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 71 毫秒
1.
Michael A Henderson   《Surface science》1998,400(1-3):203-219
The reaction of CO2 and H2O to form bicarbonate (HCO3) was examined on the nearly perfect and vacuum annealed surfaces of TiO2(110) with temperature programmed desorption (TPD), static secondary ion mass spectrometry (SSIMS) and high resolution electron energy loss spectrometry (HREELS). The vacuum annealed TiO2(110) surface possesses oxygen vacancy sites that are manifested in electronic EELS by a loss feature at 0.75 V. These oxygen vacancy sites bind CO2 only slightly more strongly (TPD peak at 166 K) than do the five-coordinated Ti4+ sites (TPD peak at 137 K) typical of the nearly perfect TiO2(110) surface. Vibrational HREELS indicates that CO2 is linearly bound at the latter sites with a νa(OCO) frequency similar to the gas phase value. In contrast, oxygen vacancies dissociate H2O to bridging OH groups which recombine to liberate H2O in TPD at 490 K. No evidence for a reaction between CO2 and H2O is detected on the nearly perfect surface. In sequentially dosed experiments on the vacuum annealed surface at 110 K, CO2 adsorption is blocked by the presence of preadsorbed H2O, adsorbed CO2 is displaced by postdosed H2O, and there is little or no evidence for bicarbonate formation in either case. However, when CO2 and H2O are simultaneously dosed, a new CO2 TPD state is observed at 213 K, and the 166 K state associated with CO2 at the vacancies is absent. SSIMS was used to tentatively assign the 213 K CO2 TPD state to a bicarbonate species. The 213 K CO2 TPD state is not formed if the vacancy sites are filled with OH groups prior to simultaneous CO2+H2O exposure. Sticking coefficient measurements suggest that CO2 adsorption at 110 K is precursor-mediated, as is known to be the case for H2O adsorption on TiO2(110). A model explaining the circumstances under which the proposed bicarbonate species is formed involves the surface catalyzed conversion of a precursor-bound H2O–CO2 van der Waals complex to carbonic acid, which then reacts at unoccupied oxygen vacancies to generate bicarbonate, but falls apart to CO2 and H2O in the absence of these sites. This model is consistent with the conditions under which bicarbonate is formed on powdered TiO2, and is similar to the mechanism by which water catalyzes carbonic acid formation in aqueous solution.  相似文献   

2.
The adsorption of D2O on Zr(0001) at 80 K and its subsequent reactions at higher temperatures have been studied by thermal desorption spectroscopy (TDS), work-function measurements (Δф), nuclear reaction analysis (NRA), LEED, infrared reflection spectroscopy (FTIR-RAS), Auger electron spectroscopy (AES), and static secondary ion mass spectroscopy (SSIMS). D2O adsorption on Zr(0001) at 80 K is accompanied by a Δф of −1.33 eV. The adsorbed D2O can be characterized into three layers by TDS: a chemisorbed layer (up to 0.23 ML), a second adsorbed layer, and an ice layer. The chemisorbed D2O dissociates into ODad and Dad at 80 K (possibly also into Oad) and no desorption products could be detected, implying that the reaction products dissolved into the zirconium at temperatures appropriate for each component. The ice layer and most of the second adsorbed layer desorb as molecular water during heating. The water adsorbed at 80 K did not form any long-range ordered structure, but a (2 × 2) LEED pattern that was formed by heating the sample to temperatures above 430 K is believed due to be an ordered oxygen superstructure.  相似文献   

3.
4.
The influence of chemical etching with HF on the nature of the surface of amorphous Ni59Nb40Pt1−xSnx alloys has been studied in situ by electrolyte electroreflectance (EER) and ex situ by X-ray photoelectron spectroscopy (XPS). The EER spectrum of the untreated alloy in 0.5 M H2SO4 shows a bipolar band, which disappears after the HF treatment yielding a structureless EER spectrum similar to that of Pt, but reappears after several hours in the 0.5 M H2SO4 electrolyte. This process of dissolution by HF of an oxide species and its reappearance after a few hours cannot be followed by XPS, since the time interval between sample withdrawal from the electrolyte and actual measurement is of a few hours as well. XPS spectra showed the presence of metallic Nb before and after the HF treatment, and that niobium pentoxide was the main species in the as-quenched alloy, but that after treatment with HF it became a minor component, the main one being NbO. The main effect of the HF treatment is to produce a platinum enrichment of the surface, as unequivocally determined by cyclic voltammetry, XPS and EER. After Ar sputtering for 9 min the XPS spectrum of the untreated alloy showed metallic Nb only, while in the HF-treated alloy the peaks of metallic Nb were swamped by those of NbO and some Nb2O5. We interpret this difference as being due to the formation by the HF attack of a porous Nb film which becomes oxidized in the electrolyte and/or during transfer to the spectrometer, and so thick that it is not eliminated by Ar sputtering for 9 min.  相似文献   

5.
The vibrational spectrum of water dissociatively adsorbed on Si(100) surfaces is obtained with surface infrared absorption spectroscopy. Low frequency spectra (< 1450 cm−1 are acquired using a buried CoSi2 layer as an internal mirror to perform external reflection spectroscopy. On clean Si(100), water dissociates into H and OH surface species as evidenced by EELS results [1] in the literature which show a Si---H stretching vibration (2082 cm−1), and SiO---H vibrations (O---H stretch at 3660 cm−1 and the Si---O---H bend and Si---O stretch of the hydroxyl group centered around 820 cm−1). In this paper, infrared (IR) measurements are presented which confirm and resolve the issue of a puzzling isotopic shift for the Si---O mode of the surface hydroxyl group, namely, that the Si---O stretch of the O---H surface species formed upon H2O exposure occurs at 825 cm−1, while the Si---O stretch of the ---OD surface species formed upon D2O exposure shifts to 840 cm−1, contrary to what is expected for simple reduced mass arguments. The higher resolution of IR measurements versus typical EELS measurements makes it possible to identify a new mode at 898 cm−1, which is an important piece of evidence in understanding the anomalous frequency shift. By comparing the results of measurements for adsorption of H162O, H182O and D2O with the results from recently performed first-principles calculations, it can be shown that a strong vibrational interaction between the Si---O stretching and Si---O---H bending functional group vibrations of the hydroxyl group accounts for the observed isotopic shifts.  相似文献   

6.
The electron-stimulated desorption (ESD) of D and H ions from condensed D2O and H2O films is investigated. Three low-energy peaks are observed in the ESD anion yield, which are identified as arising from excitation of 2B1, 2A1 and 2B2 dissociative electron attachment (DEA) resonances. Additional structure is observed between 18 and 32 eV, which may be due to ion pair formation or to DEA resonances involving the 2a1 orbital. The ion yield resulting from excitation of the 2B1 resonance increases as the film is heated. We attribute the increase in the ion yield to thermally induced hydrogen bond breaking near the surface, which enhances the lifetimes of the excited states that lead to desorption.  相似文献   

7.
A study of deuterium conductivity and diffusion in the oxide perovskite La0.9Sr0.1YO3−δ is presented in this work. Deuterium ions were implanted into La0.9Sr0.1YO3−δ (50 keV, 1×1016 atoms/cm2) and the corresponding deuterium depth profile was determined by SIMS and compared with a Monte Carlo simulation (TRIM96). This implant was used as a standard for the determination of deuterium concentration in a La0.9Sr0.1YO3−δ sample pre-treated in D2O atmosphere. In this way, it was fully confirmed that La0.9Sr0.1YO3−δ incorporates water at high temperatures. The conductivity of La0.9Sr0.1YO3−δ was measured in D2O atmosphere and compared with other proton (deuteron) conductors. Concentration and conductivity data were used in conjunction to estimate the deuterium diffusivity and the constant of reaction of (heavy) water incorporation into LaYO3. Some comments on the catalytic activity of this oxide are made.  相似文献   

8.
The surface chemical reactions of O2 and H2O on clean lithium have been studied by a combination of XPS, EELS and microgravimetry. Reactions with O2 produce a monolayer of oxide which does not passivate the surface and which allows for the growth of several monolayers of additional oxide, probably as a result of the mixing of zero-valent metal into the oxide layer. The reaction of H2O with the clean lithium surface results in the complete dissociation of the molecule and loss of hydrogen to form one monolayer of the oxide. This is followed by the formation of multilayers of hydroxide/oxide mixtures which are shown to be unstable over periods of minutes, converting back to the oxide form predominantly.  相似文献   

9.
N. Saliba  D. H. Parker  B. E. Koel   《Surface science》1998,410(2-3):270-282
Atomic oxygen coverages of up to 1.2 ML may be cleanly adsorbed on the Au(111) surface by exposure to O3 at 300 K. We have studied the adsorbed oxygen layer by AES, XPS, HREELS, LEED, work function measurements and TPD. A plot of the O(519 eV)/Au(239 eV) AES ratio versus coverage is nearly linear, but a small change in slope occurs at ΘO=0.9 ML. LEED observations show no ordered superlattice for the oxygen overlayer for any coverage studied. One-dimensional ordering of the adlayer occurs at low coverages, and disordering of the substrate occurs at higher coverages. Adsorption of 1.0 ML of oxygen on Au(111) increases the work function by +0.80 eV, indicating electron transfer from the Au substrate into an oxygen adlayer. The O(1s) peak in XPS has a binding energy of 530.1 eV, showing only a small (0.3 eV) shift to a higher binding energy with increasing oxygen coverage. No shift was detected for the Au 4f7/2 peak due to adsorption. All oxygen is removed by thermal desorption of O2 to leave a clean Au(111) surface after heating to 600 K. TPD spectra initially show an O2 desorption peak at 520 K at low ΘO, and the peak shifts to higher temperatures for increasing oxygen coverages up to ΘO=0.22 ML. Above this coverage, the peak shifts very slightly to higher temperatures, resulting in a peak at 550 K at ΘO=1.2 ML. Analysis of the TPD data indicates that the desorption of O2 from Au(111) can be described by first-order kinetics with an activation energy for O2 desorption of 30 kcal mol−1 near saturation coverage. We estimate a value for the Au–O bond dissociation energy D(Au–O) to be 56 kcal mol−1.  相似文献   

10.
Z. M. Liu  M. A. Vannice   《Surface science》1996,350(1-3):45-59
The interaction between submonolayer titania coverages and Pt foil has been studied by Auger electron spectroscopy (AES), X-ray photoelectron spectroscopy (XPS), temperature programmed desorption (TPD) and high-resolution electron energy loss spectroscopy (HREELS). The submonolayer titania can be fully oxidized to TiO2 at 923 K under 10−8 Torr O2, and partially oxidized to TiOx at lower oxidation temperatures. The oxidized surface can be reduced by annealing to 1000 K or higher, or by heating in H2 at 823 K, or by interacting with surface carbon formed from acetone decomposition. Under certain conditions (e.g., hydrogen reduction at 923 K), the surface titania can be fully reduced to metallic Ti which diffuses into bulk Pt readily. The reduced metallic Ti can resurface when the surface is oxidized at 923 K. Both XPS and HREELS data indicate the existence of subsurface oxygen, which plays an important role for the diffusion of Ti into and out of the Pt foil. Although no special interfacial active sites were revealed by HREELS studies of adsorbed acetone and CO, some TPD and XPS data suggest the presence of sites active for acetone decomposition.  相似文献   

11.
The structural and ferroelectric characteristics of SrBi2(Nb1−xWx)2O9 (x=0–0.12) ferroelectric ceramics were investigated. SrBi2(Nb1−xWx)2O9 ceramics consisted of a single-phase layered perovskite structure when x was less than 0.06. Uniform microstructure and grain size reduction were observed after the introduction of W. The maximum remanent polarization of 16 μC/cm2 appeared at x=0.03, and the coercive field decreased with increasing concentration of W. The ferroelectric behavior of SrBi2(Nb1−xWx)2O9 ceramics is interpreted based on the Raman measurement.  相似文献   

12.
13.
The oxygen ion conductivity of Y2O3---Nb2O5 with a fluorite-like structure was studied. Substitutional solid solutions of Nb2 O5 in Y2O3 lattice formed the defect fluorite phase and remarkably enhanced the oxygen ion conductivity. Doping with tetravalent cations, especially Ti4+ or Ce4+, in yttria-niobia oxide is effective in enhancing the oxygen ion conductivity. Although the n-type semiconducting property appeared below PO2 = 10−18 atm at 1243 K, the yttria-niobia mixed oxide doped with Ce4+, Ti4+, and Zr4+ stably exhibited oxygen-ion conduction in the wide range of oxygen partial pressures studied.  相似文献   

14.
We studied reaction of oxygen atoms with D-terminated Si(1 1 1) surfaces from a desorption point of view. As the D (1 ML)/Si(1 1 1) surface was exposed to O atoms D2 and D2O molecules were found to desorb from the surface. The desorption kinetics of D2 and D2O molecules exhibited a feature characterized with a quick rate jump at the very beginning of O exposure, which was followed by a gradual increase with a delayed maximum and then by an exponential decrease. The O-induced D2 desorption spectra as a function of Ts appeared to be very similar to the H-induced D2 desorption spectrum from the D/Si(1 1 1) surfaces. Possible mechanisms for the O-induced desorption reactions were discussed.  相似文献   

15.
Adsorption of N2 and N2O at various sites on Ni(7 5 5) has been investigated by density functional theory (DFT) method (periodic DMol3). Several possible adsorption structures (attaching the nitrogen atom to the surface, or lying parallel) are found for both molecules. There is a clear binding energy preference of N2 and N2O for step sites in contrast to the case of CO. It is revealed that the decomposition of N2O occurs exclusively near the step, but not on the terrace. Two decomposition channels can be considered; dissociative adsorption and spontaneous decomposition during TPD ramp. Three possible candidates for the precursor of the spontaneous decomposition of N2O during TPD ramp are discussed.  相似文献   

16.
High-Tc superconducting thin films have been deposited in situ by means of a plasma assisted metal-organic chemical vapour deposition (PAMOCVD) process on LaAlO3. An EMCORE high-speed rotating disc reactor was used to deposit the films at a substrate temperature of 600°C to 800°C. The system is equipped with a (remote) 120 W microwave plasma generator. The oxidising plasma gas is N2O and/or O2 while Ar was used as the inert carrier gas for the different metal-organics. The influence of different process parameters (such as the temperatures of the metal-organics, substrate temperature, and plasma gas composition) on the superconductive properties and on the morphology of the films was investigated. Surface morphology and composition were studied by SEM/EDX or EPMA, and AC susceptibility measurements were used to investigate the superconductive properties (Tc and Jc). X-ray diffraction measurements indicated that single-phase YBa2Cu3O7−x films were epitaxially grown with the 00l orientation perpendicular to the substrate surface. The critical temperature (Tc) of the films is about 90 K and the critical current density (Jc) is higher than 106 A/cm2 at 77 K and zero field.  相似文献   

17.
The investigations on the properties of HfO2 dielectric layers grown by metalorganic molecular beam epitaxy were performed. Hafnium-tetra-tert-butoxide, Hf(C4H9O)4 was used as a Hf precursor and pure oxygen was introduced to form an oxide layer. The grown film was characterized by X-ray photoelectron spectroscopy (XPS), atomic force microscopy (AFM), high-resolution transmission electron microscopy (HRTEM), and capacitance–voltage (CV) and current–voltage (IV) analyses. As an experimental variable, the O2 flow rate was changed from 2 to 8 sccm while the other experimental conditions were fixed. The XPS spectra of Hf 4f and O 1s shifted to the higher binding energy due to the charge transfer effect and the density of trapped charges in the interfacial layer was increased as the oxygen flow rate increased. The observed microstructure indicated the HfO2 layer was polycrystalline, and the monoclinic phases are the dominant crystal structure. From the CV analyses, k = 14–16 and EOT = 44–52 were obtained, and the current densities of (3.2–3.3) × 10−3 A/cm2 were measured at −1.5 V gate voltage from the IV analyses.  相似文献   

18.
Oxygen tracer diffusion (D*) and surface exchange rate constant (k*) have been measured, using isotopic exchange and depth profiling by secondary ion mass spectrometry (SIMS), in La1−xSrxFe0.8Cr0.2O3−δ (x=0.2, 0.4 and 0.6). Measurements were made as a function of temperature (700–1000 °C) and oxygen partial pressure (0.21–10−21 atm) in dry oxygen, water vapour and water vapour/hydrogen/nitrogen mixtures. At high oxygen activity, D* was found to increase with increasing temperature and Sr content. The activation energies for D* in air are 2.13 eV (x=0.2), 1.53 eV (x=0.4) and 1.21 eV (x=0.6). As the oxygen activity decreases, D* increases as expected qualitatively from the increase in oxygen vacancy concentration. Under strongly reducing conditions, the measured values of D* at 1000 °C range from 10−8 cm2 s−1 for x=0.2 to 10−7 cm2 s−1 for x=0.4 and 0.6. The activation energies determined at constant H2O/H2 ratio are 1.21 eV (x=0.2), 1.59 eV (x=0.4) and 0.82 eV (x=0.6).

The surface exchange rate constant of oxygen for the H2O molecule is similar in magnitude to that for the O2 molecule and both increase with increasing Sr concentration.  相似文献   


19.
J.-W. He  P.R. Norton   《Surface science》1990,230(1-3):150-158
The co-adsorption of oxygen and deuterium at 100 K on a Pd(110) surface has been studied by measurements of the change in work function (Δφ) and by thermal desorption spectroscopy (TDS). When the surface with co-adsorbed species is heated, the adsorbates O and D react to form D2O which desorbs from the surface at T > 200 K. The D2O desorption peaks shift continuously to lower temperatures as the surface D coverage (θD) increases. The maximum production of D2O is estimated to be 0.26 ML (1 ML = 9.5 × 1014 atoms cm−2), resulting from reaction in a layer containing 0.65 ML D and 0.3 ML O. The maximum work function increase caused by adsorption of D to saturation onto oxygen precovered Pd(110) decreases almost linearly with ΔφO of the oxygen precovered surface. On a surface with pre-adsorbed D however, the maximum Δφ increase contributed by oxygen adsorption decreases abruptly at ΔφD > 200 mV. This sharp change occurs at θD > 1 ML and is believed to be associated with the development of the reconstructed (1 × 2) phase of D/Pd(110).  相似文献   

20.
Nd2CuO4±δ is the non-superconducting prototype of the Re2−xMxCuO4ty family (Re=Pr, Nd, Sm and M=Ceor Th) of n-type oxide superconductors. Four-probe DC conductivity, EMF in P(O2) gradient, and thermopower measurements have been used to characterise its electric transport and defect structure between 300 and 900°C and between 5×10−4 and 1 atm oxygen partial pressure.

The results show that Nd2CuO4±δ can be oxygen under-stoichiometric (with n-type conductivity), near-stoichiometric, and over-stoichiometric (with p-type conductivity) in different T, P(O2) ranges.  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