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 31 毫秒
1.
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   

2.
The polar Zn-ZnO(0001) surface is involved in the catalysis of methanol synthesis and the water–gas-shift reaction. We use density functional theory calculations to explore the favorable binding geometries and energies of adsorption of several molecular species relevant to these reactions, namely carbon monoxide (CO), carbon dioxide (CO2), water (H2O) and methanol (CH3OH). We also consider several proposed reaction intermediates, including hydroxymethyl (CH2OH), methoxyl (CH3), formaldehyde (CH2O), methyl (CH3), methylene (CH2), formic acid (HCOOH), formate (HCOO), formyl (HCO), hydroxyl (OH), oxygen (O) and hydrogen (H). For each, we identify the preferred binding geometry at a coverage of 1/4 monolayers (ML), and report calculated vibrational frequencies that could aid in the identification of these species in experiment. We further explore the effects on the binding energy when the adsorbate coverage is lowered to 1/9 and 1/16 ML.  相似文献   

3.
By means of temperature-programmed desorption (TPD) and X-ray photoemission spectroscopy (XPS) with synchrotron radiation, we investigated the adsorption and thermal decomposition of alkanethiols (RSH, R = CH3, C2H5, and C4H9) on a GaAs(1 0 0) surface. All chemisorbed alkanethiols can deprotonate to form thiolates below 300 K via dissociation of the sulfhydryl hydrogen (-SH). Two types of thiolates species are observed on GaAs(1 0 0), according to adsorption on surface Ga and As sites. The thiolates adsorbed on a Ga site preferentially recombine with surface hydrogen to desorb as a molecular thiol at 350-385 K. The thiolate on the As site exhibits greater thermal stability and undergoes mainly dissociation of the C-S bond at ∼520 K, independent of the alkyl chain length. The decomposition of CH3S either directly desorbs CH3 or transfers the CH3 moiety onto the surface. The surface CH3 further evolves directly from the surface at 665 K. The dissociations of C2H5S and C4H9S yield surface C2H5 and C4H9, which further decompose to desorb C2H4 and C4H8, respectively, via β-hydride elimination. The complete decomposition of alkanethiol leads to the formation of surface S without deposition of carbon. Adsorption of CH3SSCH3 results in the formation of surface CH3S at initial exposures via scission of the S−S bond. Compared with the adsorption of CH3SH, the CH3S on the Ga site exhibits greater thermal stability because surface hydrogen is absent. At a high exposure, CH3SSCH3 can absorb molecularly on the surface and decompose to desorb CH3SCH3 via formation of a CH3SS intermediate.  相似文献   

4.
We have used uv photoeinission (primarily at a photon energy hv = 40.8 eV) to study chemisorption and decomposition reactions of small oxygen-containing organic molecules on clean polycrystalline Pd surfaces at 120 and 300 K. These molecules include methanol (CH3OH), dimethyl ether (CH3OCH3) formaldehyde (H2CO), acetaldehyde [H(CH3)CO], and acetone [(CH3)2CO]. Chemisorption bonding of these molecules to the Pd surface occurs primarily through the lone-pair orbitais associated with the oxygen atoms, as evidenced by chemical bonding shifts of these orbitais toward larger electron binding energy relative to the other adsorbate valence orbitals. At 300 K all the molecules studied decompose on the surface, resulting in chemisorbed CO. Since chemisorbed (as well as condensed) phases of some of these molecules (CH3OH and H(CH3)CO) are observed at low temperature, the decomposition to CO is a thermally-activated reaction. The observed orbital shifts associated with chemisorption bonding are used to make rough estimates of interaction strengths and chemisorption bond energies (within the framework of Mulliken's theory of electron donor-acceptor complexes as applied to chemisorption by Grimley). The resulting heats of chemisorption are consistent with the observed surface reactions.  相似文献   

5.
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   

6.
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   

7.
The thermal chemistry of allyl alcohol (CH2CHCH2OH) on a Ni(100) single-crystal surface was studied by the temperature programmed desorption (TPD) and the X-ray photoelectron spectroscopy (XPS). The allyl alcohol adsorbs molecularly on the metal surface at 100 K. Intact molecular desorption from the surface occurs at temperatures around 180 K, but some molecules exhibit chemical reactivity on the surface: activation of the OH, CC, and CO bonds produces η1(O)-allyloxy CH2CHCH2O(a), η2(C, C) allyl alcohol (C(a)H2C(a)HCH2OH), and η3(C, C, O)-alkoxide (C(a)H2C(a)CH2 O(a)) intermediates. Further thermal activation of allyl alcohol on the surface yields propylene (CH2CHCH3), 1-propanol (CH3CH2CH2OH), propanal (CH3CH2CHO), and combustion and dehydrogenation products (H2O, H2, and CO). Propylene desorbs from the surface at temperatures of around 270 K. Hydrogenation to the η3(C, C, O)-alkoxide intermediate leads to the production of propanal which desorbs from the surface around 320 K, while hydrogenation of the η2(C, C) allyl alcohol intermediate produces 1-propanol, which desorbs at around 310 K. The co-adsorption of hydrogen atoms on the surface enhances the formation of the saturated alcohol, while co-adsorption of oxygen enhances the formation of both the saturated alcohol and the saturated aldehydes.  相似文献   

8.
The chemisorption, condensation, desorption, and decomposition of methanol, both CH3OH and CH3OD, on a clean Ni(110) surface have been characterized using high resolution electron energy loss spectroscopy, temperature programmed reaction spectroscopy, and low energy electron diffraction. The vibrational spectrum of the saturated chemisorbed layer, 7.4 × 1014 molecules cm?2, is almost identical to the infrared spectrum of liquid or solid methanol. Condensation of multilayers of methanol is facile at 80 K. The only quasi-stable intermediate isolated during the decomposition is a methoxy species, CH3O, which decomposes thermally to CO and H. The evolution of both CO and H2 occurs in desorption limited processes.  相似文献   

9.
The adsorption and reaction of H2O with adsorbed oxygen atoms on Ag(110) was examined by UPS. In agreement with previous EELS results, H2O formed multilayers of ice upon adsorption at 140 K. The ice layers could be easily distinguished from monolayer coverages of chemisorbed H2O (present above 160 K) by UPS. The ice layers produced (1) strong attenuation of the emission from the Ag d-bands, (2) a nearly 2 eV shift of H2O valence levels to higher binding energy and (3) strong attenuation of emission from the H2O 3a1 orbital. H2O was observed to react stoichiometrically with O(a) above 250 K to produce a pure layer of adsorbed hydroxyl species. The UPS spectra for these species exhibited features at ?5.8 and ?8.7 eV, as well as strong features above the d-bands. These spectra were compared with those for OH(a) on other surfaces, and the difficulties of identifying OH by UPS due to contamination by excess H2O are discussed.  相似文献   

10.
The adsorption and reaction of H2O on clean and oxygen precovered Ni(110) surfaces was studied by XPS from 100 to 520 K. At low temperature (T<150 K), a multilayer adsorption of H2O on the clean surface with nearly constant sticking coefficient was observed. The O 1s binding energy shifted with coverage from 533.5 to 534.4 eV. H2O adsorption on an oxygen precovered Ni(110) surface in the temperature range from 150 to 300 K leads to an O 1s double peak with maxima at 531.0 and 532.6 eV for T=150 K (530.8 and 532.8 eV at 300 K), proposed to be due to hydrogen bonded Oads… HOH species on the surface. For T>350 K, only one sharp peak at 530.0 eV binding energy was detected, due to a dissociation of H2O into Oads and H2. The s-shaped O 1s intensity-exposure curves are discussed on the basis of an autocatalytic process with a temperature dependent precursor state.  相似文献   

11.
S. Zalkind  N. Shamir 《Surface science》2007,601(5):1326-1332
In the 310-790 K temperature range, the mechanism of initial oxidation by O2 is oxide island nucleation and growth. At the lower temperature range, oxygen is first chemisorbed and the oxide nucleates at coverage of ∼0.2. Increasing the temperature causes the oxide islands to nucleate at lower coverage and at 700 K and above, the oxide nucleates without any significant stage of chemisorbed oxygen. The temperature dependence shows that while the dissociation stage is not activated, the oxide nucleation and growth are thermally activated. Also, opposite to O2 adsorption, the initial H2O adsorption and oxidation rate was found to decrease with temperature. Opposite to the oxygen case, upon exposure to water vapor there is no noticeable stage of chemisorbed oxygen (or OH) and oxide is directly nucleated. Only after oxide coalescence, this tendency changes and the oxidation rate is increased with temperature.  相似文献   

12.
X-ray and ultraviolet photoelectron spectroscopic results are reported for the interaction of CH3OH with clean polycrystalline Al in the temperature range 110–500 K. Methanol is moleculary chemisorbed at low exposure and low temperature (110 K) followed by condensation at higher exposure. Bonding mechanisms and geometries in the condensed and chemisorbed layers are discussed. The multilayers desorb beginning near 170 K and the chemisorbed layer is converted into a surface methoxide. Room temperature adsorption also leads to formation of the methoxide species which is stable to ~500 K, at which point it decomposes evolving CH4 and leaves the surface oxidized.  相似文献   

13.
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   

14.
Kevin Summers 《Surface science》2007,601(6):1443-1455
The surface reactions of 2-iodopropane ((CH3)2CHI) on gallium-rich GaAs(1 0 0)-(4 × 1), was studied by temperature programmed desorption (TPD) and X-ray photoelectron spectroscopy (XPS). CH3CHICH3 adsorbs molecularly at 120 K but dissociates below room temperature to form chemisorbed 2-propyl ((CH3)2CH) and iodide (I) species. Thermal activation causes desorption of the molecular species at 240 K, and this occurs in competition with the further reactions of the (CH3)2CH and I chemisorbed species. Self-coupling of the (CH3)2CH results in the formation of 2,3-dimethylbutane ((CH3)2CH-CH(CH3)2) at 290 K. β-Hydride elimination in (CH3)2CH yields gaseous propene (CH3CHCH2) at 550 K while reductive elimination reactions of (CH3)2CH with surface hydrogen yields propane (CH3CH2CH3) at 560 K. Recombinative desorption of the adsorbed hydrogen as H2 also occurs at 560 K. We observe that the activation barrier to carbon-carbon bond formation with 2-propyls on GaAs(1 0 0) is much lower than that in our previous investigations involving ethyl and 1,1,1-trifluoroethyl species where the β-elimination process was more facile. The difference in the surface chemistry in the case of 2-propyl species is attributable to its rigid structure resulting from the bonding to the surface via the second carbon atom, which causes the methyl groups to be further away from the surface than in the case of linear ethyl and 1,1,1-trifluoroethyl species. The β-hydride and reductive elimination processes in the adsorbed 2-propyl species thus occurs at higher temperatures, and a consequence of this is that GaI desorption, which is expected to occur in the temperature range 550-560 K becomes suppressed, and the chemisorbed iodine leaves the surface as atomic iodine.  相似文献   

15.
Inelastic electron tunneling spectroscopy (IETS) has been used to probe the irreversible chemisorption of H2O, HCOOH, and CH3COOH on a thin amorphous film of Al2O3. The “clean” Al2O3 film was also probed in a similar way. The measurement involves examining the second derivative of the I–V curve for electrons tunneling between the metal electrodes in an AlAl2O3Pb junction. When chemisorbed species are present on the Al2O3, the second derivative measurement is a representation of vibronic excitations of the chemical bonds in the adspecies, i.e. the tunneling spectra are analogous to IR spectra. A “clean” Al2O3 surface is found to possess free OH groups essentially equivalent to a surface exposed to large concentrations of H2O. Saturation coverages of HCOOH and CH3COOH yield similar spectra to each other, as expected. However, chemical shifts as well as variations in intensity among the equivalent bonds in the adsorbed HCOOH and CH3COOH are noted. Tentative assignments are given for the observed peaks, and it is concluded that the acids adsorb as formate and acetate ions on the surface. Tunneling spectra for several submonolayer coverages of HCOOH (θ = 0.03, 0.08, 0.3 and 0.4) have been measured in addition to the spectrum corresponding to saturation coverage. Small shifts in certain vibronic transitions as a function of surface coverage have been noted. The shifts in the peak positions are related to variations in adsorbate-adsorbent and adsorbate-adsorbate interactions, and the variations in relative peak intensity are associated with the orientation of the adsorbed molecules. The kinetics of the chemisorption of HCOOH, i.e. the sticking probability as a function of surface coverage, has also been determined for surface coverages between 0.03 and 0.4 monolayer.  相似文献   

16.
The adsorption and desorption of glycine (NH2CH2COOH), vacuum deposited on a NiAl(1 1 0) surface, were investigated by means of Auger electron spectroscopy (AES), low energy electron diffraction (LEED), temperature-programmed desorption, work function (Δφ) measurements, and ultraviolet photoelectron spectroscopy (UPS). At 120 K, glycine adsorbs molecularly forming mono- and multilayers predominantly in the zwitterionic state, as evidenced by the UPS results. In contrast, the adsorption at room temperature (310 K) is mainly dissociative in the early stages of exposure, while molecular adsorption occurs only near saturation coverage. There is evidence that this molecularly adsorbed species is in the anionic form (NH2CH2COO). Analysis of AES data reveals that upon adsorption glycine attacks the aluminium sites on the surface. On heating part of the monolayer adsorbed at 120 K is converted to the anionic form and at higher temperatures dissociates further before desorption. The temperature-induced dissociation of glycine (<400 K) leads to a series of similar reaction products irrespective of the initial adsorption step at 120 K or at 310 K, leaving finally oxygen, carbon and nitrogen at the surface. AES and LEED measurements indicate that oxygen interacts strongly with the Al component of the surface forming an “oxide”-like Al-O layer.  相似文献   

17.
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   

18.
The decomposition of formic acid was studied on a clean Ru(101&#x0304;0) surface adsorption temperature between 100 and 460 K by means of flash thermal desorption. The decomposition products observed were H2, CO2, H2O and CO. HCOOH itself was also desorbed, although at low exposures no formic acid was observed. The H2 and CO2 products were desorbed in identical first order peaks, with a peak temperature of 395 K. The H2O product desorbed in a second order peak at 813 K, in contrast to H2O desorption from low coverage H2O adsorption which occurs in two peaks in the region of 220 and 265 K. The CO product desorbed in a first order peak at 488 K, identical to CO from CO adsorption. The dependence of the product peaks on adsorption temperature of the Ru surface was also studied. These results suggest a model involving the formation and decomposition of a surface intermediate species.  相似文献   

19.
Chemisorption of methanol on the Si(111)(7 × 7) surface has been studied at ~ 300 K using high-resolution electron energy loss spectroscopy. Methanol reacts with the Si(111) surface to form the stable surface species SiOCH3 and SiH. The methoxy species (CH3O) is bonded to the Si surface with a covalent bond formed between its oxygen end and the dangling bond of the Si(111) surface atom. A structural model for methanol chemisorbed on the Si(111) surface is proposed.  相似文献   

20.
The formation of water by the reaction of preadsorbed oxygen with hydrogen on a Pt(111) surface has been characterized, using secondary ion mass spectroscopy, below the desorption temperature of H2O (180 K). The concentration of chemisorbed water was monitored during the reaction by following the SIMS H3O+ signal. Reaction profiles were measured over a temperature range of 120 to 153 K, and an H2 pressure range of 10-9 to 10-6 Torr. Under all conditions the reaction profiles were characterized by an induction time, a region of rapid reaction, and finally a steady decline in the rate. In the rapid region, an overall activation energy of 2.9 ± 0.3 kcalmol-1 and a half-order H2 pressure dependence were observed. At low initial oxygen concentrations the induction time increased and the maximum rate decreased. The reaction was slow in the absence of gas phase hydrogen, even when the surface coverage of hydrogen was relatively high. Water and hydrogen thermal desorption spectra, measured after stopping the reaction by removal of gas phase hydrogen, were complex functions of the H2 exposure, exhibiting several peaks between 170 and 400 K. However, after an exposure large enough to drive the reaction to completion, only one H2O peak at 173 K and one H2 peak at 350 K were observed. The results indicate that only a fraction of the total H(a) on the surface was readily available for reaction during H2 exposure at T ? 153 K. the remainder either recombined to form H2 or reacted with O(a) during the thermal desorption ramp. There is good evidence for a surface rearrangement during the induction period. A model is proposed which involves the formation of water clusters that accelerate the rate.  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