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 adsorption of hydrogen on Pt (100) was investigated by utilizing LEED, Auger electron spectroscopy and flash desorption mass spectrometry. No new LEED structures were found during the adsorption of hydrogen. One desorption peak was detected by flash desorption with a desorption maximum at 160 °C. Quantitative evaluation of the flash desorption spectra yields a saturation coverage of 4.6 × 1014 atoms/cm2 at room temperature with an initial sticking probability of 0.17. Second order desorption kinetics was observed and a desorption energy of 15–16 kcal/mole has been deduced. The shapes of the flash desorption spectra are discussed in terms of lateral interactions in the adsorbate and of the existence of two substates at the surface. The reaction between hydrogen and oxygen on Pt (100) has been investigated by monitoring the reaction product H2O in a mass spectrometer. The temperature dependence of the reaction proved to be complex and different reaction mechanisms might be dominant at different temperatures. Oxygen excess in the gas phase inhibits the reaction by blocking reactive surface sites. At least two adsorption states of H2O have to be considered on Pt (100). Desorption from the prevailing low energy state occurs below room temperature. Flash desorption spectra of strongly bound H2O coadsorbed with hydrogen and oxygen have been obtained with desorption maxima at 190 °C and 340 °C.  相似文献   

2.
The adsorption and reaction of water on clean and oxygen covered Ag(110) surfaces has been studied with high resolution electron energy loss (EELS), temperature programmed desorption (TPD), and X-ray photoelectron (XPS) spectroscopy. Non-dissociative adsorption of water was observed on both surfaces at 100 K. The vibrational spectra of these adsorbates at 100 K compared favorably to infrared absorption spectra of ice Ih. Both surfaces exhibited a desorption state at 170 K representative of multilayer H2O desorption. Desorption states due to hydrogen-bonded and non-hydrogen-bonded water molecules at 200 and 240 K, respectively, were observed from the surface predosed with oxygen. EEL spectra of the 240 K state showed features at 550 and 840 cm?1 which were assigned to restricted rotations of the adsorbed molecule. The reaction of adsorbed H2O with pre-adsorbed oxygen to produce adsorbed hydroxyl groups was observed by EELS in the temperature range 205 to 255 K. The adsorbed hydroxyl groups recombined at 320 K to yield both a TPD water peak at 320 K and adsorbed atomic oxygen. XPS results indicated that water reacted completely with adsorbed oxygen to form OH with no residual atomic oxygen. Solvation between hydrogen-bonded H2O molecules and hydroxyl groups is proposed to account for the results of this work and earlier work showing complete isotopic exchange between H216O(a) and 18O(a).  相似文献   

3.
Using molecular-beam relaxation techniques and isotopic exchange experiments, the water-formation reaction on Pd(111) has been shown to proceed via a Langmuir-Hinshelwood mechanism. The reaction product H2O is emitted from the surface with a cosine distribution. The rate-determining step is the formation of OHad in the reaction Oad + Had → OHad. The activation energy for this step is 7 kcal/mole with a pre-exponential factor, v, of 4 × 10?8 cm2 atom?1 sec?1. This value for v lies well below that observed for simple second-order desorption of dissociatively adsorbed diatomic gases, but is roughly of the order of that obtained for the oxidation of CO on Pd(111). The formation of H2O proceeds differently under conditions of excess O2 or H2. In an excess of H2, the kinetics is dominated by the transport of atomic hydrogen between the bulk and the surface as was found for the H?D exchange reaction on Pd(111). In an excess of O2, diffusion of hydrogen into the bulk is blocked by adsorbed oxygen and the hydrogen reservoir available for reaction at the surface is decreased by several orders of magnitude. This results in a drastic reduction of the reaction rate which can be reversed by increasing the partial pressure of H2.  相似文献   

4.
On metals such as Zr, during hydrogen exposure, dissolution competes with desorption; this competition can be probed by thermal desorption at different heating rates. In the case of desorption from preadsorbed hydrogen, only ∼1% of the hydrogen can be desorbed even at heating rates of >1010 K s−1. Recent measurements of the dynamics of hydrogen released by water dissociation on Zr(0 0 0 1) [G. Bussière, M. Musa, P.R. Norton, K. Griffiths, A.G. Brolo, J.W. Hepburn, J. Chem. Phys. 124 (2006) 124704] have shown that the desorbing hydrogen originates from the recombinative desorption of adsorbed H-atoms and that over 25% of the water collisions lead to hydrogen desorption. To gain further insight into the desorption and dissolution of hydrogen and in an attempt to resolve the paradox of the different desorption yields from H2 vs. H2O exposures, we report new measurements of the laser induced thermal desorption (LITD) of hydrogen from Zr(0 0 0 1) at initial temperatures down to 90 K. The low temperature was chosen because work function measurements suggested that hydrogen adsorbed into only the outermost (surface site) of the two available adsorption sites (surface and subsurface), from which we postulated much more efficient desorption at high heating rates compared to desorption from the sub-surface sites. However, hydrogen desorption by LITD from Zr(0 0 0 1) at 90 K still only accounts for 1% of the adsorbed species, the remainder dissolving into the bulk at LITD heating rates. The different yields alluded to above remain unexplained (Bussière, 2006).  相似文献   

5.
Optical second-harmonic generation (SHG) from silicon surfaces may be resonantly enhanced by dangling-bond-derived surface states. The resulting high sensitivity to hydrogen adsorption combined with unique features of SHG as an optical probe has been exploited to study various kinetical and dynamical aspects of the adsorption system H2/Si. Studies of surface diffusion of H/Si(111)7×7 and recombinative desorption of hydrogen from Si(111)7 × 7 and Si(100)2 × 1 revealed that the covalent nature of hydrogen bonding on silicon surfaces leads to high diffusion barriers and to desorption kinetics that strongly depend on the surface structure. Recently, dissociative adsorption of molecular hydrogen on Si(100)2×1 and Si(111)7×7 could be observed for the first time by heating the surfaces to temperatures between 550 K and 1050 K and monitoring the SH response during exposure to a high flux of H2 or D2. The measured initial sticking coefficients for a gas temperature of 300K range from 10–9 to 10–5 and strongly increase as a function of surface temperature. These results demonstrate that the lattice degrees of freedom may play a decisive role in the reaction dynamics on semiconductor surfaces.  相似文献   

6.
We have studied desorption of 13CO and H2O and desorption and reaction of coadsorbed, 13CO and H2O on Au(310). From the clean surface, CO desorbs mainly in, two peaks centered near 140 and 200 K. A complete analysis of desorption spectra, yields average binding energies of 21 ± 2 and 37 ± 4 kJ/mol, respectively. Additional desorption states are observed near 95 K and 110 K. Post-adsorption of H2O displaces part of CO pre-adsorbed at step sites, but does not lead to CO oxidation or significant shifts in binding energies. However, in combination with electron irradiation, 13CO2 is formed during H2O desorption. Results suggest that electron-induced decomposition products of H2O are sheltered by hydration from direct reaction with CO.  相似文献   

7.
The interaction of H2O with Zircaloy-4 (Zry-4) is investigated using Auger electron spectroscopy (AES) and temperature programmed desorption (TPD) methods. Following adsorption of H2O at 150 K the Zr(MNV) and Zr(MNN) Auger features shift by ∼6.5 and 4.5 eV, respectively, indicating surface oxidation. Heating H2O/Zry-4 results in molecular desorption of water at both low and high temperatures. The low-temperature desorption is attributed to ice multilayers, whereas, three overlapping high-temperature features are presumably due to recombinative desorption. This high-temperature desorption begins before the surface oxide is dissolved, continues upon its removal, and is atypical for water/metal systems. Unexpectedly, no significant desorption of hydrogen is observed near 400 K, as is typically observed following O2 adsorption on Zr-based materials. However, we do observe that H2O adsorption on Zry-4 surfaces roughened by argon ion sputtering results in H2 desorption.  相似文献   

8.
The kinetics of simultaneous hydrogen and deuterium thermal desorption from PdHxDy has been investigated. A novel experimental approach for the study of the transition state (TS) characteristics of the surface recombination reaction is proposed based on the analysis of the H and D partitioning into H2, HD and D2 molecules. It has been found that the hydrogen molecular isotopes distribution is determined by the energy differences of the corresponding TS of the atom-atom recombination reactions. On the other hand, the mechanisms and activation energies of the desorption process have been obtained. At 420 K, the desorption reaction changes from a surface recombination limiting mechanism during desorption from β-PdHxDy to a reaction limited by the rate of β to α phase transformation during the two phase coexistence. Surface recombination reaction becomes again rate limiting above 480 K, due to a change in the catalytic properties of the Pd surface. TS energies obtained from the kinetic analysis of the thermal desorption spectra are in good accordance with those obtained from the analysis of the H2, HD and D2 distributions. Anomalous TS energies have been observed for the H-D recombination reaction, which may be related to the heteronuclear character of this molecule.  相似文献   

9.
《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   

10.
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   

11.
The oxidation of methanol was studied on a Ag(110) single-crystal by temperature programmed reaction spectroscopy. The Ag(110) surface was preoxidized with oxygen-18, and deuterated methanol, CH3OD, was used to distinguish the hydroxyl hydrogen from the methyl hydrogens. Very little methanol chemisorbed on the oxygen-free Ag(110) surface, and the ability of the silver surface to dissociatively chemisorb methanol was greatly enhanced by surface oxygen. CH3OD was selectively oxidized upon adsorption at 180 K to adsorbed CH3O and D218O, and at high coverages the D218O was displaced from the Ag(110) surface. The methoxide species was the most abundant surface intermediate and decomposed via reaction channels at 250, 300 and 340 K to H2CO and hydrogen. Adsorbed H2CO also reacted with adsorbed CH3O to form H2COOCH3which subsequently yielded HCOOCH3 and hydrogen. The first-order rate constant for the dehydrogenation of D2COOCH3 to DCOOCH3 and deuterium was found to be (2.4 ± 2.0) × 1011 exp(?14.0 ± 0.5 kcalmole · RT)sec?1. This reaction is analogous to alkoxide transfer from metal alkoxides to aldehydes in the liquid phase. Excess surface oxygen atoms on the silver substrate resulted in the further oxidation of adsorbed H2CO to carbon dioxide and water. The oxidation of methanol on Ag(110) is compared to the previous study on Cu(110).  相似文献   

12.
《Surface science》1994,303(3):L385-L391
The oxygen-exchange reaction between N16O and 18O2 coadsorbed on Pt(111) has been studied by temperature-programmed desorption (TPD). Reaction products of N18O and 18O16O are desorbed from Pt(111) initially saturated with 18O2 at 94 K followed by exposure of N16O. Three distinct desorption peaks are observed in N18O TPD spectra at 145, 310, and 340 K, and two peaks in 18O16O at 155 K and between 600 and 1000 K. In contrast, the exchange reaction is greatly suppressed when oxygen molecules are replaced with oxygen adatoms at three-fold hollow sites of Pt(111). These results strongly suggest that adsorbed oxygen molecules are responsible for the exchange reaction. NO2 or NO3 is postulated as a reaction intermediate. However, since desorption signals corresponding to these species are not detected, the oxygen-exchanged products are not due to the cracking processes of the higher order nitrogen oxides in the mass spectrometer. Thus, the reaction proceeds via the intermediate that is dissociated during the elevation of surface temperature.  相似文献   

13.
The oxidation of H2C16O by adsorbed 18O was studied on an Cu(110) sample by temperature programmed reaction spectroscopy. Formaldehyde exchanged its oxygen with surface 18O upon adsorption to yield H2C18O(a) and 16O(a). Formaldehyde was also oxidized by surface 16O and 18O atoms to H2COO which subsequently released one of the hydrogen atoms to form HCOO. The evolution of H2 from the Cu(110) surface was desorption limited, and the low pre-exponential factor for the recombination of the surface hydrogen atoms suggested stringent requirement on the trajectories of the colliding partners. The formate was very stable and dissociated at elevated temperatures to simultaneously yield H2 and CO2. The surface concentration of 18O exerted a pronounced affect on the activity of the oxidation of formaldehyde on Cu(110).  相似文献   

14.
《Surface science》1986,172(3):544-556
Thermal programmed desorption, ultraviolet photoelectron spectroscopy, and X-ray photoelectron spectroscopy have been used to study the reaction of H2O with stoichiometric and partially reduced single crystals of α-Fe2O3. On the stoichiometric surface only ice condensation below 200 K was observed. Oxygen deficient surfaces were prepared by Ar bombardment giving rise to a decrease in the work function of the crystal of up to 1 eV. On these surfaces OH species were formed as detected by UPS that were stable up to 320 K. Annealing the defective surfaces between 475 and 700 K increased the work function by values between 0.5 and 0.7 eV respectively. These surfaces contained reduced Fe2+ species in subsurface layers as shown by UPS and XPS, but were inactive towards H2O chemisorption. The Fe2+ species were stable for long periods of time at temperatures of up to 775 K. Potassium deposited on the surface forms a strongly bound monolayer compound. With H2O it produced a complex that resulted in H2 evolution upon annealing.  相似文献   

15.
Surface ions generated by electron stimulated desorption from mass spectrometer ion source grids are frequently observed, but often misidentified. For example, in the case of mass 19, the source is often assumed to be surface fluorine, but since the metal oxide on grid surfaces has been shown to form water and hydroxides, a more compelling case can be made for the formation of hydronium. Further, fluorine is strongly electronegative, so it is rarely generated as a positive ion. A commonly used metal for ion source grids is 316L stainless steel. Thermal vacuum processing by bakeout or radiation heating from the filament typically alters the surface composition to predominantly Cr2O3. X-ray photoelectron spectral shoulders on the O 1s and Cr 2p3/2 peaks can be attributed to adsorbed water and hydroxides, the intensity of which can be substantially increased by hydrogen dosing. On the other hand, the sub-peak intensities are substantially reduced by heating and/or by electron bombardment. Electron bombardment diode measurements show an initial work function increase corresponding to predominant hydrogen desorption (H2) and a subsequent work function decrease corresponding to predominant oxygen desorption (CO). The fraction of hydroxide concentration on the surface was determined from X-ray photoelectron spectroscopy and from the deconvolution of temperature desorption spectra. Electron stimulated desorption yields from the surface show unambiguous H3O+ peaks that can be significantly increased by hydrogen dosing. Time of flight secondary ion mass spectrometry sputter yields show small signals of H3O+, as well as its constituents (H+, O+ and OH+) and a small amount of fluorine as F, but no F+ or F+ complexes (HF+, etc.). An electron stimulated desorption cross-section of σ+ ∼ 1.4 × 10−20 cm2 was determined for H3O+ from 316L stainless steel for hydrogen residing in surface chromium hydroxide.  相似文献   

16.
《Surface science》1996,364(2):L580-L586
The adsorption and decomposition of formic acid on NiO(111)-p(2 × 2) films grown on Ni(111) single crystal surface were studied by temperature-programmed desorption (TPD) spectroscopy. Exposure of formic acid at 163 K resulted in both molecular adsorption and dissociation to formate. The adsorbed formate underwent further dissociation to H2, CO2 and CO. H2 and CO2 desorbed at the same temperatures of 340, 390 and 520 K, while CO desorbed at 415 and 520 K. The desorption features varied with the formic acid exposure. Two reaction channels were identified for the decomposition of formate under equilibrium with gas-phase formic acid with a pressure of 2.5 × 10−4Pa, one preferentially producing H2 and CO2 with an activation energy of 22 ± 2 kJ mol−1 and the other preferentially producing CO and H2O with an activation energy of 16 ± 2 kJ mol−1. The order of both reaction paths was 0.5 with respect to the pressure of formic acid.  相似文献   

17.
The adsorption and reaction of hydrogen sulfide, H2S, have been studied on cerium oxide thin films that were vapor deposited on Ru(0 0 0 1). The behavior of the H2S was examined as a function of Ce oxidation state. H2S weakly chemisorbs on fully oxidized CeO2 desorbing near 155 K. Hydrogen from the H2S reacts with the surface O to desorb as water between 200 K and 450 K. When ca. 20% of the Ce4+ is reduced to Ce3+ more H2S dissociates to -OH and -SH and water is produced near 580 K. When the ceria is ca. 70% reduced, water formation is suppressed and H2 desorbs near 580 K. S 2p photoelectron spectroscopy indicates the decomposition of H2S into -SH and then -S as the sample is annealed from 100 K to 600 K. O 1s photoemission indicated the presence of H2O and -OH.  相似文献   

18.
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   

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.
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   

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