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.
Using a surface ionisation ion microscope the desorption parameters and the diffusion constant of potassium were measured on stepped W(100) surfaces. The activation energy of ionic desorption as well as the corresponding prefactor do not depend on the step density; the mean adsorption lifetime τ can be expressed as τ=1.6×10?14s exp(2.44 eV/kT).Whereas the surface diffusion of potassium on “flat” W(100) and on W(S)-[9(100)×(110)] was found to be isotropic, on W(S)- [5(100)×(110)] and W(S)-[3(100)×(110)] it occurs preferentially parallel to the step direction. The diffusion constant D for this direction has roughly the same value for all investigated surfaces: D=7.8×10?2 cm2s?1 exp(?0.42 eV/kT). For the direction perpendicular to the steps D⊥ depends on the step density, whereby the activation energy as well as the prefactor increase with increasing step density.  相似文献   

2.
J. Goniakowski  C. Noguera   《Surface science》1995,340(3):191-204
The electronic and atomic structure of vicinal MgO surfaces are studied using a quantum self-consistent method associated with a geometry optimization code. 10n, (n + 1)0n and n1n surfaces, with periodic monoatomic steps separating {001} or {101} terraces, are considered. Diatomic steps along the {10n} orientation and periodic kinks on the {3 1 10} surface are also modelled. We assign most electronic peculiarities of stepped surfaces to the values of the Madelung potential acting on the under-coordinated atoms, which is a function of their first and second coordination numbers. An analytical model is proposed to explain the bond contractions around these atoms. Finally the microscopic contributions to the step energy are discussed, together with the strength of the step interaction as a function of their separation.  相似文献   

3.
GaP(001) cleaned by argon-ion bombardment and annealed at 500°C showed the Ga-stabilized GaP(001)(4 × 2) structure. Only treatment in 10?5 Torr PH3 at 500°C gave the P-stabilized GaP(001)(1 × 2) structure. The AES peak ratio PGa is 2 for the (4 × 2) and 3.5 for the (1 × 2) structure. Cs adsorbs with a sticking probability of unity up to 5 × 1014 Cs atoms cm?2 and a lower one at higher coverages. The photoemission measured with uv light of 3660 Å showed a maximum at the coverage of 5 × 1014 atoms cm?2. Cs adsorbs amorphously at room temperature, but heat treatment gives ordered structures, which are thought to be reconstructed GaP(001) structures induced by Cs. The LEED patterns showed the GaP(001)(1 × 2) Cs structure formed at 180°C for 10 h with a Cs coverage of 5 × 1014 atoms cm?2, the GaP(001)(1 × 4) Cs formed at 210°C for 10 hours with a Cs coverage of 2.7 × 1014 atoms cm?2, the GaP(001)(7 × 1) and the high temperature GaP(001)(1 × 4), the latter two with very low Cs content. Desorption measurements show three stability regions: (a) between 25–150°C for coverages greater than 5 × 1014 atoms cm?2, and an activation energy of 1.2 eV; (b) between 180–200°C with a coverage of 5 × 1014 atoms cm?2, and an activation energy of 1.8 eV; (c) between 210–400°C with a coverage of 2.7 × 1014 atoms cm?2, and an activation energy of 2.5 eV.  相似文献   

4.
The surface diffusion of palladium on the curved part of a tungsten crystal is studied by field electron microscopy. The variation of the local coverage distribution is measured by a probe-hole device on the stepped surface region around (001). The measured data allow a determination of the mass transport surface diffusion coefficient D of Pd on W across atomic steps as a function of temperature, coverage and step density. D has been found (1) to be constant for a given step density and for coverages lower than about 5 × 1014 Pd adatoms/cm2, (2) to increase for higher coverages, and (3) to increase with increasing step density for a given coverage. The activation energy of the process is nearly constant (about 24 kcal/mol) for all coverages up to about 6 × 1014 adatoms/cm2, while the pre-exponential factor of D increases with increasing step density. Interpretation of the results gives some information on the diffusion mechanism.  相似文献   

5.
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   

6.
The influence of surface defects on the adsorption of CO by rhenium is investigated using LEED, AES and linear temperature programmed desorption. On both surfaces, thermal desorption reveals two adsorption states, the lower temperature α state being resolved into two substates, and one β state, all desorbing with first order kinetics. The α state is unaffected by the surface texture, its maximum population being the same on both surfaces, around 4 × 1014 molecules cm?2, similar to the value found for poly crystalline rhenium. On the other hand, the β state is strongly dependent on surface structure. On Re(0001) a maximum of 4 × 1013 molecules cm?2 was found, and 2 × 1014 molecules cm?2 on the stepped surface. The adsorption is activated and can be increased, by heating to 550 K, to 2 × 1014 molecules cm?2 on the basal plane and 3.5 × 1014 molecules cm?2 on the stepped surface. Ordered structures are now seen in LEED. Comparison of these results with previous results from polycrystalline rhenium indicate that the dissociation of β-CO on the latter surface must occur at defects other than steps.  相似文献   

7.
Large anisotropy effects have been observed for the spreading of Pd on W single crystal planes exhibiting regular step structures. The spreading rate was studied as function of step density and step direction at temperatures around 1000 K. The terraces had always (110) orientation. Step densities as low as 3.3 × 104cm?1 corresponding to an average terrace width of 3000 Å already cause enhanced spreading rates parallel to the step direction. Steps parallel to the [001] direction showed the largest promoting action. The rate perpendicular to steps does not depend markedly on step density. The diffusion behaviour on the “step free” (110) plane turned out to be isotropic. A model is proposed which assumes the mass transport to occur in the second Pd layer. The increase of the diffusion rate along steps is correlated with the much larger density of diffusing atoms in step sites as compared to terrace sites due to the difference in binding energy.  相似文献   

8.
The in-plane magnetic anisotropy of Fe/NiO bilayers was studied quantitatively as a function of NiO thickness using the magneto-optical Kerr effect with a rotating field. For NiO thicker than the ordering transition thickness, the total in-plane fourfold anisotropy of the Fe layer decreases with NiO thickness in Fe/NiO/Au(001), but increases in Fe/NiO/MgO(001). Our result indicates that the exchange coupling in an Fe/NiO bilayer might induce an additional in-plane fourfold anisotropy, and the opposite thickness dependent behaviors may be attributed to the different Ni2+ antiferromagnetic spin orientations for NiO films grown on Au(001) and MgO(001) surfaces.  相似文献   

9.
By means of a real-space tight-binding calculation we obtain the spin-polarization of vicinal (1, 0, 2n?1) vanadium surface, and of one V layer deposited on vicinal (1, 0, 2n?1) iron surfaces (n=1,2,3,4). These geometries can be viewed as a staircase on the (001) surface with steps of monoatomic height andn-atoms width. Forn≤2, the pure V stepped surfaces do not show magnetism which is consistent with the absence of spinpolarization at the (101) surface. By contrast, magnetism is always obtained when a V monolayer is deposited on Fe stepped substrate, which is due to the hybridization of d-orbitals of V and Fe. Furthermore, in all the cases where magnetism is obtained, the surface V atoms at the edge of the step result to be antiferromagnetically coupled with all kink atoms. The effect of this local defect on the magnetic structure seems to remain when approaching the (001) surface (n→∞).  相似文献   

10.
The dissociative chemisorption of oxygen and water is reported on both (111) and [6(111) × (100)] crystal faces of gold. The oxide formation becomes rapid above 500°C at pressures of about 10?6 torr. The resulting gold oxide is bound strongly. It is similar in structure to the corresponding sulphide and is stable on both surfaces to 800°C in vacum. Ethylene, cyclohexene, n-heptane, benzene did not chemisorb on gold under low pressure conditions on either the (111) or on the stepped gold surface while naphthalene exhibited dissociative chemisorption on both types of surfaces. Hydrocarbon fragments are bound strongly to the gold surface but the activation energy for dissociative adsorption of light hydrocarbon molecules appears to be high.  相似文献   

11.
The copper deposition on single crystal ZnO(101̄0) and MgO(001) surfaces has been studied by electron energy loss spectroscopy (EELS) in UHV at room temperature. The initial deposited Cu (well below one monolayer) induces a loss peak at about 2 eV on both oxide surfaces and at 4.3 eV on the MgO(001) surface. Based upon heat treatment and oxidation experiments the 2 eV structure is assigned to the electronic resonance of Cu(I) from the Cu deposit on the oxide matrix substrates. On the basis of the experiments the colour-center-related loss peaks, at 2.6 eV for MgO(001) and at 1.9 eV for ZnO(101̄0), are believed to be due to electronic resonance of a Vs center, and the metal ion vacancies are suggested to be active centers which interact with the submonolayer copper deposits. Finally, the electronic energy loss spectra from the Cu-covered oxide surfaces are discussed in the framework of electronic band structures.  相似文献   

12.
The adsorption and diffusion of Pd monomers and dimers on the (001) surfaces of strongly correlated nickel oxides were investigated using density functional theory combined with the on-site Coulomb repulsion U. The results were compared with those of Pd on nonmagnetic MgO(001). For the Pd monomer, the most stable adsorption site was found to be near the surface O atom. The surface diffusion of the Pd monomer occurred by a hopping process over surface hollow sites. The diffusion energy barrier was 0.21 eV, which was lower than that for Pd on MgO(001). In the case of the Pd dimer, the smallest and stable cluster, the most stable adsorption structure had a flat geometry, with both Pd atoms sitting above the neighboring surface O atoms. The surface diffusion of the Pd dimer occurred by rotational and sliding processes, in contrast to that of the Pd dimer on MgO(001). The diffusion energy barriers ranged from 0.33 to 0.36 eV. The values for the surface diffusion of Pd dimers on NiO(001) were lower than those of Pd on MgO(001). This suggests that Pd dimers move more rapidly on NiO(001) than on MgO(001), and that the sintering of Pd clusters closely related to catalytic activities can occur more easily compared to that of Pd on MgO(001).  相似文献   

13.
Epitaxially grown GaAs(001), (111) and (1?1?1?) surfaces and their behaviour on Cs adsorption are studied by LEED, AES and photoemission. Upon heat treatment the clean GaAs(001) surface shows all the structures of the As-stabilized to the Ga-stabilized surface. By careful annealing it is also possible to obtain the As-stabilized surface from the Ga-stabilized surface, which must be due to the diffusion of As from the bulk to the surface. The As-stabilized surface can be recovered from the Ga-stabilized surface by treating the surface at 400°C in an AsH3 atmosphere. The Cs coverage of all these surfaces is linear with the dosage and shows a sharp breakpoint at 5.3 × 1014 atoms cm?2. The photoemission reaches a maximum precisely at the dosage of this break point for the GaAs(001) and GaAs(1?1?1?) surface, whereas for the GaAs(111) surface the maximum in the photoemission is reached at a higher dosage of 6.5 × 1014 atoms cm?2. The maximum photoemission from all surfaces is in the order of 50μA Im?1 for white light (T = 2850 K). LEED measurements show that Cs adsorbs as an amorphous layer on these surfaces at room temperature. Heat treatment of the Cs-activated GaAs (001) surface shows a stability region of 4.7 × 1014 atoms cm?2 at 260dgC and one of 2.7 × 1014 atoms cm?2 at 340°C without any ordering of the Cs atoms. Heat treatment of the Cs-activated GaAs(111) crystal shows a gradual desorption of Cs up to a coverage of 1 × 1014 atoms cm?2, which is stable at 360°C and where LEED shows the formation of the GaAs(111) (√7 × √7)Cs structure. Heat treatment of the Cs-activated GaAs(1?1?1?) crystal shows a stability region at 260°C with a coverage of 3.8 × 1014 atoms cm?2 with ordering of the Cs atoms in a GaAs(1?1?1?) (4 × 4)Cs structure and at 340°C a further stability region with a coverage of 1 × 1014 at cm?2 with the formation of a GaAs(1?1?1?) (√21 × √21)Cs structure. Possible models of the GaAs(1?1?1?) (4 × 4)Cs, GaAs(1?1?1?)(√21 × √21)Cs and GaAs(111) (√7 × √7)Cs structures are given.  相似文献   

14.
Direct reflectance, thermoreflectance, and electroreflectance have been measured for MnO, CoO, and NiO above the fundamental edge. Spectra of all three materials support a model containing both localized and one-electron band states. In MnO peaks with temperature coefficients of ~ 10?3 eV/K were observed at 5.7 and 6.9 eV, temperature-independent structure occurred at 5.4, 6.3, and 7.2 eV, and spectral features with indeterminate temperature dependence were seen at 4.6 and 5.5 eV. The temperature-dependent structure was assigned to one-electron interband transitions associated with anion 2p and cation 4s states; the temperature independent structure was assigned to crystal field split localized interionic transitions between the 3d-states of neighboring Mn ions. Thermoreflectance spectra for CoO exhibited temperature dependent structure (9.5 × × 10?4 eV/K at 5.0, 6.0, and 7.2 eV. A strong, temperature dependent electroreflectance oscillation was seen in NiO near 6.2 eV. On the basis of these data the interband gap between the anion 2p and cation 4s bands was determined to be 5.7 eV in MnO, 6.0 eV in CoO, and 6.2 eV in NiO.  相似文献   

15.
The adsorption of CO, O2, and H2O was studied on both the (111) and [6(111) × (100)] crystal faces of iridium. The techniques used were LEED, AES, and thermal desorption. Marked differences were found in surface structures and heats of adsorption on these crystal faces. Oxygen is adsorbed in a single bonding state on the (111) face. On the stepped iridium surface an additional bonding state with a higher heat of adsorption was detected which can be attributed to oxygen adsorbed at steps. On both (111) and stepped iridium crystal faces the adsorption of oxygen at room temperature produced a (2 × 1) surface structure. Two surface structures were found for CO adsorbed on Ir(111); a (√3 × √3)R30° at an exposure of 1.5–2.5 L and a (2√3 × 2√3)R30° at higher coverage. No indication for ordering of adsorbed CO was found on the Ir(S)-[6(111) × (100)] surface. No significant differences in thermal desorption spectra of CO were found on these two faces. H2O is not adsorbed at 300 K on either iridium crystal face. The reaction of CO with O2 was studied on Ir(111) and the results are discussed. The influence of steps on the adsorption behaviour of CO and O2 on iridium and the correlation with the results found previously on the same platinum crystal faces are discussed.  相似文献   

16.
Ba0.5Sr0.5TiO3 (BSTO) films have been grown heteroepitaxially on (001) MgO substrates by pulsed laser deposition (PLD) to fabricate microwave phase shifters for the wide frequency range 45 MHz–50 GHz. Both as-grown and ex situ annealed films have a cube on cube epitaxial relationship with ?100?BSTO//?100?MgO. Threading dislocations are the dominant defects, mostly with Burgers vectors b = ?101?. Growth at 10?1 mbar oxygen pressure, compared to 10?4 mbar, resulted in significantly better properties. Ex situ annealing of the film grown at 0.1 mbar resulted in a reduction of 40% in threading dislocation density and a 40% increase in dielectric tunability.  相似文献   

17.
The microhardness of single-crystal samples of ZnSe: Cr2+ with a chromium concentration in the range from 3.3 × 1017 to 4.0 × 1019 cm?3 has been studied. The microhardness as a function of the load on the indenter on the faces (111), (1 $\bar 1$ 0), and (001) of the ZnSe: Cr2+ and ZnSe samples has been measured. It has been established that doping of zinc selenide with chromium leads to a decrease in the anisotropy of the mechanical properties and stabilization of the cubic sphalerite structure.  相似文献   

18.
Abstract

MgO implanted at room temperature with 150keV In+ ions and doses ranging from 1015 to 1017 ions cm?2 was studied by optical absorption and transmission electron microscopy (TEM). Creation of defects in the anionic sublattice (F-, F+-, F2-centers) and in the cationic sublattice (V?-centers) are observed. Subsequent annealings of the implanted crystals have shown different behaviours depending on the implanted dose. For medium dose (2 × 1016 ions cm?2), the formation of In3+ species seems to be the preponderant phenomenon. At higher dose (8 × 1016 ions cm?2), metallic precipitates are formed between 400 and 700°C. The identification of these precipitates has been achieved using TEM: they are formed of a metallic alloy Mg3In with a hexagonal structure and their orientation relationship with respect to the MgO matrix is: (0001)Mg3In//(111)Mgo and [1120]Mg3In// [l10]MgO.  相似文献   

19.
This paper reports structural, optical and cathodoluminescence characterizations of sintered Zn1?x Mg x O composite materials. The effects of MgO composition on these film properties have been analyzed. X-ray diffraction (XRD) confirms that all composites are polycrystalline with prominent hexagonal wurtzite structure along two preferred orientations (002) and (101) for the crystallite growth. Above doping content x = 10 %, the formation of the hexagonal ZnMgO alloy phase and the segregation of the cubic MgO phase start. From reflectance and absorption measurements, we determined the band gap energy which tends to increase from 3.287 to 3.827 eV as the doping content increases. This widening of the optical band gap is explained by the Burstein–Moss effect which causes a significant increase of electron concentration (2.89 × 1018?5.19 × 1020 cm?3). The luminescent properties of the Zn1?x Mg x O pellets are studied by cathodoluminescence (CL) at room and liquid nitrogen temperatures under different electron beam excitations. At room temperature, the CL spectra of the Zn1?x Mg x O composites exhibit a dominant broad yellow-green light band at 2.38 eV and two ultraviolet emission peaks at 3.24 and 3.45 eV corresponding to the luminescence of the hexagonal ZnO and ZnMgO structures, respectively. For the doped ZnO samples, it reveals also new red peaks at 1.72 and 1.77 eV assigned to impurities’ emissions. However, the CL spectra recorded at 77 K show the presence of excitonic emission peaks related to recombination of free exciton (X A), neutral donor-bound excitons (D0X) and their phonon replicas. The CL intensity and energy position of the green, red and ultraviolet emission peaks are found to depend strongly on the MgO doping content. The CL intensity of the UV and red emissions is more enhanced than the green light when the MgO content increases. CL imaging analysis shows that the repartition of the emitting centers in Zn1?x Mg x O composites is intimately connected to the film composition and surface morphology.  相似文献   

20.
A kinetic study of D2 formation from HCOOD decomposition on Ni(110) (2 × 1)C was performed using the flash desorption technique. The surface structure and surface composition were monitored by low energy electron diffraction (LEED) and Auger electron spectroscopy (AES). Flash curves were obtained using initial coverage and heating rate variations. D2 formation exhibited a single second-order rate-determining step. Three different techniques were employed in obtaining the activation energy, two of which did not require the assumption of reaction order. Using an average value of 12.6 kcal/mole for the activation energy the pre-exponential factor was calculated to be 2.7 × 10?4 cm2 molecules?1 sec?1. Good agreement was achieved with the theoretically generated second-order flash curves only up to the peak temperature. The discrepancy on the high temperature side was explained using the model proposed by Clavenna and Schmidt utilizing a coverage dependent pre-exponential factor.  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