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 875 毫秒
1.
The mechanisms for defect formation stimulated by the adsorption of water molecules in the surface of YBa2Cu3O7 ceramic are studied, together with the types of defects and their distributions. It is found that a water layer physically bound to the surface reduces the rates of annihilation and capture of positrons, changes the amount of barium and copper on the surface by a factor of two, and inhibits diffusive jumps of nickel atoms. A layer of adsorbed water excites subthreshold formation of 1021 cm−3 interstitial Ba and Cu1 atoms and transitions of oxygen from O1 to O5, and vice versa in the volume of crystallites, and the migration of defects and accumulation of Ba atoms in the surface layer, which block diffusive jumps of Ni within the volume of the crystals. These effects are related to the excitation of collective, low-frequency weakly damped motion of heavy holes in the crystal volume when defects are formed on the surface by physically adsorbed H2O molecules, which is accompanied by Coulomb repulsion of cations from intermediate layers into interstitials and the migration of defects in the field of the collective excitations. Zh. éksp. Teor. Fiz. 116, 586–603 (August 1999)  相似文献   

2.
Pyrite (FeS2) oxidation during coal combustion is one of the main sources of harmful SO2 emission from coal-fired power plants. Density functional theory (DFT) study was performed to uncover the evolution mechanism of SOx formation during pyrite oxidation. The results show that chemisorption mechanism is responsible for O2, SO2 and SO3 adsorption on FeS2 surface. The presence of formed oxidation layer (Fe2O3) weakens the interaction between O2 molecule and FeS2 surface. The adsorbed O2 molecule easily dissociates into active surface O atom for SOx formation. The dissociation reaction of O2 is activated by 77.38?kJ/mol, and exothermic by 138.46?kJ/mol. Compared to the further oxidation of SO2 into SO3, SO2 prefers to desorb from FeS2 surface. The dominant reaction pathway of SO2 formation from the oxidation of the outermost FeS2 surface layer is a three-step process: surface lattice S oxidation, SO2 desorption and replenishment of S vacancy by activated surface O atom. The elementary reaction of surface lattice S oxidation has an activation energy barrier of 197.96?kJ/mol, and is identified as the rate-limiting step. SO2 formation from the further oxidation of bulk FeS2 layer is controlled by a four-step process: bulk lattice S migration, lattice S oxidation, SO2 desorption and surface O atom deposition. Migration of lattice S from bulk position to the outermost surface shows a high activation energy barrier of 175.83?kJ/mol. The deposition process of surface O atom is a relatively easy step, and is activated by 21.05?kJ/mol.  相似文献   

3.
This paper approaches the evolution of adsorption layer with temperature under different water concentrations. EDX and X-ray diffraction (XRD) indicated that under low water concentration (LWC), there was a sharp decrease in TiO2 quantity curve and grain size curve between 40 °C and 60 °C. But under high water concentration (HWC), TiO2 quantity and grain size both rose continually. In transmission electron microscopy (TEM) picture, the characteristic of samples of LWC and HWC at low temperature were similar, whereas they became quite different at high temperature. These phenomena manifest that the effects of adsorption and reaction alter at different conditions. Under LWC, adsorption mechanism dominates the process. Under HWC, the chief mechanism at low temperature was similar to that of LWC and at high temperature the major reaction region switches from adsorption to alcohol bulk. Based on the characteristics of adsorption on silica surface, the phenomenology theory dealt with formation of adsorption layer and its evolution with temperature was proposed.  相似文献   

4.
Oxygen adsorption on the Si(110) surface has been studied by Auger electron spectroscopy. For a clean annealed surface chemisorption occurs, with an initial sticking probability of ~6 × 10?3. In this case the oxygen okll signal saturates and no formation of SiO2 can be detected from an analysis of the Si L2,3VV lineshape. With electron impact on the surface during oxygen exposure much larger quantities are adsorbed with the formation of an SiO2 surface layer. This increased reactivity towards oxygen is due to either a direct effect of the electron beam or to a combined action of the beam with residual CO during oxygen inlet, which creates reactive carbon centers on the surface. Thus in the presence of an electron beam on the surface separate exosures to CO showed adsorption of C and O. For this surface subsequent exposure in the absence of the electron beam resulted in additional oxygen adsorption and formation of SiO2. No adsorption of CO could be detected without electron impact. The changes in surface chemistry with adsorption are detectable from the Si L2,3VV Auger spectrum. Assignments can be made of two main features in the spectra, relating to surface and bulk contributions to the density of states in the valence band.  相似文献   

5.
The interaction of the Cr(111) surface with O2 was studied by means of X-ray and UV photoemission and also work function measurements. A strong oxygen adsorption was found even at very low exposures, suggesting a high sticking coefficient. Previous treatments of the clean surface such as argon-ion bombardment or annealing result in significant changes of the surface structure reflected on work function and adsorption kinetics. No work function change was observed in the initial stage of adsorption, ruling out a model of chemisorption on top. In this range the sticking coefficient remains also constant, supporting a model of rapid regeneration of the genuine surface sites and incorporation of oxygen into the lattice. But in contrast with non transition metals like Cs or Sr, oxygen absorbed at room temperature in Cr, remains essentially in the topmost layers of the surface. At room temperature this initial stage of oxygen incorporation is followed by chemisorption on the corrosion film obtained when the uppermost layers are saturated with oxygen. The oxide layer has a stoichiometry close to Cr2O3 at saturation, but the detailed electronic structure depends on previous thermal treatments. Exposures at room temperature lead to a thin (about 9 Å), probably amorphous corrosion layer with a maximum work function change Δφ = +0.9 eV. Adsorption followed by heating at 500° C results in a much thicker corrosion film with a limiting work function decrease of Δφ = ?1.2 eV. The XP and UP spectra differ significantly in both cases and suggest a Fermi level shift of nearly 1 eV connected with oxygen adsorption on the Cr2O3 surface. The thickness of the corrosion film may be further increased by heating at 500°C in oxygen. The usual XPS spectra of bulk chromium sesquioxide are then clearly observed.  相似文献   

6.
The chemisorption and subsequent reaction of bromine on Cr(110 has been studied by Auger spectroscopy, LEED, Δφ, and thermal desorption measurements. For gas doses of < 7.5 × 1018 molecules m?2, very efficient dissociative chemisorption leads to a series of well-ordered, out-of-registry compression structures. Uniquely, however, the overlayer falls back into registry at saturation coverage; at this point the appearance of glide symmetry indicates that the three-fold coordinated adsorption sites are occupied exclusively. Brominemetal charge transfer occurs during adsorption (in contrast to Cr(100)). On raising the temperature at low coverages, the surface phase decomposes by evaporation as CrBr molecules; at higher coverages the desorption product switches to CrBr2. Continuous growth of bulk CrBr2 sets in at high gas exposures, this corrosion reaction proceeding at a rate which is ten times slower than the rate of overlayer formation. The chromium dibromide layer also evaporates as CrBr2(g). Structural relationships with related metal-halogen systems are discussed.  相似文献   

7.
IR-reflection spectroscopy (IRS) under grazing incidence and UV-photoemission (UPS) with Hel and Hell radiation were used to study the adsorption of D2O and H2O, respectively, on Pt(111) under UHV conditions. In IRS three different vibrational bands in the OD stretch region yield information about the orientation of hydrogen-bonded water molecules and the formation of water clusters. Initially formed water multimers grow to clusters and finally build up multilayers. From UPS the adsorption of molecular water in the whole coverage range is confirmed. Both UPS and IRS show that the water molecules in the “first layer” are bound through their oxygen end to the surface.  相似文献   

8.
H2O adsorption on clean Ni(110) surfaces at T ≦ 150 K leads at coverages below θ ? 0.5 to the formation of chemisorbed water dimers, bound to the Ni substrate via both oxygen atoms. The linear hydrogen bond axis is oriented parallel to the [001] surface directions. With increasing H2O coverage (θ ≧ 0.5), the accumulation of further hydrogen bonded water molecules induces some modification of the dimer configuration, producing at θ ? 1 a two-dimensional hydrogen bonded network with a slightly distorted ice lattice structure and long range order.  相似文献   

9.
The state of the surface of amorphous ice with a specific surface area of about 160 m2/g obtained by the condensation of water vapor at 77 K is studied by IR spectroscopy. As the temperature increases to 130–160 K, absorption bands of surface hydroxyl groups vanish, whereas changes in bands characteristic of hydroxyl groups in the bulk of ice are indicative of a phase transition of ice from amorphous to the polycrystalline structure. The surface sites of amorphous ice are characterized with low-temperature adsorption of carbon monoxide. It is shown that there are two types of CO adsorption sites, free hydroxyl groups and oxygen atoms of surface coordinately unsaturated water molecules. Upon adsorption of nitrogen, methane, and carbon monoxide, in addition to the perturbation of surface OH groups, reversible changes in the spectrum are observed in the region of vibrations of bulk hydroxyls, which indicate that the strength of hydrogen bonds between water molecules in the surface layer of icy particles increases approaching the strength of these bonds in the crystal and that the ice surface becomes less amorphous. These results indicate that the properties of the ice surface layer substantially depend on the presence of adsorbed molecules.  相似文献   

10.
The evolution of a polycrystalline magnesium surface during oxidation at room temperature has been studied by Metastable Impact Electron Spectroscopy (MIES). This technique allowed us to follow the metal-to-insulator transformation of the top layer of the surface. An electronic signal corresponding to a metallic behavior of the surface evidences the presence of under-stoichiometric MgO species on the surface. The total covering by oxygen of the Mg surface uppermost layer, obtained at around 10 L of oxygen deposition, does not correspond to a fully insulating surface. An insulating surface is obtained after 30 L of oxygen deposition. Depositions of CO2 on a clean and a preoxidized polycrystalline Mg surface have been analyzed to give information about the composition of the surface and its evolution. CO2 adsorption in the form of CO32− compounds on preoxidized Mg is more efficient than on clean Mg. Oxygen species, corresponding to chemisorbed oxygen less bounded than oxygen in the MgO lattice, allows the formation of CO32−. Therefore, it is concluded that during oxygen deposition at room temperature, MgO islands and chemisorbed oxygen species coexist on the surface. Moreover, the larger the oxygen predeposition is, the less CO32− compounds are formed, meaning a decrease of available chemisorbed oxygen sites. From oxidation measurements at high temperature (420 K), we show that MgO islands and uncovered Mg domain coexist. Further, no under-stoichiometric compound features have been observed. The high temperature allows the direct formation of oxide MgO species in islands.  相似文献   

11.
本文利用程序升温脱附技术(TPD)研究了乙醛吸附在锐钛矿型TiO2(001)-(1×4)表面的化学性质. 实验结果表明完整晶格位点对乙醛反应表现极为惰性,而表面上的还原型缺陷位点在热驱动下可有效地使乙醛分子通过碳-碳偶联反应生成2-丁酮和丁烯. 提出了乙醛在锐钛矿型TiO2(001)-(1×4)表面偶联反应主要是通过表面还原型缺陷位吸附成对的乙醛分子,因为表面已有的钛原子对还原型缺陷为乙醛分子提供了合适的吸附位点.  相似文献   

12.
《Solid State Ionics》1988,27(4):221-225
The protonic conductivity of three acid hydrates, HUO2PO4·4H2O, Ce(HPO4)2·nH2O,V2O5·nH2O has been studied as a function of the partial pressure of water. The evolution of their conductivity is related to experimental water adsorption isotherms and to the theoretical Brunauer adsorption isotherms. Conclusions can then be drawn about the origin of the conductivity (bulk-type or surface conductivity) and about the type of water molecules involved in the protonic diffusion process. These three compounds are each representative of a different class of protonic conductors, the lattice hydrates, the particle hydrates, and the swelling hydrates respectively.  相似文献   

13.
Ag vapor deposited on Rh(100) is investigated as a possible model for the corresponding bimetallic cluster catalyst. Ag overlayer growth follows the Frank-Van der Merwe mechanism at 300 K and the Stranski-Krastanov mechanism at 640 K. The first Ag overlayer grows epitaxially with the Rh(100) surface and forms two-dimensional islands. The second layer reconstructs forming a pseudohexagonal lattice structure with a 19% greater Ag atom density than the first layer. Neither alloy formation nor dissolution of Ag into the Rh crystal lattice is observed. The presence of Ag decreases the capacity of the Rh(100) surface for both D2 and CO chemisorption. In both cases, loss of adsorption sites on the Rh surface is attributed to physical site blocking by islands of Ag. Very weakly bound CO adsorption sites are observed which are attributed to CO bonding to Ag atoms in the first Ag overlayer. The Ag/Rh(100) system is found to be very similar to the analogous SiO2 supported catalyst.  相似文献   

14.
Results from structural and morphological studies, measurements of the sheet electrical resistance, and estimating resistivity ρm of a graphite-like conducting surface layer formed upon high-dose irradiation of the (111) face of a synthetic diamond with Ar+ ions at an energy of 30 keV and a target temperature of 400°C are presented. It is found that the orienting effect of the diamond lattice is visible in the suppression of the formation of graphite crystallites with axis c perpendicular to the surface. The thickness of the modified layer is 40–50 nm, and its sheet resistance is 0.5 kΩ/sq. Resistivity ρm = 20–25 μΩ m of the modified layer lies within the range of ρ values of graphite and glassy carbon materials.  相似文献   

15.
The adsorption and condensation of benzene on ZnO(101&#x0304;0) was investigated by thermal desorption spectroscopy and LEED. The first monolayer shows an ordered c(2 × 2) super-structure. First order desorption is observed. The desorption energy and frequency factor decrease from 73 to 56 kJ mole?1 and from ~1015 to ~1012 s?1, respectively, with coverage increasing to 0.85. The second layer is more weakly bound. Two-dimensional (2D) island formation is deduced from peak shape analysis. Near completion, the second layer converts to a more tightly bound configuration as deduced from a sudden shift of the desorption peak and the formation of an additional c(4 × 3) LEED pattern. This pattern which can be identified as a property of bulk benzene is preserved upon epitaxial growth of the 3D benzene crystal. Angular resolved UPS measurements indicate the benzene molecules of the first layer to be arranged in an oblique position of low symmetry.  相似文献   

16.
The balance between water-metal interactions and water-water hydrogen bonding (HBs) controls the process of water adsorption on metallic surfaces. In other hand, the yield of oxygen evolution reaction (OER) is dependent on the binding energy of H2O at electrode surface. Therefore, on a specific metal substrate, attenuation of HBs may be a promising route for improving OER. In this study, the computational and experimental evidences indicate that the performance of ultrasonically irradiated deionized water (USI-DW), participated in water oxidation reaction (WOR), is different from its in the intact bulk water. To date, establishing of new electrocatalysts with lower overpotentials (η) and higher current densities (J) in OER have been mostly considered based on metals and oxide materials. Here, we ultrasonically agitated the water clusters formed by strong HBs, and as a sustainable improvement route explored its particular effects on the efficiency of OER. The molecular modeling (MM) of the (H2O)n clusters (n = 1–100 molecules), the corresponding IR spectra, the molecular orbitals energy levels and the adsorption of free and cluster confined H2O molecules on the Pt surface were studied by the appropriate quantum mechanical (QM) methods. The result of deconvolution of FTIR spectra recorded for USI-DW in the –OH stretching region (∼2600–3900 cm−1) properly confirmed the expected increase of the single water molecules. The reduction in overpotentials was 82 ± 8 mV and 158 ± 12 mV, to reach the J of 1 mA cm−1 at the typical pHs 12.2 and 13.1, respectively.  相似文献   

17.
A study is reported of the temperature dependences of the hyperfine (HF) interaction parameters in a ~200-nm thick surface layer and in the bulk of macroscopic hexagonal ferrite crystals of the Sr-M type (SrFe12O19 and SrFe10.2Al1.8O19). The method used for the measurements is Mössbauer spectroscopy with simultaneous detection of gamma quanta, characteristic x-ray emission, and electrons, which permits direct comparison of the HF parameters in the bulk and the near-surface layers of a sample. As follows from the experimentally determined temperature dependences of the effective magnetic fields, the fields at the nuclei of the iron ions located in a ~200-nm thick near-surface layer decrease with increasing temperature faster than those of the ions in the bulk. The transition to paramagnetic state in a ~200-nm thick surface layer was found to occur 3° below the bulk Curie temperature. This offers the first experimental evidence for the transition to paramagnetic state in a surface layer of macroscopic ferromagnets to take place below the Curie temperature T c for the bulk of the crystal. It has been established that the transition temperature T c (L) of a thin layer at a depth L from the surface of a crystal increases as one moves away from the surface to reach T c at the inner boundary of the surface layer called critical. In the vicinity of T c one observes a nonuniform state, with the crystal being magnetically ordered in the bulk but disordered on the surface. The experimental data obtained were used to construct a phase diagram of surface and bulk states for macroscopic magnets near the Curie (or Néel) temperature.  相似文献   

18.
H. Niehus 《Surface science》1983,130(1):41-49
Low-energy He+ ion scattering (IS) in combination with AES, LEED and work function measurements has been applied for the determination of surface reconstruction of Cu(111) upon oxygen adsorption. IS data clearly indicate that oxygen is not significant incorporated into the bulk at room temperature adsorption, however the surface shows reconstruction by displacement of Cu atoms by 0.3 Å. The disappearance of structure of both CuIS and OIS in the “?in pattern” demonstrate the development of a disordered layer of reconstruction centres. At saturation coverage, a rough and dis-ordered oxygen-copper surface layer is present.  相似文献   

19.
Constant-area and fully relaxed molecular dynamics methods are employed to study the properties of the surface and point defects at and near {001} surfaces of bulk and thin-film Ni, Al and Ni3Al respectively. The surface tension is larger than the surface energy for all {001} surfaces considered in the sequence: Al (1005?mJ?m?2)<?Ni3Al (mixed Ni–Al plane outermost, 1725?mJ?m?2)<?Ni3Al (all-Ni-atoms plane outermost, 1969?mJ?m?2)<?Ni (1993?mJ?m?2). For a surface of bulk Ni3Al crystal with a Ni–Al mixed plane outermost, Al atoms stand out by 0.0679?Å compared with the surface Ni atoms and, for the all-Ni-atoms surface, Al atoms in the second layer stand out by 0.0205?Å compared with Ni atoms in the same layer. Vacancy formation energies are about half the bulk values in the first layer and reach a maximum in the second layer where the atomic energy is close to the bulk value but the change in embedding energy of neighbouring atoms before and after vacancy formation is greater than that in the bulk. Both the vacancy formation energy and the surface tension suggest that the fourth layer is in a bulk state for all the surfaces. The formation energy of adatoms, antisite defects and point-defect pairs at and near {001} surfaces of Ni3Al are also given.  相似文献   

20.
The thermodynamic properties of the adsorption of xenon on the stepped Pd(s)[8(100)×(110)] surface have been studied over a wide range of pressure (5×10?11 to 1×10?4 Torr) and temperature (40–140 K). We have measured adsorption isobars using AES in order to evaluate the surface coverage. By choosing pressure and temperature we have studied under equilibrium conditions, the successive adsorption of xenon on the steps and on the terraces until the first layer is formed, the condensation of the second layer as well as the formation of xenon multilayers. For a small range of pressure and temperature, adsorption takes place only on the atomic steps. The LEED pattern shows that only every other site along the steps is occupied. The extrapolated initial heat of adsorption for steps is EadS = 10.2 kcal/mol, decreasing monotonically by about 2 kcal/mol as the relative coverage of the step sites increases. The dipole moment of the Xe atoms adsorbed on steps is 1.12 D. During adsorption on the terraces the LEED observations suggest that the xenon adlayer is non-localized up to completion of the hexagonally close packed monolayer. The initial heat of adsorption on the terraces, EadT is 8.2 kcal/mol and decreases continuously to a value of 6.9 kcal/mol for a complete monolayer due to lateral repulsive interactions between the adsorbed xenon atoms. The induced dipole moment of Xe on terraces is reduced to 0.49 D. The 5p12 binding energy of Xe adsorbed on terrace sites is 0.3 eV smaller than that of Xe occuping step sites. The differential molar entropy of the adsorbed layer on the terraces as a function of coverage compares fairly well with the calculated value for an ideally mobile two-dimensional gas. No indication of the growth of two-dimensional xenon islands has been found under these conditions. The isosteric heat of adsorption for the second layer is Eadsec = 5.8 kcal/mol independently of the coverage. The condensation of the second layer is a first order two-dimensional gas ? two-dimensional solid phase transition in opposition to the continuous nature of the adsorption of the first layer (extending over a wide range of temperature for a given pressure). The induced dipole moment is further reduced for the Xe second layer to a value of 0.11 D. Finally, the condensation of multilayers proceeds with a latent heat of transformation of Econd = 3.8 kcal/mol in excellent agreement with the known bulk value for the heat of sublimation of xenon. The line shape of the NVV low energy Auger transitions of xenon or the UPS binding energies of the Xe 5p32,12 spectra allow a clear distinction between first, second and higher layer Xe atoms. We have also established the temperature/pressure conditions for equilibrium between first, second and bulk xenon layers, i.e. a so-called “roughening point”.  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