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 439 毫秒
1.
The interaction of water vapour with clean as well as with oxygen precovered Ni(110) surfaces was studied at 150 and 273 K, using UPS, ΔΦ, TDS, and ELS. The He(I) (He(II)) excited UPS indicate a molecular adsorption of H2O on Ni(110) at 150 K, showing three water-induced peaks at 6.5, 9.5 and 12.2 eV below EF (6.8, 9.4 and 12.7 eV below EF). The dramatic decrease of the Ni d-band intensity at higher exposures, as well as the course of the work function change, demonstrates the formation of H2O multilayers (ice). The observed energy shift of all water-induced UPS peaks relative to the Fermi level (ΔEmax = 1.5 eVat 200 L) with increasing coverage is related to extra-atomic relaxation effects. The activation energies of desorption were estimated as 14.9 and 17.3 kcal/mole. From the ELS measurements we conclude a great sensitivity of H2O for electron beam induced dissociation. At 273 K water adsorbs on Ni(110) only in the presence of oxygen, with two peaks at 5.7 and 9.3 eV below EF (He(II)), being interpreted as due to hydroxyl species (OH)δ? on the surface. A kinetic model for the H2O adsorption on oxygen precovered Ni(110) surfaces is proposed, and verified by a simple Monte Carlo calculation leading to the same dependence of the maximum amount of adsorbed H2O on the oxygen precoverage as revealed by work function measurements. On heating, some of the (OH)δ? recombines and desorbs as H2O at ? 320 K, leaving behind an oxygen covered Ni surface.  相似文献   

2.
A. Spitzer  H. Lüth 《Surface science》1985,160(2):353-361
The adsorption of H2O on clean and oxygen precovered Cu(110) is studied at temperatures between 90 and 300 K by XPS. On the oxygen precovered surface three O(1s) emission lines are detected at lower temperature. They originate from adsorbed atomic oxygen, from OH groups, and from H2O molecules. For an oxygen coverage of half a monolayer, XPS indicates that during H2O decomposition the preadsorbed oxygen does not directly participate in the OH formation. After water adsorption on the clean surface three O(1s) emission lines are found, which are due to H2O molecules, “disturbed” H2O molecules, and OH groups. The XPS results are directly correlated with information about the adsorbates obtained by UPS. Coverages are determined from the XPS spectra for the detected species.  相似文献   

3.
《Surface science》1987,182(3):499-520
Photoelectron spectroscopy (UPS), thermal desorption spectroscopy (TDS), isotope exchange experiments, work function change (δφ) and LEED were used to study the adsorption and dissociation behavior of H2O on a clean and oxygen precovered stepped Ni(s)[12(111) × (111)] surface. On the clean Ni(111) terraces fractional monolayers of H2O are adsorbed weakly in a single adsorption state with a desorption peak temperature of 180 K, just above that of the ice multilayer desorption peak (Tm = 155 K). In the angular resolved UPS spectra three H2O induced emission maxima at 6.2, 8.5 and 12.3 eV below EF were found for θ ≈ 0.5. Angular and polarization dependent UPS measurements show that the C2v symmetry of the H2O gas-phase molecule is not conserved for H2O(ad) on Ni(s)(111). Although the Δφ suggest a bonding of H2O to Ni via the negative end of the H2O dipole, the O atom, no hints for a preferred orientation of the H2O molecular axes were found in the UPS, neither for the existence of water dimers nor for a long range ordered H2O bilayer. These results give evidence that the molecular H2O axis is more or less inclined with respect to the surface normal with an azimuthally random distribution. H2O adsorption at step sites of the Ni(s)(111) surface leads in TDS to a desorption maximum at Tm = 225 K; the binding energy of H2O to Ni is enhanced by about 30% compared to H2O adsorbed on the terraces. Oxygen precoverage causes a significant increase of the H2O desorption energy from the Ni(111) terraces by about 50%, suggesting a strong interaction between H2O and O(ad). Work function measurements for H2O+O demonstrate an increase of the effective H2O dipole moment which suggests a reorientation of the H2O dipole in the presence of O(ad), from inclined to a more perpendicular position. Although TDS and Δφ suggest a significant lateral interaction between H2O+O(ad), no changes in the molecular binding energies in UPS and no “isotope exchange” between 18O(ad) and H216O(ad) could be observed. Also, dissociation of H2O could neither be detected on the oxygen precovered Ni(s)(111) nor on the clean terraces.  相似文献   

4.
A. Spitzer  H. Lüth 《Surface science》1982,120(2):376-388
The water adsorption on clean and oxygen precovered Cu(110) surfaces is studied by means of UPS, LEED, work function measurements and ELS. At 90 K on the clean surface molecular water adsorption is indicated by UPS. The H2O molecules are bonded at the oxygen end and the H-O-H angle is increased as compared with the free molecule. In the temperature range between 90 and 300 K distorted H2O molecules and adsorbed hydroxyl species (OH) are detected, which are desorbed at room temperature. On an oxygen covered surface hydroxyl groups are formed by dissociation of adsorbed water molecules at a lower temperature than on the clean surface. Multilayers of condensed water are found below 140 K in both cases.  相似文献   

5.
The adsorption of H2O on Al(111) has been studied by ESDIAD (electron stimulated desorption ion angular distributions), LEED (low energy electron diffraction), AES (Auger electron spectroscopy) and thermal desorption in the temperature range 80–700 K. At 80 K, H2O is adsorbed predominantly in molecular form, and the ESDIAD patterns indicate that bonding occurs through the O atom, with the molecular axis tilted away from the surface normal. Some of the H2O adsorbed at 80 K on clean Al(111) can be desorbed in molecular form, but a considerable fraction dissociates upon heating into OHads and hydrogen, which leaves the surface as H2. Following adsorption of H2O onto oxygen-precovered Al(111), additional OHads is formed upon heating (perhaps via a hydrogen abstraction reaction), and H2 desorbs at temperatures considerably higher than that seen for H2O on clean Al(111). The general behavior of H2O adsorption on clean and oxygen-precovered Al(111) (θO ? monolayer) is rather similar at low temperature, but much higher reactivity for dissociative adsorption of H2O to form OH adsis noted on the oxygen-dosed surface around room temperature.  相似文献   

6.
《Surface science》1986,177(1):191-206
The adsorption and dissociation of H2O on Rh(111) and Rh foil surfaces have been studied in UHV using Auger electron, electron energy loss (in the electronic range) and thermal desorption spectroscopy. H2O adsorbs weakly on clean Rh samples at 110 K. The adsorption is accompanied by the appearance of a broad loss feature at 14–14.5 eV. At higher exposures new losses appeared at 8.6 and 10.5 eV. The desorption of H2O took place in two stages, with Tp = 183 K (β, chemisorption) and 158 K (α, multilayer formation). There was no indication of dissociation of H2O on a clean Rh(111) surface. Similar results were obtained for a clean Rh foil. However, when small amounts of boron segregated on the surface of Rh, they exerted a dramatic influence on the adsorptive properties of this surface and caused the dissociation of H2O. This was exhibited by the formation of H2, by the buildup of surface oxygen, by the appearance of an intense new loss at 9.4 eV, identified as B-O surface species, and by the development of “boron-oxide”-like Auger fine structure.  相似文献   

7.
The adsorption of H2O on clean and K-covered Pt(111) was investigated by utilizing Auger, X-ray and ultra-violet photoemission spectroscopies. The adsorption on Pt(111) at 100–150 K was purely molecular (ice formation) in agreement with previous work. No dissociation of this adsorbed H2O was noted on heating to higher temperatures. On the other hand, adsorption of H2O on Pt(111) + K leads to dissociation and to the formation of OH species which were characterized by a work function increase, an O 1s binding energy of 530.9 eV and UPS peaks at 4.7 and 8.7 eV below the Fermi level. The amount of OH formed was proportional to the K coverage for θK > 0.06 whereas no OH could be detected for θ? 0.06. Dissociation of H2O occurred already at T = 100 K, with a sequential appearance of O 1s peaks at 531 and 533 eV representing OH and adsorbed H2O, respectively. At room temperature and above only the OH species was observed. Annealing of the surface covered with coadsorbed K/OH indicated the high stability of this OH species which could be detected spectroscopically up to 570 K. The adsorption energy of H2O coadsorbed with K and OH on Pt(111) is increased relative to that of H2O on Pt. The work function due to this adsorbed H2O increases whereas it decreases for H2O on Pt(111). The energy shifts of valence and O1s core levels of H2O on Pt + K as deduced from a comparison of gas phase and adsorbate spectra are 2.8–4.2 eV compared to ≈ 1.3–2.3 eV for H2O on Pt (111). This increased relaxation energy shift suggests a charge transfer screening process for H2O on Pt + K possibly involving the unoccupied 4a1 orbital of H2O. The occurrence of this mode of screening would be consistent with the higher adsorption energy of H2O on Pt + K and with its high propensity to dissociate into OH and H.  相似文献   

8.
Photoelectron spectroscopic studies of the oxidation of Ni(111), Ni(100) and Ni(110) surfaces show that the oxidation process proceeds at 295 and 485 K in two distinct steps: a fast dissociative chemisorption of oxygen followed by oxide nucleation and lateral oxide growth to a limiting coverage of 3 NiO layers. The oxygen concentration in the 295 K saturated oxygen layer on Ni(111) was confirmed by 16O(d,p) 17O nuclear microanalysis. At 295 and 485 K the oxide growth rates are in the order Ni(110) > Ni(111) > Ni(100). At 77 K the oxygen uptake proceeds at the same rate on all three surfaces and shows a continually decreasing sticking coefficient to saturation at ~2.1 layers (based upon NiO). An O 1sb.e. = 529.7 eV is associated with NiO, and O ls b.e.'s of ~531.5 and 531.3 eV can be associated, respectively, with defect oxide (Ni2O3) or (in the presence of H2O) with an NiO(H) species. The binding energies (Ni 2p, O 1s) of this NiO(H) species are similar to those for Ni(OH)2. Defect oxides are produced by oxidation at 485 K, or by oxidation of damaged films (e.g. from Ar+ sputtering) and evaporated films. Wet oxidation (or exposure to air) of clean nickel surfaces and oxides, and exposure of thick oxide to hydrogen at high temperature results in an O 1s b.e. ~531.3 eV species. Nuclear microanalysis 2H(3He,p) 4He indicates the presence of protonated species in the latter samples. Oxidation at 77 K yields O 1s b.e.'s of 529.7 and ~531 eV; the nature of the high b.e. species is not known. Both clean and oxidised nickel surfaces show a low reactivity towards H2O; clean nickel surfaces are ~103 times less reactive to H2O than to oxygen.  相似文献   

9.
A study of the adsorption/desorption behavior of CO, H2O, CO2 and H2 on Ni(110)(4 × 5)-C and Ni(110)-graphite was made in order to assess the importance of desorption as a rate-limiting step for the decomposition of formic acid and to identify available reaction channels for the decomposition. The carbide surface adsorbed CO and H2O in amounts comparable to the clean surface, whereas this surface, unlike clean Ni(110), did not appreciably adsorb H2. The binding energy of CO on the carbide was coverage sensitive, decreasing from 21 to 12 kcalmol as the CO coverage approached 1.1 × 1015 molecules cm?2 at 200K. The initial sticking probability and maximum coverage of CO on the carbide surface were close to that observed for clean Ni(110). The amount of H2, CO, CO2 and H2O adsorbed on the graphitized surface was insignificant relative to the clean surface. The kinetics of adsorption/desorption of the states observed are discussed.  相似文献   

10.
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   

11.
H2S, H2 and S adsorbed on Ru(110) have been studied by angle-integrated ultraviolet photoemission (UPS) as part of a study of the effect of adsorbed sulfur, a common catalytic poison, on this Ru surface. For low exposures of H2S at 80 K, the work function rises to a value 0.16 eV above that of clean Ru(110) while the associated UPS spectra (hν = 21.2 eV) exhibit features similar to those of H(ads) and S(ads) and different from those of molecular H2S. We conclude that H2S dissociates completely at low coverages on Ru(110) at 80 K. At intermediate exposures the work function drops and the UPS spectra show new features which are attributed to the presence of an adsorbed SH species. This appears to be the first direct observation of this surface complex. At higher exposures the work function saturates at a value 0.36 eV below the clean value; the UPS spectra change markedly and indicate the adsorption of molecular H2S. Heating adsorbed H2S leaves a stable layer of S(ads) on Ru(110). The surface with adsorbed sulfur strongly modifies the adsorption at 80 K of a number of molecules relative to the clean Ru(110) surface.  相似文献   

12.
Electron energy loss spectra on a (110) nickel surface exhibit characteristic changes upon adsorption of H2, CO and O2. The clean surface shows only the surface and bulk plasmon losses at 8 eV and 18 eV respectively. Adsorption of CO produces two new loss peaks at 13.5 eV and 5.5 eV. Loss peaks due to hydrogen adsorption at 15 eV and 7.5 eV show a strong correlation with the well known adsorption characteristics of this system. The oxygen induced losses are different for chemisorbed O on Ni and NiO. In any case the chemisorption-induced losses are well established for primary energies below 120eV. In the loss spectra with higher excitation energies only a drastic decrease of the surface plasmon loss peak-height is visible. If the new losses can be attributed to one-electron excitations from molecular orbital levels due to the chemisorption bond, with assumptions of the final state of the excited electron a determination of the postition of these levels can be made. In case of CO and H2 reasonable results are evaluated.  相似文献   

13.
Elastic and direct-inelastic scattering as well as dissociative adsorption and associative desorption of H2 and D2 on Ni(110) and Ni(111) surfaces were studied by molecular beam techniques. Inelastic scattering at the molecular potential is dominated by phonon interactions. With Ni(110), dissociative adsorption occurs with nearly unity sticking probability s0, irrespective of surface temperature Ts and mean kinetic energy normal to the surface 〈 E 〉. The desorbing molecules exhibit a cos θe angular distribution indicating full thermal accommodation of their translation energy. With Ni(111), on the other hand, s0 is only about 0.05 if both the gas and the surface are at room temperature. s0 is again independent of Ts, but increases continuously with 〈 E⊥ 〉 up to a value of ~0.4 forE⊥ 〉 = 0.12 eV. The cos5θe angular distribution of desorbing molecules indicates that in this case they carry off excess translational energy. The results are qualitatively rationalized in terms of a two-dimensional potential diagram with an activation barrier in the entrance channel. While the height of this barrier seems to be negligible for Ni(110), it is about 0.1 eV for Ni(111) and can be overcome through high enough translational energy by direct collision. The results show no evidence for intermediate trapping in a molecular “precursor” state on the clean surfaces, but this effect may play a role at finite coverages.  相似文献   

14.
The temperature-programmed reaction (TPR) method, high-resolution electron energy loss spectroscopy (HREELS), and molecular beam method were used to elucidate the role surface reconstruction, subsurface oxygen (Osubs), and COads concentration play in the low-temperature oxidation of CO on the Pt(100), Pt(410), Pd(111), and Pd(110) surfaces. The possibility of the formation of so-called hot oxygen atoms, which arise at the surface at the instant of dissociation of O2ads molecules and can react with COads at low temperatures (~150 K) to form CO2, was examined. It was revealed that, when present in high concentration, COads initiates the phase transition of the Pt(100)-(hex) reconstructed surface into the (1 × 1) non-reconstructed one and blocks fourfold hollow sites of oxygen adsorption (Pt4-Oads), thereby initiating the formation of weakly bound oxygen (Pt2-Oads), active in CO oxidation. For the Pt(410), Pd(111), and Pd(110) surfaces, the reactivity of Oads with respect to CO was demonstrated to be dependent on the surface coverage of COads. The 18Oads isotope label was used to determine the nature of active oxygen reacting with CO at ~150–200 K. It was examined why a COads layer produces a strong effect on the reactivity of atomic oxygen. The experimental results were confirmed by theoretical calculations based on the minimization of the Gibbs energy of the adsorption layer. According to these calculations, the COads layer causes a decrease in the apparent activation energy E act of the reaction due to changes in the type of coordination and in the energy of binding of Oads atoms to the surface.  相似文献   

15.
The chemisorption of both CO and O2 on a clean tungsten ribbon has been studied using an ultrahigh vacuum X-ray photoelectron spectrometer. For CO, the energy and intensity of photoemission from O(1s) and C(1s) core levels have been studied for various adsorption temperatures.At adsorption temperatures of ~100 K., the “virgin”-CO state was the dominant adsorbed species. Conversion of this state to more strongly-bound β-CO is observed upon heating the adsorbed layer to ~320K. Thermal desorption of CO at 300?T?640 K causes sequential loss of α1-CO and α2-CO as judged by the disappearance of O(1s) and C(1s) photoelectron peaks characteristic of these states.Oxygen adsorption at 300K gives a single main O(ls) peak at all coverages, although at high oxygen coverages there exist small auxiliary peaks at ~2eV lower kinetic energy. The photoelectron C(1s) and O(1s) binding energies observed for these adsorbed species are all lower than for gaseous molecules containing C and O atoms. For CO adsorption states there is a systematic decrease in photoelectron binding energy as the strength of adsorption increases. These observations are in general accord with expectations based on electronic relaxation effects in condensed materials.  相似文献   

16.
The interaction of Cs and O2 on MoS2(0001) has been studied both in the alternate adsorption and the codeposition mode by LEED, AES, TDS and WF measurements at 170 and 300 K. Oxygen does not interact with Cs when θCs?0.04 at 300 K or θCs?0.08 at 170 K, where Cs is known to adsorb as strongly ionized, individual adatoms. The interaction at higher θCs, where Cs is known to form clusters on MoS2(0001), leads to clusters of a Cs/O complex characterized by a Cs(563 eV)/O(512 eV) Auger peak ratio of 1.1–1.3. The minimum WF is 2.1 eV at 300 and 170 K upon alternate adsorption, and 1.7 eV at both T upon codeposition. Upon heating, oxygen and Cs desorb independently, as no oxide desorption is observed. The Cs TDS spectrum is shifted to lower T in the presence of oxygen and a new desorption peak appears at ~ 880 K. The differences in the Cs/O interaction between MoS2(0001) and other semiconductors and metals are attributed to the Cs clustering and the inertness of MoS2(0001) to O2 adsorption.  相似文献   

17.
《Surface science》1986,177(3):515-525
The adsorption of bromine on the (110) surface of silver has been studied by ultraviolet ( = 21.2 and 40.8 eV) photoelectron spectroscopy in the temperature range of 100–300 K. Four different adsorption and reaction states could be detected. For fractional monolayer coverages Br2 adsorbs dissociatively on the Ag(110) surface. The chemisorption of bromide leads to new emission features at about 3 and 5.2 eV below EF, which are assigned as occupied antibonding structures (3 eV) and as bonding Br4px, y orbitals (5.2 eV). At 100 K, further bromide adsorption leads to the formation of an AgBr layer with molecular adsorbed bromine on top of this corrosion layer. The He I spectrum is dominated by structures at 3.5, 5.8 and 7.5 eV which are due to emission from the πg, πu and σg molecular orbitals of Br2. The buildup of the AgBr layer is clearly demonstrated by desorbing the molecular bromine at about 150 K. The resulting spectrum of the AgBr layer shows peaks at 2.5 and 3.4 eV with p- and mixed-in d-character and peaks at 4.1, 5.2 and 6.1 eV which are primarily d-like. Heating of the AgBr layer up to 300 K results in a transformation from a 2D layer into a 3D agglomeration of larger AgBr clusters on top of a Br/Ag(110) chemisorption layer.  相似文献   

18.
《Surface science》1987,179(1):59-89
Molecular adsorption is observed on a Ni(110) surface at 80 K. The relative binding energies of the valence ion states as determined by ARUPS are consistent with those in the gas as well as in the condensed phase, and indicate that the electronic structure of the adsorbed molecule is only slightly distorted upon adsorption at this temperature. The adsorbate spectra show E versus k dispersions indicating some long-range order in the adsorbate. The variations in relative ionisation probabilities of the ion states as a function of electron emission angle suggest that the molecular axis is oriented parallel to the surface within ± 20°. Upon heating the adsorbate to above 100 K. (i.e. 140 K) the spectrum changes. A new species causing an increase in work function by 1 eV can be identified. Comparison with calculations suggests that it is an anionic bent CO; molecule. Electron energy loss studies on this intermediate species support the proposed bent CO2 geometry and favour a coordination site with C2v symmetry. The bent CO2 moiety is stable up to 230 K. Further heating to room temperature leads to dissociation of the bent CO2 molecule into adsorbed CO and O. The CO molecule is oriented with its axis perpendicular to the surface. The bent CO2 species appears to be a precursor to dissociation. Results on CO2 adsorption on an oxygen precovered surface show that CO2 interacts with oxygen at 85 K. Upon heating the co-adsorbate to near room temperature a reaction product is formed the nature of which cannot yet be clearly identified.  相似文献   

19.
The interaction of O2, CO2, CO, C2H4 AND C2H4O with Ag(110) has been studied by low energy electron diffraction (LEED), temperature programmed desorption (TPD) and electron energy loss spectroscopy (EELS). For adsorbed oxygen the EELS and TPD signals are measured as a function of coverage (θ). Up to θ = 0.25 the EELS signal is proportional to coverage; above 0.25 evidence is found for dipole-dipole interaction as the EELS signal is no longer proportional to coverage. The TPD signal is not directly proportional to the oxygen coverage, which is explained by diffusion of part of the adsorbed oxygen into the bulk. Oxygen has been adsorbed both at pressures of less than 10-4 Pa in an ultrahigh vacuum chamber and at pressures up to 103 Pa in a preparation chamber. After desorption at 103 Pa a new type of weakly bound subsurface oxygen is identified, which can be transferred to the surface by heating the crystal to 470 K. CO2 is not adsorbed as such on clean silver at 300 K. However, it is adsorbed in the form of a carbonate ion if the surface is first exposed to oxygen. If the crystal is heated this complex decomposes into Oad and CO2 with an activation energy of 27 kcal/mol(1 kcal = 4.187 kJ). Up to an oxygen coverage of 0.25 one CO2 molecule is adsorbed per two oxygen atoms on the surface. At higher oxygen coverages the amount of CO2 adsorbed becomes smaller. CO readily reacts with Oad at room temperature to form CO2. This reaction has been used to measure the number of O atoms present on the surface at 300 K relative to the amount of CO2 that is adsorbed at 300 K by the formation of a carbonate ion. Weakly bound subsurface oxygen does not react with CO at 300 K. Adsorption of C2H4O at 110 K is promoted by the presence of atomic oxygen. The activation energy for desorption of C2H4O from clean silver is ~ 9 kcal/mol, whereas on the oxygen-precovered surface two states are found with activation energies of 8.5 and 12.5 kcal/mol. The results are discussed in terms of the mechanism of ethylene epoxidation over unpromoted and unmoderated silver.  相似文献   

20.
Adsorption and reactivity of carbon dioxide at the clean and oxygen precovered Ni(110) surface has been studied by means of EELS and LEED. On the clean surface two different types of CO2 molecules have been observed by EELS at 135 K, one being the undisturbed linear configuration. With increasing temperature the linear molecule changes into a different species which precedes dissociation at 220 K into CO and O. EELS and LEED data of the intermediate species support the assumption that it is a bent CO2 anion adsorbed in C2v symmetry with twofold oxygen coordination to the surface. Oxygen preadsorption stabilizes the linear CO2 molecule up to higher temperatures which does not convert into a bent species in this case. Instead, a reaction product of CO2 and O is found and interpreted as a carbonate species.  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