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.
The methanol decomposition and oxidation on a Pd(111) single crystal have been investigated in situ using ambient-pressure X-ray photoelectron spectroscopy (XPS) and mass-spectrometry (MS) in the temperature range of 300–600 K. It was found that even in the oxygen presence the methanol decomposition on palladium proceeds through two competitive routes: fast dehydrogenation to CO and H2, and slow decomposition of methanol via the C–O bond scission. The rate of the second route is significant even in the millibar pressure range, which leads to a blocking of the palladium surface by carbon and to a prevention of the further methanol conversion. As a result, no gas phase products of methanol decomposition were detected by mass-spectrometry at 0.1 mbar CH3OH in the whole temperature range. The methanol C–O bond scission produces CHx species, which fast dehydrogenate to atomic carbon even at room temperature and further partially dissolve in the palladium bulk at 400 K with the formation of the PdCx phase. According to in situ XPS data, the PdCx phase forms even in the oxygen excess. The application of an in situ XPS–MS technique unambiguously shows a good correlation between a decrease in the surface concentration of all carbon-containing species and the rate of methanol conversion. Since these carbon species have a high reactivity towards oxygen, heating of Pd(111) above 450 K in a methanol–oxygen mixture yields CO, CO2, and water. The product distribution indicates that the main route of methanol conversion is the dehydrogenation of methanol to CO and hydrogen. However, under the experimental conditions used, hydrogen is completely oxidized to water, while CO is partially oxidized to CO2. No palladium oxide was detected by XPS in these conditions.  相似文献   

2.
The adsorption and reaction behaviors of CF3CH2I on Ag(111) were systematically studied by density functional theory (DFT) calculations. Physical adsorption of CF3CH2I on Ag(111) occurs due to the weak interactions between surface Ag atoms and iodine atom of CF3CH2I; while strong chemisorption occurs for CF3CH2 fragment on Ag(111). Electronic analysis indicates that the singly occupied molecular orbital (SOMO) of CF3CH2 strongly interacts with the surface Ag atoms. It is very interesting to find that the most stable structures of CF3CH2 on Ag(111) locate at the top site, instead of the hollow sites. This might be attributed to the facts that CF3CH2 adsorbed at the top site will maximize the sp3-type hybridization, and the possible weak interaction between the fluorine lone pair electrons of p orbitals for CF3CH2 and surface Ag(111) occurs, which is supported by the charge density difference (CDD) analysis with a low isosurface value. We propose that the charge density difference (CDD) analysis with a high or low isosurface value can be widely applied to analyze the strong or weak electronic interactions upon adsorption. Transition state calculations suggested that the energy barrier of CF bond rupture for CF3CH2I on Ag(111) (1.44 eV) is much higher than that of CI bond breakage for CF3CH2I (0.43 eV); and the activation energy of the CF bond dissociation for CF3CH2(a) is 0.67 eV.  相似文献   

3.
Assem Bakry  Ahmed M. El-Naggar 《Optik》2013,124(24):6501-6505
Phosphorus doped hydrogenated amorphous silicon (a-Si:H) films were prepared by decomposition of silane using RF plasma glow discharge. Both DC dark conductivity measurements, and spectrophotometric optical measurements through the range 200–3000 nm were recorded for the prepared films. The DC conductivity activation energy Ea decreased from 0.8 eV for the undoped sample to 0.34 eV for the highest used doping value. The optical energy gap Eg decreased ranging from 1.66 eV to 1.60 eV. The refractive index n, the density of charge carriers N/m* and the plasma frequency ωp showed an opposite behavior, i.e. an increase in value with doping. Fitting the dispersion values to Sellmeier equation led to the determination of the material natural frequency of oscillating particles. A correlation between the changes in these parameters with the doping has been attempted.  相似文献   

4.
Blue light emitting chromophores have been separated from silica spheres by soaking them into acetone for 120 days. The luminescent chromophores were not obtained from other solvents, including ether, methanol, ethanol, 2-propanol, chloroform and tetrahydrofuran. According to the Fourier transform infrared spectrum, the luminescent material is composed of C–OH, –CH2, –CH3, C=O, and Si–O–Si. UV–visible absorption peak of the chromophore is at 5.17 eV (240 nm). Field emission scanning electron microscope images show small cracks on the surface of aged spheres. The luminescence peak was at 2.81 eV (441 nm) for excitation energy between 3.88 and 3.35 eV and slightly shifted toward lower energy for excitation energy lower than 3.35 eV. The deconvoluted luminescent spectrum shows two emission bands at 3.08 and 2.74 eV, which are well-matched the oxygen deficient center model. Compared to the absorption peak (5.17 eV) and the emission peak (2.81 eV), large Stokes shift (2.36 eV) is observed.  相似文献   

5.
Local defects present in CeO2 ? x films result in a mixture of Ce3+ and Ce4+ oxidation states. Previous studies of the Ce 3d region with XPS have shown that depositing metal nanoparticles on ceria films causes further reduction, with an increase in Ce3+ concentration. Here, we compare the use of XPS and resonant photoemission spectroscopy (RESPES) to estimate the concentration of Ce3+ and Ce4+ in CeO2 ? x films grown on Pt (111), and the variation of this concentration as a function of Pd deposition. Due to the nature of the electronic structure of CeO2 ? x, resonant peaks are observed for the 4d–4f transitions when the photon energy matches the resonant energy; (hν = 121.0 eV) for Ce3+ and (hν = 124.5 eV) for Ce4+. This results in two discrete resonant photoemission peaks in valence band spectra. The ratio of the difference of these peaks with off-resonance scans gives an indication of the relative contribution of Ce3+. Results from RESPES indicate reduction of CeO2 ? x on deposition of Pd, confirming earlier findings from XPS studies.  相似文献   

6.
S. ?zkaya  M. ?akmak  B. Alkan 《Surface science》2010,604(21-22):1899-1905
The surface reconstruction, 3 × 2, induced by Yb adsorption on a Ge (Si)(111) surface has been studied using first principles density-functional calculation within the generalized gradient approximation. The two different possible adsorption sites have been considered: (i) H3 (this site is directly above a fourth-layer Ge (Si) atom) and (ii) T4 (directly above a second-layer Ge (Si) atom). We have found that the total energies corresponding to these binding sites are nearly the same, indeed for the Yb/Ge (Si)(111)–(3 × 2) structure the T4 model is slightly energetic by about 0.01 (0.08) eV/unitcell compared with the H3 model. In particular for the Ge sublayer, the energy difference is small, and therefore it is possible that the T4, H3, or T4H3 (half of the adatoms occupy the T4 adsorption site and the rest of the adatoms are located at the H3 site) binding sites can coexist with REM/Ge(111)–(3 × 2). In contrast to the proposed model, we have not determined any buckling in the Ge = Ge double bond. The electronic band structures of the surfaces and the corresponding natures of their orbitals have also been calculated. Our results for both substrates are seen to be in agreement with the recent experimental data, especially that of the Yb/Si(111)–(3 × 2) surface.  相似文献   

7.
Theoretical calculations focused on the geometry, stability, electronic and magnetic properties of small palladium clusters Pdn (n=1–5) adsorbed on the NiAl(1 1 0) alloy surface were carried out within the framework of density functional theory (DFT). In agreement with the experimental observations, both Ni-bridge and Al-bridge sites are preferential for the adsorption of single palladium atom, with an adsorption energy difference of 0.04 eV. Among the possible structures considered for Pdn (n=1–5) clusters adsorbed on NiAl(1 1 0) surface, Pd atoms tend to form one-dimensional (1D) chain structure at low coverage (from Pd1 to Pd3) and two-dimensional (2D) structures are more stable than three-dimensional (3D) structures for Pd4 and Pd5. Furthermore, metal-substrate bonding prevails over metal–metal bonding for Pd cluster adsorbed on NiAl(1 1 0) surface. The density of states for Pd atoms of Pd/NiAl(1 1 0) system are strongly affected by their chemical environment. The magnetic feature emerged upon the adsorption of Pd clusters on NiAl(1 1 0) surface was due to the charge transfer between Pd atoms and the substrate. These findings may shade light on the understanding of the growth of Pd metal clusters on alloy surface and the construction of nanoscale devices.  相似文献   

8.
《Solid State Ionics》2006,177(13-14):1117-1122
We report a comparative study of transport and thermodynamic properties of single-crystal and polycrystalline samples of the ionic salt CsH5(PO4)2 possessing a peculiar three-dimensional hydrogen-bond network. The observed potential of electrolyte decomposition ≈ 1.3 V indicates that the main charge carriers in this salt are protons. However, in spite of the high proton concentration, the conductivity appears to be rather low with a high apparent activation energy Ea  2 eV, implying that protons are strongly bound. The transport anisotropy though is not large, correlates with the crystal structure: the highest conductivity is found in the [001] direction (σ130 °C 5.6 × 10 6 S cm 1) while the minimal conductivity is in the [100] direction (σ130 °C 10 −6 S cm 1). The conductivity of polycrystalline samples appears to exceed the bulk one by 1–3 orders of magnitude with a concomitant decrease of the activation energy (Ea  1.05 eV), which indicates that a pseudo-liquid layer with a high proton mobility is formed at the surface of grains. Infrared and Raman spectroscopy used to study the dynamics of the hydrogen-bond system in single-crystal and polycrystalline samples have confirmed the formation of such a modified surface layer in the latter. However, no bulk phase transition into the superionic disordered phase is observed in CsH5(PO4)2 up to the melting point Tmelt 151.6 °C, in contrast to its closest relative compound CsH2PO4.  相似文献   

9.
By performing density functional theory calculations, this work clarifies the sites and energetics of both the non-dissociative and dissociated adsorptions of CH3SH on clean Au(1 1 1) and Au(1 1 1) with intrinsic defects. It was found that the adsorption on defect-free Au(1 1 1) is most stable for non-dissociative CH3SH. Its direct molecular dissociation to form CH3S/Au and H/Au is barred by an activation barrier of 0.9 eV. However, the presence of neighboring Auad can assist the dissociation reaction to form CH3S–Auad–H by lowering the energy barrier to 0.6 eV. As for the dissociated CH3S, the surface geometry of two CH3S joined by a Auad is the most favorable one.  相似文献   

10.
The reaction of PdCl2 with anticancer-alkylating agent mechlorethamine hydrochloride (CH3NH(C2H4Cl)2 = HN2 x HCl), in the molar ratio 1:2, affords the complex [CH3NH(C2H4Cl)2]2[PdCl4] ([H2N2]2[PdCl4]). Novel Pd(II) complex and the complex precursor mechlorethamine hydrochloride were tested for their antiradical property. Both present weak interaction with 2, 29-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid) radical cation (ABTS). Assays with soybean lipoxygenase and with superoxide anion radicals in vitro showed very high radical scavenging activity of the complex, whereas the complex precursor mechlorethamine hydrochloride presents lower inhibition. Hydrolytic activity of new complex with N-acetylhistidylglycine (AcHis–Gly) was also studied. It was established that regioselective cleavage of the amide bond of the investigated dipeptide had occurred after heating at 60 °C and at pH = 1.5 for 36 h.  相似文献   

11.
We investigate effects of annealing on GaSb quantum dots (QDs) formed by droplet epitaxy. Ga droplets grown on GaAs are exposed to Sb molecular beam and then annealed at Ta=340–450 °C for 1 min to form GaSb QDs. An atomic force microscope study shows that with the increase of Ta, the average diameter of dots increases by about 60%, while their density decreases to about 1/3. The photoluminescence (PL) of GaSb QDs is observed at around 1 eV only for those samples annealed above Ta=380 °C, which indicates that the annealing process plays an important role in forming high quality GaSb QDs.  相似文献   

12.
The polycrystalline sample of Na1/2Dy1/2TiO3 ceramic was prepared by a standard high-temperature solid-state reaction technique. X-ray structural analysis confirmed the formation of single-phase (with minor secondary phase) compound in the orthorhombic (distorted tetragonal) crystal system at room temperature. Study of surface morphology by scanning electron microscope exhibits uniform distribution of rectangular/cubical grains with less voids. The elemental composition of the prepared compound was confirmed by energy dispersive X-ray spectroscopy microanalysis. Detailed studies of dielectric properties exhibit a dielectric anomaly at 94 °C suggesting a possible ferroelectric–paraelectric phase transition in the compound. The activation energy (Ea), calculated from the temperature dependence of ac conductivity plot, was found to be small (∼0.1 eV) in low temperature and large (∼0.5 eV) in high temperature region.  相似文献   

13.
The interactions of methyl and methylene radicals on Cu(111) were investigated with XPS, AES and HREELS under various exposure conditions. The CH2 and CH3 radicals are generated through a hot nozzle source with ketene and azomethane gases. It is shown that with substrate at 300 K, the impinging CH3 radicals are trapped mainly as CH3(ads), while a part of the adsorbate decomposes to form CH2(ads) and H(ads). H atoms are found to desorb at about 380 K, while the chemisorbed hydrocarbon adspecies desorb at about 420 K. In drastic contrast, exposing the clean Cu surface to methylene radicals results not only in the trapping of CH2(ads), but also in the formation of complex aromatic species. The adlayer is sensitive to annealing at elevated temperatures. Desorption and partial conversion to methylidyne take place at around 420 K. The CH(ads) species can survive up to 700 K and then decomposes to form residual carbon above 800 K. In both radical-Cu(111) systems, surface coverage appears to saturate near one monolayer. The relative concentrations of different surface species in the adlayer, however, depend on the amount of radical exposure. The reaction properties of the two systems are compared and discussed.  相似文献   

14.
Electronic structure of the Ba/3C–SiC(111) interface has been detailed studied in situ in an ultrahigh vacuum using synchrotron radiation photoemission spectroscopy with photon energies in the range of 100–450 eV. The 3C–SiC(111) samples were grown by a new method of epitaxy of low-defect unstressed nanoscaled silicon carbide films on silicon substrates. Valence band photoemission and both the Si 2p, C 1s core level spectra have been investigated as a function of Ba submonolayer coverage. Under Ba adsorption two induced surface bands are found at binding energies of 2 eV and 6 eV. It is obtained that Ba/3C–SiC(111) interface can be characterized as metallic-like. Modification of both the Si 2p and C 1s surface-related components were ascertained and shown to be provided by redistribution effect of electron density between Ba adatoms and both the Si surface and C interface atoms.  相似文献   

15.
Pyrazolo[1,5-a]pyrimidines were synthesized via the ultrasonic sonochemical method using the cyclocondensation reaction of 4-alkoxy-1,1,1-trifluoro-3-alken-2-ones [CF3C(O)CH = C(R)(OMe) – where R = Me, Bu, i-Bu, Ph, 4-Me–C6H4, 4-F–C6H4, 4-Cl–C6H4, 4-Br–C6H4, naphth-2-yl and biphen-4-yl] – with 3-amino-5-methyl-1H-pyrazole in the presence of EtOH for 5 min. This methodology has several advantages, for example, it is a simple procedure, it has an easy work-up, mild conditions, short reaction times (5 min) and produces satisfactory yields (61–98%).  相似文献   

16.
Michael A. Henderson 《Surface science》2010,604(19-20):1800-1807
The photochemical properties of the Cr-terminated α-Cr2O3(0001) surface were explored using methyl bromide (CH3Br) as a probe molecule. CH3Br adsorbed and desorbed molecularly from the Cr-terminated α-Cr2O3(0001) surface without detectable thermal decomposition. Temperature programmed desorption (TPD) revealed a CH3Br desorption state at 240 K for coverages up to 0.5 ML, followed by more weakly bound molecules desorbing at 175 K for coverages up to 1 ML. Multilayer exposures led to desorption at ~ 130 K. The CH3Br sticking coefficient was unity at 105 K for coverages up to monolayer saturation, but decreased as the multilayer formed. In contrast, pre-oxidation of the surface (using an oxygen plasma source) led to capping of surface Cr3+ sites and near complete removal of CH3Br TPD states above 150 K. The photochemistry of chemisorbed CH3Br was explored on the Cr-terminated surface using post-irradiation TPD and photon stimulated desorption (PSD). Irradiation of adsorbed CH3Br with broad band light from a Hg arc lamp resulted in both photodesorption and photodecomposition of the parent molecule at a combined cross section of ~ 10? 22 cm2. Photodissociation of the CH3–Br bond was evidenced by both CH3 detected in PSD and Br atoms left on the surface. Use of a 385 nm cut-off filter effectively shut down the photodissociation pathway but not the parent molecule photodesorption process. From these observations it is inferred that d-to-d transitions in α-Cr2O3, occurring at photon energies < 3 eV, do not significantly promote photodecomposition of adsorbed CH3Br. It is unclear to what extent band-to-band versus direct CH3Br photolysis play in CH3–Br bond dissociation initiated by more energetic photons.  相似文献   

17.
X-irradiation of single crystals of Tp–GeH3 (Tp: triptycene) led to the trapping of the radical Tp–GeH2. The angular variations of the resulting EPR spectra were recorded at 300 and 77 K. The drastic temperature dependence of the spectra was caused by both a strong anisotropy of the g-tensor and a rotation of the GeH2 moiety around the C–Ge bond. The determination of the EPR tensors as well as the analysis of this motion required to take the presence of disorder in the crystal into account. In accordance with DFT calculations, Tp–GeH2 is shown to be pyramidal and to adopt, in its lowest energy structure, a staggered conformation. Rotation around the C–GeH2 bond is blocked at 90 K and is almost free above 110 K. The experimental barrier, obtained after simulation of the EPR spectra as a function of the rotational correlation time, is equal to 1.3 kcal mol−1, which is slightly inferior to the barrier calculated by DFT (3.6 kcal mol−1). Calculations performed on Tp–CH3, Tp–GeH3 and Tp–GeH2 show that the rotation barrier ΔErot around the C–Ge bond drastically decreases by passing from the germane precursor to the germanyl radical and that ΔErot increases by passing from the germane to its carbon analogous. Structural parameters involved in these barrier differences are examined.  相似文献   

18.
High-purity specimens of Li6CaLa2Ta2O12 and Li6BaLa2Ta2O12 have been successfully synthesized by solid-state reactions. The analytical chemical compositions of these samples were in good agreement with the nominal compositions of Li6CaLa2Ta2O12 and Li6BaLa2Ta2O12. The Rietveld refinements verified that these compounds have the garnet-type framework structure with the lattice constants of a = 12.725(2) Å for Li6CaLa2Ta2O12 and a = 13.001(4) Å for Li6BaLa2Ta2O12. All of the diffraction peaks of X-ray powder diffraction patterns were well indexed on the basis of cubic symmetry with space group Ia-3d. To make a search for Li sites, the electron density distributions were precisely examined by using the maximum entropy method. Li+ ions occupy partially two types of crystallographic site in these compounds: (i) tetrahedral 24d sites, and (ii) distorted octahedral 96h sites, the latter of which are the vacant sites of the ideal garnet-type structure. The present Li6CaLa2Ta2O12 and Li6BaLa2Ta2O12 samples exhibit the conductivity σ = 2.2 × 10? 6 S cm? 1 at 27 °C (Ea = 0.50 eV) and σ = 1.3 × 10? 5 S cm? 1 at 25 °C (Ea = 0.44 eV), respectively.  相似文献   

19.
《Solid State Ionics》2006,177(19-25):1985-1989
The application of the electrophoretic deposition (EPD) technique to the preparation of high quality electrolyte films for intermediate temperature solid oxide fuel cells (IT-SOFCs) was investigated. Films of La0.83Sr0.17Ga0.83Mg0.17O2.83 (LSGM) were deposited on Pt and La0.8Sr0.2MnO3 (LSM) substrates from suspensions in acetone/ethanol (3:1 by volume) mixture solvent and sintered at 1300 °C. Pt supported LSGM films, 10–20 μm thick, exhibited good adhesion to the Pt substrate, well-distributed microporosity and some surface roughness. LSM supported films exhibited cracking after sintering at 1300 °C for 3 h. Up to 900 °C the bulk conductivity of the Pt supported LSGM film showed the same behaviour of LSGM pellet (Ea = 0.93 eV and 0.99 eV, respectively). The LSGM film exhibited lower bulk electrical conductivity than the latter (4.1 × 10− 3 and 4.4 × 10− 2 Ω− 1 cm− 1, respectively, at 700 °C). This difference should be ascribed to the slight Ga depletion in the LSGM film. An important issue remains the selection of adequate electrode for LSGM electrolyte films.  相似文献   

20.
Density functional theory calculations have been performed to investigate the structural and electronic properties of bulk Co2C and the stability of low index Co2C surfaces. We found that the formation of Co2C is exothermic with the formation energy of ? 0.81 eV/Co2C with respect to Co under the presence of syngas (mixture of CO and H2). While formed Co2C can be decomposed further to metal Co and graphite carbon with modest energy gain of 0.37 eV/Co2C. This suggests that Co2C is only metastable in Fischer–Tropsch synthesis, which agrees well with experimental findings. The density of states (DOSs) reveals that the Co2C is paramagnetic and strong metallic-like. The difference of charge density analysis indicates that the bond of Co2C is of the mixtures of metallic, covalent, and ionic properties. A variety of low index Co2C surfaces with different terminations are studied. We find that the surface energy of low index stoichiometric Co2C highly relies on the surface area, the number of coordination of surface atoms and the surface dipole, with the decreased stability order of (101) > (011) > (010) > (110) > (100) > (001) = (111). Our results indicate that under Co-poor condition, the formation of non ? stoichiometric surface (011) and (111) without terminated cobalt is energetically more favorable, while under Co-rich condition the formation of non ? stoichiometric (111) surface with cobalt overlayer are preferential.  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