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.
2.
The adsorption of potassium and the coadsorption of potassium and oxygen on the Pt(111) and stepped Pt(755) crystal surfaces were studied by AES, LEED, and TDS. Pure potassium adlayers were found by LEED to be hexagonally ordered on Pt(111) at coverages of θ = K0.9–;1. The monolayer coverage was 5.4 × 1014K atoms/cm2 (0.36 times the atomic density of the Pt(111) surface). Orientational reordering of the adlayers, similar to the behavior of noble gas phase transitions on metals, was observed. The heat of desorption of K decreased, due to depolarization effects, from 60 kcal/mole at θK <0.1, to 25 kcal/mole at θK = 1 on both Pt(111) and Pt(755). Exposure to oxygen thermally stabilizes a potassium monolayer, increasing the heat of desorption from 25 to 50 kcal/mole. Both potassium and oxygen were found to desorb simultaneously indicating strong interactions in the adsorbed overlayer. LEED results on Pt(111) further indicate that a planar K2O layer may be formed by annealing coadsorbed potassium and oxygen to 750 K.  相似文献   

3.
The adsorption of benzene and naphthalene on the Rh(111) single-crystal surface has been studied by low-energy electron diffraction (LEED), Auger electron spectroscopy (AES) and thermal desorption spectroscopy (TDS). Both benzene and naphthalene form two different ordered surface structures separated by temperature-induced phase transitions: benzene transforms from a (3113) structure, which can also be labelled c(23 × 4)rect, to a (3 × 3) structure in the range of 363–395 K, while naphthalene transforms from a (33 × 33)R30° structure to a (3 × 3) structure in the range 398–423 K. Increasing the temperature further, these structures are found to disorder at about 393 K for benzene and about 448 K for naphthalene. Then, a first H2 desorption peak appears at about 413 K for benzene and 578 K for naphthalene and is interpreted as due to the occurrence of molecular dissociation. All these phase transitions are irreversible. The ordered structures are interpreted as due to flat-lying or nearly flat-lying intact molecules on the rhodium surface, and they are compared with similar structures found on other metal surfaces. Structural models and phase transition mechanisms are proposed.  相似文献   

4.
5.
Bromine gas adsorbs atomically on W(100) at room temperature to a saturation concentration of θ = 0.88 relative to the surface tungsten atom density (1019 m?2). Below θ ~ 0.4, a c(2 × 2) overlayer is formed. Beyond this a (34√2 × √2)R45° structure is preferred and this saturates at θ = 0.67. Higher surface bromine concentrations result in hexagonal variable compression structures on W(100). The sequence begins w structures on W(100). The sequence begins with a c(4 × 2) coincidence mesh which at higher coverages is compressed in one 〈0,1〉 substrate direction. At certain compressions the overlayer achieves p(5 × 2), c(6 × 2), p(7 × 2) coincident configurations and perhaps c(8 × 2) at saturation. This would correspond to θ = 0.875 and is the closest coincidence structure to a perfect hcp overlayer. Bromine prefers a rectangular overlayer geometry on W(100) and compression into an hexagonal array greatly reduces the overlayer stability. The nn repulsions incurred limit room temperature adsorption as the overlayer compresses to perfect hep. Halogen behaviour on W(100) is compared with that on Fe(100). Most differences can be explained in terms of geometrical and bond strength differences but chlorine on W(100) appears to be an exception to this rule.  相似文献   

6.
We present a detailed LEED study of the Ge(111) surface. The low-temperature structure is c(2×8). This structure undergoes a first order transition to an incommensurate reconstruction at 300°C. The diffraction from the high-temperature phase is consistent with either of two simple models. One is a 2 × 1 reconstruction modulated by antiphase walls which are oriented primarily perpendicular to the undoubled direction of the 2 × 1 unit mesh. The second is a 2 × 2 reconstruction modulated by a honeycomb arrangement of intersecting antiphase walls produced by shearing the 2 × 2 mesh by one primitive translation vector. Fluctuations in the I(2 × 2) structure preserve the total wall length and number of wall intersections; the free energy of this structure should be dominated by the entropy. The wall separation and correlation lengths for the high-temperature phase decrease continuously with increasing temperature above the transition.  相似文献   

7.
The Mg(0001) face is subjected to oxygen adsorption from 0 to 103 L. Three characteristic stages of oxygen adsorption are detected from 0 to 10 L. The AES signal of clean Mg decays exponentially against exposure with slopes α ai such that αA2 (0.75 → 3 L) >αA1 (0 → 0.75 L)>αA3 (3 → 10 L). For increasing exposures, they correspond to: (1) a clear (1 × 1)-Mg(0001), (2) a diffuse (1 × 1)-Mg(0001) and (3) a (1 × 1) with a weaker (1 × 1)-R30°-MgO(111) LEED patterns, respectively. At the end of the third stage, a supplementary (7 × 72)?R19°?MgO(111) pattern is observed. In ELS, a very fast intensity decrease of energy loss peaks due to surface and bulk plasmon excitations of the clean metal is recorded during the first stage. The energy loss peak due to the oxidized surface plasmon excitation reaches a maximum intensity at the end of the second stage. Energy loss peaks to be attributed to excitations in bulk MgO appear during the third stage. The work function of the sample decreases and shows a minimum around 6 L, and then slowly increases. Beyond 10 L, a logarithmic relation between oxide thickness and exposure seems to exist. These results are interpreted by the following sequential processes: stage 1: random oxygen chemisorption followed by oxygen incorporation (αA1); stage 2: assembling into islands and lateral island growth (αA2); stage 3: oxide formation (αA3) and stage 4: oxide thickening. Lattice models describing these processes are proposed and discussed. The influence of surface roughness on the results is emphasized.  相似文献   

8.
S. Mróz  A. Mróz 《Surface science》1981,109(2):444-450
Measurements of the surface Debye temperature for the (110) nickel face were performed in such a manner that difficulties in the interpretation known from the literature seem to have been avoided. A specular beam could reach a fluorescent screen for incidence angles 30° < β < 70°. A strong (00) beam was observed for (β ≈ 60° and electron energy E = 100–110 eV. For this beam, the dependence of its intensity on the sample temperature was measured and the effective Debye temperature was calculated to be equal to not more than 216 ± 10 K. Owing to the large β, the penetration depth of electrons should be equal to about 1.2 Å and only the first layer of atoms is responsible for diffraction in this case. Besides, it is shown that single scattering processes are dominant in our experimental conditions. So, the effective Debye temperature measured should be very close to the surface Debye temperature.  相似文献   

9.
CuSi(111) interfaces have been examined as a function of substrate temperature and deposition time using different surface sensitive techniques. At room temperature a layer by layer growth was found. Strong intermixing of Cu and Si takes place and beyond 10 monolayers an ordered α Si phase in Cu was found. At high temperature, the growth follows the layer plus island or Stranski-Krastanov growth mode. Three-dimensional crystallites, probably consisting of Cu3Si, grow epitaxially on an ordered intermediate layer.  相似文献   

10.
We have deposited two monolayers of Sn onto Rh(111) single crystal. After the deposition, no ordered structure was revealed by low energy electron diffraction (LEED). We oxidized the obtained system in a low-pressure oxygen atmosphere at 420 K. The oxidized sample was then gradually heated to study the thermal stability of the oxide layer. We characterized the system by synchrotron radiation stimulated photoelectron spectroscopy and LEED. Valence band and core level photoelectron spectra of rhodium, tin and oxygen were used to study the oxidation of the Sn-Rh(111) surface and its behaviour upon annealing. A low stoichiometric oxide of Sn was created on the surface. The oxidation process did not continue towards creation of SnO(2) with higher oxygen dose. The annealing at 970 K caused decomposition of the surface oxide of Sn and creation of an ordered (√3 × √3)R30° Sn-Rh(111) surface alloy.  相似文献   

11.
The interaction between single crystalline Fe0.84Cr0.16 (100) or (110) and oxygen gas in the pressure range 10?9 to 10?7 torr was studied at room temperature and at 800 K, using LEED and AES. The interaction starts with a chromium-oxygen reaction next to the alloy surface, followed by an iron—oxygen reaction outside the chromium-oxygen layer. At 800 K these reactions are connected with redistribution of cations between the interior of the alloy and the surface region, whereas at room temperature only a redistribution of cations within the surface region is observed. Different symmetries and lattice parameters of oxides which grow epitaxially on Fe0.84Cr0.16 (100) are compared with the corresponding surface compositions. It is found that the formation of spinel-like oxide layers is favoured by lower values of the Cr/Fe surface ratio.  相似文献   

12.
Auger electron spectroscopy (AES), low energy electron diffraction (LEED) and work function (Kelvin probe) measurements have been used to study the initial interaction of clean Al(111), (100) and (110) surfaces with oxygen at room temperature. The oxidation process was found to be surface orientation dependent, but a common feature has been always observed on the three low-index surfaces: they show two distinct phases, i.e. a chemisorbed phase followed then by an oxidized phase. From analysis of AES, LEED and Kelvin probe results, an adsorption mechanism of O on Al for each surface orientation is proposed.  相似文献   

13.
The condensation of gold onto clean and contaminated, single crystal, tantalum (100) surfaces has been followed by using LEED and AES. On a contaminated surface gold condenses as crystallites in a (211) surface orientation with some degree of preferred, azimuthal orientation. On a clean surface gold condenses in an ordered overlayer. Up to approximately 34 monolayer the structure conforms to the (1 × 1) tantalum surface. Beyond this, the observed LEED structure may be interpreted initially in terms of a TaAu overlayer made up of 90° rotated domains with (001)TaAu//(100)Ta and 〈 10 〉 TaAu// 〈 11 〉 Ta, and then in terms of a gold overlayer in a “distorted (111)” orientation. Annealing of these gold films always results in the formation of a (1 × 1) TaAu overlayer of small crystallite size.  相似文献   

14.
《Surface science》1986,171(2):303-330
In this paper, we describe the apparatus which allows us to study by LEED and AES the surface of a single crystal electrode before and after it is submitted to voltammetric experiments sheltered from atmospheric air. This apparatus is used to contribute to the study of the old problem of platinum electrode “activation”. Firstly, we determine which voltammogram shape does actually correspond to the first hydrogen electrochemical adsorption-desorption process onto a clean, well ordered Pt(111) surface. Secondly, we study by LEED and AES the changes in the Pt(111) electrode surface structure induced by a few electrochemical oxygen adsorption- desorption cycles.  相似文献   

15.
The Ag induced superstructures on the Si(111) surface have been studied by low energy electron diffraction constant momentum transfer averaging (LEED/CMTA) technique. The vertical displacements of the atoms are determined from the analysis of the specularly reflected (00) beam intensities. Unexpected behavior of the Ag atoms is clarified: For the √3 × √3-Ag surface it is verified that the Ag atoms are embedded in the first double layer of Si, leading to a considerable rearrangement of the substrate. In contrast, for the 3 × 1-Ag surface, the Ag atoms are riding on the Si surface and the reconstruction of the substrate is small.  相似文献   

16.
The {0001} polar surfaces of ZnO single crystals have first been examined after a chemical treatment involving HCl and H3PO4 and a 24 hr bakeout at 250 °C. The impurities detected on the (0001)-O surface with AES were carbon, chlorine, phosphorus and to a lesser extent sulphur. On the (0001)-Zn surface, carbon, chlorine and sulphur were the dominant impurities, while the phosphorus signal was less important. These results were confirmed by XPS measurements on frehsly etched surfaces. The AES spectra were recorded as distribution curves N(E). Averaging, curve-fitting and related numerical techniques were used to obtain high resolution spectra, enabling the identification of the phosphorus L1-transitions. The etched surfaces were cleaned progressively using argon ion bombardment and ohmic heating. It has been consistently observed that the clean surfaces exhibit primitive (1 × 1) structures. Superstructures such as (3 × 3) on the (0001)-O surface, and (43 × 43) and (3 × 3) on the (0001)-Zn surface, were repeatedly observed at discrete spots of contaminated surfaces. A clear correlation with impurities as observed by AES however could not be found. Facetting was observed after prolonged heating.  相似文献   

17.
H. Niehus 《Surface science》1979,87(2):561-580
Oxygen and carbon monoxide adsorption on clean W(111) surfaces have been studied by angular resolved ESD emission (ESDIAD). In addition, the specimen could be characterized in situ with AES and LEED. Adsorption was performed at room temperature. The electron stimulated desorption yielded O+ ions from the two investigated adsorption layers. Upon oxygen adsorption followed by subsequent annealing at least eight different ESDIAD patterns have been obtained. However, a convincing interpretation on the basis of the surface geometry can only be presented for three patterns produced without annealing as well as for one pattern at a very high annealing temperature. The difficulties are a consequence of complex structure changes which the surface undergoes in the intermediate annealing temperature range. This may influence the little known neutralisation probability of the desorbing ions. In this special case ESDIAD probably reflects in contrast to LEED a picture of some specific adsorption sites (minority species) and therefore, no clear correlation of the two techniques can be seen. ESDIAD from carbon monoxide shows four different patterns and supports the model of linear bonded CO molecules at room temperature with oxygen in the “standing up” position. At T > 900 K, CO starts to dissociate and results in similar ESDIAD patterns as obtained from O2 adsorption.  相似文献   

18.
19.
Features of copper sulfide growth on a Cu(111) surface resulting from sulfur impurity segregation were studied by means of AES, REELS and LEED in the range T = 300−1075 K. It was shown that the 2D sulphide formation is a multistage process that consists of the segregant accumulation in atomic similarity state up to a certain critical coverage (depending on temperature), followed by a comparatively rapid formation of 2D-phase domains due to chemical interaction and atomic ordering, and the slow growth of new phase domains up to an equilibrium coverage.  相似文献   

20.
C. Binns  C. Norris 《Surface science》1982,115(2):395-416
LEED, AES, UPS and EELS have been used to observe the growth of thallium in the Stranski-Krastanov mode on copper(100). In the sub-monolayer range, thallium forms three commensurate ordered overlayer structures. The transition from the first overlayer structure to the second involves a transition from a one-dimensional to a two-dimensional structure. The resulting change in co-ordination number reduces the binding energy of the thallium 5d core levels and increases their width.  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