEED, AES, ELS, APS etc.) are stressed.  相似文献   

7.
The growth of ultrathin films of Y2O3(111) on Pt(111) has been studied using scanning tunneling microscopy (STM), X-ray photoemission spectroscopy (XPS), and low energy electron diffraction (LEED). The films were grown by physical vapor deposition of yttrium in a 10? 6 Torr oxygen atmosphere. Continuous Y2O3(111) films were obtained by post-growth annealing at 700 °C. LEED and STM indicate an ordered film with a bulk-truncated Y2O3(111)–1 × 1 structure exposed. Furthermore, despite the lattices of the substrate and the oxide film being incommensurate, the two lattices exhibit a strict in-plane orientation relationship with the [11?0] directions of the two cubic lattices aligning parallel to each other. XPS measurements suggest hydroxyls to be easily formed at the Y2O3 surface at room temperature even under ultra high vacuum conditions. The hydrogen desorbs from the yttria surface above ~ 200 °C.  相似文献   

8.
The adsorption of hydrogen on Pt (100) was investigated by utilizing LEED, Auger electron spectroscopy and flash desorption mass spectrometry. No new LEED structures were found during the adsorption of hydrogen. One desorption peak was detected by flash desorption with a desorption maximum at 160 °C. Quantitative evaluation of the flash desorption spectra yields a saturation coverage of 4.6 × 1014 atoms/cm2 at room temperature with an initial sticking probability of 0.17. Second order desorption kinetics was observed and a desorption energy of 15–16 kcal/mole has been deduced. The shapes of the flash desorption spectra are discussed in terms of lateral interactions in the adsorbate and of the existence of two substates at the surface. The reaction between hydrogen and oxygen on Pt (100) has been investigated by monitoring the reaction product H2O in a mass spectrometer. The temperature dependence of the reaction proved to be complex and different reaction mechanisms might be dominant at different temperatures. Oxygen excess in the gas phase inhibits the reaction by blocking reactive surface sites. At least two adsorption states of H2O have to be considered on Pt (100). Desorption from the prevailing low energy state occurs below room temperature. Flash desorption spectra of strongly bound H2O coadsorbed with hydrogen and oxygen have been obtained with desorption maxima at 190 °C and 340 °C.  相似文献   

9.
 采用电子束蒸发方法在大面积玻璃基底和钽基底上沉积六硼化镧薄膜阴极。分别对玻璃基底上沉积的六硼化镧薄膜的生长取向、附着力与不同蒸发角度(0°, 30°,45°和60°)的关系进行了研究;对钽基底上沉积的六硼化镧薄膜阴极的逸出功进行了研究。结果表明:在基底温度为250 ℃时,制备的六硼化镧薄膜具有(100)晶面择优生长的特点;蒸发角度为45°时,六硼化镧薄膜(100)晶面的晶格常数与靶材相差最小,晶粒较小;根据优化的工艺制备的六硼化镧薄膜阴极的逸出功为2.56 eV。  相似文献   

10.
A structural study of the different stages during NiO(100) sulphurization by H2S was carried out by RHEED, LEED and AES. On exposure to H2S (PH2S < 10?5 Torr) The “clean” surface, obtained by UHV cleavage, was found to react with H2S to produce islands of Ni(100) covered with an ordered c(2 × 2) S structure up to 300°C. Growth of Ni3S2 islands occurs on increasing the temperature and the exposure to H2S.  相似文献   

11.
The initial stages of oxidation of the (100) surface of a single crystal alloy specimen of approximate atomic composition Ni 59, Fe 41 (at%) have been studied by Auger spectroscopy and electron diffraction techniques. The clean alloy surface shows only a slight iron enrichment over the temperature range of the oxidation studies (373–873 K). Oxidation studies were performed over the O2 pressure range 5 × 10?9 to 1 × 10?6 Torr. Within these experimental conditions the rate of oxygen uptake was found to be linear in pressure and essentially independent of temperature. LEED studies showed that a chemisorbed c(2 × 2) structure preceded the formation of surface oxides. The interaction of oxygen with the surface induced a marked segregation of iron and this was particularly pronounced at elevated temperatures. Chemical shifts were observed in the low energy Ni and Fe Auger spectra during oxidation; these were similar to those previously observed in separate studies of the oxidation of pure Ni and of pure Fe. At the higher temperatures the initial oxide layer grew epitaxially apparently as a (111) cubic oxide on the (100) substrate. The Ni to Fe concentration ratio in oxides several layers thick was found to depend on the temperature of the reaction; at higher temperatures the oxide were more Fe-rich. The Fe to Ni ratio in oxides produced at lower temperatures could be increased by annealing. At large O2 exposures (about 5000 L) a transition was observed in the structure of the oxide layer.  相似文献   

12.
Clean and oxidized surfaces of tellurium films have been studied using electron-excited Auger electron spectroscopy, X-ray photoelectron spectroscopy, energy loss spectroscopy, and electron-stimulated desorption. The results for clean tellurium are in general agreement with previous studies, but the oxidation studies have provided new information. Reaction between oxygen gas and tellurium was found to be an activated process requiring tellurium temperatures in excess of 60°C to produce detectable oxide for 30 min exposures to ~800 Torr of oxygen. Increasing the temperature to 200°C produced a 10.6 A layer of TeO2. This layer was rapidly removed by electron irradiation with a cross-section of ~3 × 10?18 cm2 for electron-stimulated desorption of oxygen by 2 keV electrons; however, there was evidence for reduced cross sections for thinner TeO2 layers.  相似文献   

13.
The de Haas-van Alphen effect has been observed in the hexaborides of La, Ce, Pr and Nd with the direction of the field parallel to [100]. The frequencies for the heavier rare-earth compounds are close to those for LaB6; large values for the mass compared to the ones in LaB6 and strong anisotropy of scattering has been observed.  相似文献   

14.
The initial oxidation of Al(111) has been studied with photoemission and surface EXAFS for single crystal surfaces. We find that the oxidation is pressure dependent and that at pressures below 2×10?7 Torr molecular oxygen chemisorbs while at pressures above 1×10?6 Torr the chemisorption is disassociative. These results are discussed in the context of other LEED and surface EXAFS studies.  相似文献   

15.
Adlayers of oxygen, carbon, and sulfur on W(211) have been characterized by LEED, AES, TPD, and CO adsorption. Oxygen initially adsorbs on the W(211) surface forming p(2 × 1)O and p(1 × 1)O structures. Atomic oxygen is the only desorption product from these surfaces. This initial adsorption selectively inhibits CO dissociation in the CO(β1) state. Increased oxidation leads to a p(1 × 1)O structure which totally inhibits CO dissociation. Volatile metal oxides desorb from the p(1 × 1)O surface at 1850 K. Oxidation of W(211) at 1200 K leads to reconstruction of the surface and formation of p(1 × n)O LEED patterns, 3 ? n ? 7. The reconstructed surface also inhibits CO dissociation and volatile metal oxides are observed to desorb at 1700 K, as well as at 1850 K. Carburization of the W(211) surface below 1000 K produced no ordered structures. Above 1000 K carburization produces a c(6 × 4)C which is suggested to result from a hexagonal tungsten carbide overlayer. CO dissociation is inhibited on the W(211)?c(6×4)C surface. Sulfur initially orders into a c(2 × 2)S structure on W(211). Increased coverage leads to a c(2×6)S structure and then a complex structure. Adsorbed sulfur reduces CO dissociation on W(211), but even at the highest sulfur coverages CO dissociation was observed. Sulfur was found to desorb as atomic S at 1850 K for sulfur coverages less than 76 monolayers. At higher sulfur coverages the dimer, S2, was observed to desorb at 1700 K in addition to atomic sulfur desorption.  相似文献   

16.
The interaction of CO, O2, H2, N2, C2H4 and C6H6 with an Ir(110) surface has been studied using LEED, Auger electron spectroscopy and flash desorption mass spectroscopy. Adsorption of oxygen at 30°C produces a (1× 2) structure, while a c(2 × 2) structure is formed at 400°C. Two peaks have been detected in the thermal desorption spectrum of oxygen following adsorption at 30°C. The heat of adsorption of hydrogen is slightly higher on Ir(110) than on Ir(111). Adsorption of carbon monoxide at 30°C produces a (2 × 1) surface structure. The main CO desorption peak is found around 230, while two other desorption peaks are observed around 340 and 160°C. At exposures between 250 and 500°C carbon monoxide adsorption yields a c(2 × 2) structure and a desorption peak around 600°C. Carbon monoxide is adsorbed on an Ir(110) surface partly covered with oxygen or carbon in a new binding state with a significantly higher desorption temperature than on the clean surface. Adsorption of nitrogen could not be detected on either clean or on carbon covered Ir(110) surfaces. The hydrocarbon molecules do not form ordered surface structures on Ir(110). The thermal desorption spectra obtained after adsorption of C6H6 or C2H4 are similar to those reported previously for Ir(111) consisting mostly of hydrogen. Heating the (110) surface above 700°C in the presence of C6H6 or C2H4 results in the formation of an ordered carbonaceous overlayer with (1 × 1) structure. The results are compared with those obtained previously on the Ir(111) and Ir(755) or stepped [6(111) × (100)] surfaces. The CO adsorption results are discussed in relation to data on similar surfaces of other Group VIII metals.  相似文献   

17.
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   

18.
In analogy with the case of Sr on Si [Y. Liang, S. Gan, M. Engelhard, Appl. Phys. Lett. 79 (2001) 3591], we studied surface crystallinity and oxidation behaviour of clean and Ba terminated Ge(1 0 0) surfaces as a function of oxygen pressure and temperature. The structural and chemical changes in the Ge surface layer were monitored by LEED, XPS and real-time RHEED. In contrast to the oxidation retarding effect, observed for 1/2 monolayer of Sr on Si, the presence of a Ba termination layer leads to a pronounced increase in Ge oxidation rate with respect to clean Ge. In fact, while the Ge(1 0 0) surface terminated with 1/2 ML Ba amorphizes for a pO2 of 10−2 Torr, LEED indicates that clean Ge forms a thin (4.5 Å), 1 × 1 ordered oxide upon aggressive O2 exposure (150 Torr, 200 °C, 30 min). We briefly discuss the origins for the difference in behaviour between Ba on Ge and Sr on Si.  相似文献   

19.
The structure formed during solid-state synthesis in thin bilayer Al/Ni films with the ratio Al: Ni = 60: 40 (at %) has been investigated. The films were obtained by thermal evaporation in vacuum with a residual pressure of 10?5–10?6 Torr. Solid-state synthesis was performed by diffusion reaction. The sequence of phase formation upon vacuum annealing of bilayer Al/Ni films has been established: Al + Ni → Al3Ni + Ni (T ann = 180°C) → Al3Ni2 (T ann = 220°C).  相似文献   

20.
Hydrogenated amorphous silicon surfaces, atomically clean and subsequently oxidized to up to 20 Å oxide thickness, were studied using AES and UPS. The oxidation was made in O2 in the pressure range 10?9 Torr to 5 atm and at 23 and 300°C. The oxidation rate at 23°C was found to be the same as that of crystalline silicon while at 300°C it was appreciably faster. Changes in the d N(E)dE AES Si LVV line shape near 80 eV upon oxidation could be correlated to changes in the silicon-oxygen bonding level observed in UPS. The detailed line shape of the AES Si LVV transition indicates that at 300°C a more homogeneous oxide is produced than at 23°C.  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