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 46 毫秒
1.
The structure and NO reactivity of Zr-deposited Pd surfaces were investigated by X-ray photoelectron spectroscopy, low-energy electron diffraction, infrared reflection absorption spectroscopy, and temperature-programmed desorption. Zr on Pd(1 0 0) was oxidized to ZrO2 by exposure to O2 at 773 K. Heating at 823 K in a vacuum led to decomposition of ZrO2 to Zr metal and O2. The activation energy for ZrO2 decomposition changed remarkably at ΘZr = 0.4. For ΘZr > 0.4, a hexagonal structure was observed for ZrO2/Pd(1 0 0); no ordered structure was observed for ΘZr < 0.4. Deposited Zr had no significant effect on the adsorption and decomposition of NO on Pd(1 0 0) but resulted in a creation of new NO dissociation sites on Pd(3 1 1). Zr on Pd(3 1 1) was oxidized to ZrOX by oxygen produced from NO dissociation. Heating at 823 K in a vacuum led to decomposition of ZrOX to Zr metal and O2.  相似文献   

2.
The nature of the interaction of isocyanic acid (HNCO) with the active centers at the ideal anatase TiO2 (1 0 1) surface were studied using ab initio density functional theory (DFT) method with a cluster model. Two types of adsorption of isocyanic acid are found to be likely at (1 0 1) surface – dissociative and molecular adsorption. Only molecular adsorption of HNCO leads to the direct weakening and further splitting of the NC bond, which is a necessary step for the hydrolysis of isocyanic acid. During molecular adsorption of HNCO, an energetically stable intermediate surface complex is created with an adsorption energy of −1.33 eV, in which the HNCO skeleton is changing due to new strong bonds between C–Os and N–Tis. Based on the existence of this intermediate complex a probable reaction pathway for the hydrolysis of HNCO over the ideal anatase (1 0 1) surface was developed. A surface oxygen vacancy was formed after the decomposition of the intermediate complex and CO2 desorption. Afterwards, water adsorbs at the oxygen vacancy site and NH3 is successively formed. The HNCO hydrolysis over TiO2 was found to be energetically favorable with global energy gain of about −2.08 eV.  相似文献   

3.
The electrical conductivity, optical and metal–semiconductor contact properties of the MEH-PPV:C70 organic semiconductor have been investigated. The electrical conductivity results show that the MEH-PPV:C70 film is an organic semiconductor. The optical band gap of the film was found to be 2.06 eV and the fundamental absorption edge in the film is formed by the direct allowed transitions. The refractive index dispersion curve of the film obeys the single oscillator model and Ed and Eo dispersion parameters were found to be 10.61 and 3.89 eV, respectively. The electrical characterization of the ITO/MEH-PPV:C70 diode have been investigated by current–voltage characteristics. ITO/MEH-PPV:C70 diode indicates a non-ideal current–voltage behavior with ideality factor n (2.50) and barrier height φB (0.90 eV) values.  相似文献   

4.
Diffusion of dysprosium on the (1 1 1) facet of a tungsten micromonocrystal was investigated by means of spectral analysis of field emission current fluctuations. The experimental spectral density functions of the current fluctuations were analysed by using Gesley and Swanson’s theoretical spectral density function, which enables to determine the surface diffusion coefficient D for dysprosium. Derived from the temperature dependence of D, the diffusion activation energy E is presented for some Dy coverages θ(1 1 1). In the temperature range 400–600 K, the E first drops from 1.25 eV per atom at θ(111)≈0.25 ML to 0.48 eV per atom at θ(111)≈1 ML (corresponding to the minimum of the work function of the system), then increases to 1.03 eV per atom at θ(111)≈1.3 ML. The results are discussed from the aspects of the substrate structure and interaction in the adsorbed layer.  相似文献   

5.
The adsorption and thermal decomposition of C2H2 on Rh{111} is compared to the atomically stepped Rh{331} surface over a temperature range of 300 to 800 K. Using X-ray photoelectron spectroscopy (XPS) we find that the C 1s spectra as a function of C2H4 exposure exhibit a shift in binding energy (Eb) from 283.5 eV at 1 L C2H4 exposure on both surfaces to 283.8 eV on Rh{33 and to 284.1 eV on Rh{111} at saturation coverage (4 L). Careful analysis of the C 1s Eb value and full width at half maximum as a function of surface temperature after a 10 L exposure of C2H4 at 300 K reveals that a species consistent with a C2H adsorbate composition is formed between 400 and 450 K on Rh{111}. This species is also observed on Rh{331} although at the lower temperature of 375 K. Computer peak deconvolution of the C 1s spectra between 500 and 700 K suggests that a CHads or Cads surface fragment is formed and increases in concentration at the expense of the C2H species as the surface temperature increases. Above 750 K a graphite overlayer is formed on both surfaces. This overlayer, however, exhibits a low degree of carbon π-character bonding on Rh{331}. The adsorption and decomposition mechanisms suggest that the 300 K C2H4 adsorbate on Rh{331} is ethylidyne and that the stepped surface is more thermally reactive than the flat Rh{111} surface.  相似文献   

6.
The low work-function ZrO/W(100) surface was examined with the aim of understanding the reducing mechanism of the work function. Low-energy electron diffraction (LEED) was employed to analyze the surface atomic arrangement, and X-ray photoelectron spectroscopy (XPS) was used to identify the surface chemical condition. The ZrO/W(100) surface was made as follows: (i) around three monolayers of Zr were deposited on a clean W(100) surface, (ii) the sample was heat treated in an oxygen ambience of 1.3x10−5 Pa for several tens of minutes at 1500 K, and (iii) the sample was flash heated at 2000 K in ultrahigh vacuum (UHV). During heat treatment in O2, the deposited Zr was oxidized to ZrO2, and the LEED pattern formed was p(2×1). The work function increased to 5.3 eV. Subsequent flash heating in UHV changed the p(2×1) LEED pattern into a c(4×2) pattern, and transformed ZrO2 into the so-called Zr–O complex, the oxidized level of which is between ZrO2 and metallic Zr. A drastic decrease in the work function to 2.7 eV ensued. The angular dependence of XPS showed that the Zr–O complex segregated within a few monolayers at the surface.  相似文献   

7.
Superconductivity reported at 110 K, and recently at 150–170 K, in the infinite-layered (SrxCa1−x)CuO2 system is investigated by means of the full potential linear muffin-tin orbital (FLMTO) method. As in other high-Tc cuprates, the electronic structure of the parent compounds, CaCuO2 and SrCuO2, and of the separately calculated composition Sr0.7Ca0.3CuO2 using the experimental lattice parameters, displays strong 2D features including a low density of states at EF (lower than in the other cuprates) and a simple 2D Fermi surface (rounded square) with strong nesting due to a single hybridized Cu d-O p band. As in La2CuO4, a major van Hove saddle-point singularity exists near EF. The drastic changes of the Fermi surface when Ca/Sr vacancies shift the Fermi energy to the van Hove singularity may have a strong influence on the superconducting properties of the compounds and indicate the need for Sr/Ca vacancies in inducing the high Tc.  相似文献   

8.
U. Myler  K. Jacobi 《Surface science》1989,220(2-3):353-367
The Si(113) surface of a p-type sample was studied by AES, LEED and ARUPS. On the clean surface, a (3 × 2) and a (3 × 1) LEED pattern coexist for a large range of annealing temperatures. Annealing to 900 K results in (3 × 1), while temperatures higher than 1050 K favour the (3 × 2) superstructure. ARUPS reveals two weakly dispersing surface resonances around 0.9 and 2.6 eV below EF which are connected with the (3 × 2) and (3 × 1) structures, respectively. The work function was determined as φ = 4.81 eV and the photoionization threshold as ξ = 5.36 eV. The bands are bent downwards by 0.43 eV at room temperature.  相似文献   

9.
用密度泛函理论的总能计算研究了金属铜(100)面的表面原子结构以及氮原子的c(2×2)吸附状态.研究结果表明:在Cu(100) c(2×2)-N表面系统中,氮原子处于四度配位的空洞(FFH)位置,距离最表面铜原子层的垂直距离为0.20?,最短的Cu—N键长度为1.83?.结构优化的计算否定了被吸附物导致的表面再构模型,即c(2×2)元胞的两个铜原子在垂直于表面方向发生相对位移,一个铜原子运动到氮原子之上的模型.该吸附表面的功函数约为4.65eV, 氮原子的平均吸附能为4.92 eV(以孤立氮原子为能量参考点).计算结果还说明,Cu—N杂化形成的表面局域态的位置在费米面以下约1.0 eV附近出现,氮原子和第一层以及第二层铜原子均有不同程度的杂化作用.该结果为最近有关该表面的STM图像的争论提供了判据性的第一性原理计算结果. 关键词: Cu(100) c(2×2)-N 表面吸附态 密度泛函总能计算  相似文献   

10.
X. -C. Guo  R. J. Madix   《Surface science》2004,550(1-3):81-92
The adsorption of oxygen and carbon dioxide on cesium-reconstructed Ag(1 1 0) surface has been studied with scanning tunneling microscopy (STM) and temperature programmed desorption (TPD). At 0.1 ML Cs coverage the whole surface exhibits a mixture of (1 × 2) and (1 × 3) reconstructed structures, indicating that Cs atoms exert a cooperative effect on the surface structures. Real-time STM observation shows that silver atoms on the Cs-covered surface are highly mobile on the nanometer scale at 300 K. The Cs-reconstructed Ag(1 1 0) surface alters the structure formed by dissociative adsorption of oxygen from p(2 × 1) or c(6 × 2) to a p(3 × 5) structure which incorporates 1/3 ML Ag atoms, resulting in the formation of nanometer-sized (10–20 nm) islands. The Cs-induced reconstruction facilitates the adsorption of CO2, which does not adsorb on unreconstructed, clean Ag(1 1 0). CO2 adsorption leads to the formation of locally ordered (2 × 1) structures and linear (2 × 2) structures distributed inhomogeneously on the surface. Adsorbed CO2 desorbs from the Cs-covered surface without accompanied O2 desorption, ruling out carbonate as an intermediate. As a possible alternative, an oxalate-type surface complex [OOC–COO] is suggested, supported by the occurrence of extensive isotope exchange between oxygen atoms among CO2(a). Direct interaction between CO2 and Cs may become significant at higher Cs coverage (>0.3 ML).  相似文献   

11.
A. Kis  K. C. Smith  J. Kiss  F. Solymosi   《Surface science》2000,460(1-3):190-202
The adsorption and dissociation of CH2I2 were studied at 110 K with the aim of generating CH2 species on the Ru(001) surface. The methods used included X-ray photoelectron spectroscopy (XPS), ultraviolet photoelectron spectroscopy (UPS), temperature programmed desorption (TPD), Auger electron spectroscopy (AES) and work function measurements. Adsorption of CH2I2 is characterized by a work function decrease (0.96 eV at monolayer), indicating that adsorbed CH2I2 has a positive outward dipole moment. Three adsorption states were distinguished: a multilayer (Tp=200 K), a weakly bonded state (Tp=220 K) and an irreversibly adsorbed state. A new feature is the formation of CH3I, which desorbs with Tp=160 K. The adsorption of CH2I2 at 110 K is dissociative at submonolayer, but molecular at higher coverages. Dissociation of the monolayer to CH2 and I proceeded at 198–230 K, as indicated by a shift in the I(3d5/2) binding energy from 620.6 eV to 619.9 eV. A fraction of adsorbed CH2 is self-hydrogenated into CH4 (Tp=220 K), and another one is coupled to di-σ-bonded ethylene, which — instead of desorption — is converted to ethylidyne at 220–300 K. Illumination of the adsorbed CH2I2 initiated the dissociation of CH2I2 monolayer even at 110 K, and affected the reaction pathways of CH2.  相似文献   

12.
The adsorption of NO on single gold atoms and Au2 dimers deposited on regular O2− sites and neutral oxygen vacancies (Fs sites) of the MgO(1 0 0) surface have been studied by means of DFT calculations. For Au1/MgO the adsorption of NO is stronger when the Au atom is supported on an anionic site than when it is on a Fs site, with adsorption binding energies of 1.1 and 0.5 eV, respectively. In the first case the spin density is mainly concentrated on the metal atom and protruding from the surface. In such a way, an active site against radicals such as NO is generated. On the Fs site, the presence of the vacancy delocalizes the spin into the substrate, weakening its coupling with NO. For Au2/MgO, as this system has a closed-shell configuration, the NO molecules bonds weakly with Au2. Regarding the N–O stretching frequencies, a very strong shift of 340–400 cm−1 to lower frequencies is observed for Au1/MgO in comparison with free NO.  相似文献   

13.
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   

14.
Impedance spectroscopy was used to study the oxygen reaction kinetics of La0.8Sr0.2MnO3 (LSM)-based electrodes on Y2O3-stabilized ZrO2 (YSZ) electrolytes. Three types of electrodes were studied: pure LSM, LSM–YSZ composites, and LSM/LSM–YSZ bilayers. The electrodes were formed by spin coating and sintering on single-crystal YSZ substrates. Measurements were taken at temperatures ranging from 550 to 850°C and oxygen partial pressures from 1×10−3 to 1 atm. An arc whose resistance Rel had a high activation energy, Ea=1.61±0.05 eV, and a weak oxygen partial pressure dependence, (PO2)−1/6, was observed for the LSM electrodes. A similar arc was observed for LSM–YSZ electrodes, where Rel(PO2)−0.29 and the activation energy was 1.49±0.02 eV. The combination of a high activation energy and a weak PO2 dependence was attributed to oxygen dissociation and adsorption rate-limiting steps for both types of electrodes. LSM–YSZ composite cathodes showed substantially lower overall interfacial resistance values than LSM, but exhibited an additional arc attributed to the resistance of YSZ grain boundaries within the LSM–YSZ. At 850°C and low PO2, an additional arc was observed with size varying as (PO2)−0.80 for LSM and (PO2)−0.57 for LSM–YSZ, suggesting that diffusion had become an additional rate limiting step. Bilayer LSM/LSM–YSZ electrodes yielded results intermediate between LSM and LSM–YSZ. The results showed that most of the improvement in electrode performance was achieved for a LSM–YSZ layer only ≈2 μm thick. However, a decrease in the grain-boundary resistance would produce much better performance in thicker LSM–YSZ electrodes.  相似文献   

15.
Zn1−xMnxS epilayers were grown on GaAs (1 0 0) substrates by hot-wall epitaxy. X-ray diffraction (XRD) patterns revealed that all the epilayers have a zincblende structure. The optical properties were investigated using spectroscopic ellipsometry at 300 K from 3.0 to 8.5 eV. The obtained data were analyzed for determining the critical points of pseudodielectric function spectra, (E) = 1(E) + i2(E), such as E0, E0 + Δ0, and E1, and three E2 (Σ, Δ, Γ) structures at a lower Mn composition range. These critical points were determined by analytical line-shapes fitted to numerically calculated derivatives of their pseudodielectric functions. The observation of new peaks, as well as the shifting and broadening of the critical points of Zn1−xMnxS epilayers, were investigated as a function of Mn composition by ellipsometric measurements for the first time. The characteristics of the peaks changed with increasing Mn composition. In particular, four new peaks were observed between 4.0 and 8.0 eV for Zn1−xMnxS epilayers, and their characteristics were investigated in this study.  相似文献   

16.
The oxidation of CoGa(1 0 0) at 700 K was studied by means of high resolution electron energy loss spectroscopy (EELS), scanning tunneling microscopy, low energy electron diffraction and Auger electron spectroscopy (AES). At 700 K, thin well-ordered β-Ga2O3 films grow on CoGa(1 0 0). The EEL spectrum of the Ga-oxide films exhibit Fuchs–Kliewer phonons at 305, 455, 645, and 785 cm−1. For low oxygen exposure (<0.2 L), the growth of oxide-islands starts at step edges and on defects. The oxide films have the shape of long, rectangular islands and are oriented in the [1 0 0] and [0 1 0] directions of the substrate. For higher oxygen exposure, islands of β-Ga2O3 are found also on the terraces. After an exposure of 200 L O2 at 700 K, the CoGa(1 0 0) surface is homogeneously covered with a thin film of β-Ga2O3.  相似文献   

17.
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   

18.
Zirconium oxide (ZrO2) is one of the leading candidates to replace silicon oxide (SiO2) as the gate dielectric for future generation metal-oxide-semiconductor (MOS) based nanoelectronic devices. Experimental studies have shown that a 1–3 monolayer SiO2 film between the high permittivity metal oxide and the substrate silicon is needed to minimize electrical degradation. This study uses density functional theory (DFT) to investigate the initial growth reactions of ZrO2 on hydroxylated SiO2 by atomic layer deposition (ALD). The reactants investigated in this study are zirconium tetrachloride (ZrCl4) and water (H2O). Exchange reaction mechanisms for the two reaction half-cycles were investigated. For the first half-reaction, reaction of gaseous ZrCl4 with the hydroxylated SiO2 surface was studied. Upon adsorption, ZrCl4 forms a stable intermediate complex with the surface SiO2–OH* site, followed by formation of SiO2–O–Zr–Cl* surface sites and HCl. For the second half-reaction, reaction of H2O on SiO2–O–Zr–Cl* surface sites was investigated. The reaction pathway is analogous to that of the first half-reaction; water first forms a stable intermediate complex followed by evolution of HCl through combination of a Cl atom from the surface site and an H atom from H2O. The results reveal that the stable intermediate complexes formed in both half-reactions can lead to a slow film growth rate unless process parameters are adjusted to lower the stability of the complex. The energetics of the two half-reactions are similar to those of ZrO2 ALD on ZrO2 and as well as the energetics of ZrO2 ALD on hydroxylated silicon. The energetics of the growth reactions with two surface hydroxyl sites are also described.  相似文献   

19.
The behavior of zirconium atoms at the W(100) surface associated with oxygen adsorption at different sample temperatures has been studied by Auger electron spectroscopy (AES), ion scattering spectroscopy (ISS), and the relative change of the work function (Δф) measured by the onset of the secondary electron energy distribution. The results have revealed: (i) adsorption of zirconium onto the W(100) surface followed by the elevation of the sample temperature up to 1710 K in an oxygen partial pressure of 2.7 × 10−4 induces complete diffusion of zirconium atoms into the W(100) substrate; (ii) further exposure of oxygen induces co-existence of oxygen and tungsten on the surface at 1710 K, resulting in a work function of 4.37 eV; (iii) keeping the sample temperature at 1710 K, simple evacuation of the system has resulted in surface segregation of zirconium atoms to the surface to form a zirconium atomic layer on the top-most surface, reducing the work function to 2.7 eV. The results have revealed that this specific behavior of zirconium atoms at high temperature assures, with very good reproducibility, the highly stable performance and long service life of Zr---O/W(100)-emitters in practical use, even in a low vacuum of 10−6 Pa.  相似文献   

20.
The technique of polarity reversal of the external electric extraction field (strength: 102 V/cm) was applied to study the relaxation of the thermal ion emission from the KCl(0 0 1) single crystal surface. Transient currents of the K+ and K2Cl+ ions upon switching from the emission suppression to the ion extraction mode were recorded as a function of the evaporation time, the temperature, and the time of field reversal. The temperature dependence of the time constants of the K+ ions obtained from the exponential decreases of the emission currents to their steady-state emission resulted as logτh(s)=−(13.39±0.56)+(12.42±0.49)103/T in a high temperature interval of 826–930 K after a prolonged heating period and as logτl(s)=−(20.65±1.04)+(16.77±0.81)103/T in a low temperature interval of 750–801 K at the initial stage of evaporation, with corresponding activation energies of Eh(K+)=2.47±0.14 eV and El(K+)=3.32±0.16 eV, respectively. The transient currents can be interpreted by a partial adsorption of the suppressed ion currents at the kinks of the surface steps. The differences in the high- and low-temperature runs may be attributed to a strong coarsening of the surface at higher temperatures, which occurs as a bunching of monosteps to macrosteps and/or to an enrichment and segregation of divalent impurities at the surface. The transient behavior of the molecular K2Cl+ ions seems to be strongly correlated with that of the K+ ions. This correlation is possibly caused by changes of the strength or the sign of the local electrical field connected with the excess charge at the kinks.  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