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 554 毫秒
1.
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   

2.
This paper is the first of three articles devoted to the CO/Mo(110) chemisorption. The experimental study of adsorption and desorption kinetics was performed by several methods: thermal desorption, low energy electron diffraction and Auger electron spectroscopy. The adsorption of CO on Mo(110) presents two different states. For these two states the desorption kinetics are first order ones, the desorption energies and frequency factors have been determined (E1 = 99 kcal mole?1, E2 = 50 kcal mole?1, v1 = 1019 s?1, v2, = 5 × 1010 s?1). The dependence of sticking coefficient on surface coverage θ was investigated and was found different for the two states of adsorption. LEED shows that the adsorption is not ordered. AES investigation suggests that in the two states C and O have different positions with respect to MO atoms.  相似文献   

3.
The interaction of oxygen with a Pt(110) crystal surface has been investigated by thermal desorption mass spectroscopy, LEED and AES. Adsorption at room temperature produces a β-state which desorbs at ~800 K. Complete isotopic mixing occurs in desorption from this state and it populates with a sticking probability which varies as (1 ? θ)2, both observations consistent with dissociative adsorption. The desorption is second order at low coverage but becomes first order at high coverage. The saturationcoverage is 3.5 × 1014 mol cm?2. The spectra have been computer analysed to determine the fraction desorbing by first (β1) and second (β2) order kinetics as a function of total fractional coverage θ using this fraction as the only adjustable parameter. The β1 desorption commences at θ ~ 0.25 and β1 and β2 contribute equally to the desorption at saturation. The kinetic parameters for β1 desorption were calculated from the variation of peak temperature with heating rate as ν1 = 1.7 × 109 s?1 and E1 = 32 kcal mole?1 whereas two different methods of analysis gave consistent parameters ν2 = 6.5 × 10?7 cm2 mol?1 s?1 and E2 = 29 and 30 kcal mole?1 for β2 desorption. The kinetics of desorptior are discussed in terms of the statistics for occupation of near neighbour sites. While many fea tures of the results are consistent with this picture, it is concluded that simple models considering either completely mobile or immobile adlayers with either strong or zero adatom repulsion are not completely satisfactory. The thermal desorption surface coverage has been correlated with the AES measurements and it has been possible to use the AES data for PtO as an internal standard for calibration of the AES oxygen coverage determination. At low temperature (170 K) oxygen populates an additional molecular α-state. Adsorption into the α- and β-states is competitive for the same sites and pre-saturation of the β-state at 300 K excludes the α-state. This, together with the AES observation that the adsorption is enhanced and faster at 450 than 325 K suggests a low activation energy for adsorption into the β-state.  相似文献   

4.
The interactions between a molecular beam of SiO(g) and a clean and an oxidized tungsten surface were examined in the surface temperature range 600 to 1700 K by mass spectrometrically determined sticking probabilities, by flash desorption mass spectrometry (FDMS) and by Auger electron spectroscopy (AES). The sticking probability, S, of SiO has been determined as a function of coverage and of surface temperature for the clean and the oxidized tungsten surface. Over the temperature range studied and at zero coverage S = 1.0 and 0.88 for the clean and oxidized tungsten surfaces respectively. The results are consistent with both FDMS and AES. For coverage up to one monolayer there is one major adsorption state of SiO on the clean tungsten surface. FDMS shows that Tm = constant (Tm is the surface temperature at which the desorption rate is maximum) and that desorption from this state is described by a simple first order desorption process with activation energy, Ed = 85.3 kcal mole?1 and pre-exponential factor, ν = 2.1 × 1014 sec?1. AES shows that the 92 eV peak characteristic of silicon dominates. In contrast on the oxidized tungsten surface, Tm shifts to higher temperatures with increasing coverage. The data indicate a first order desorption process with a coverage dependent activation energy. At low coverage (θ ? 0.14) there is an adsorption state with Ed = 120 kcal mole?1 and ν = 7.6 × 1019, while at θ = 1.0, Ed = 141 kcal mole?1. This variation is interpreted as due to complex formation on the surface. AES shows that on oxidized tungsten, in contrast to clean tungsten, the dominant peaks occur at 64 and 78 eV, and these peaks are characteristic of higher oxidation states of silicon. Thus, it is concluded that SiO exists in different binding states on clean and oxidized tungsten surfaces.  相似文献   

5.
The effect of adsorbed Na on the surface conductivity, Δσ, and surface recombination velocity, S, of a clean (114)Ge surface is studied. The surface conductivity is a complicated function of the surface Na concentration, NNa; at NNa ≈ 1.5 × 1013 atoms/cm2, it has a minimum; at ca. (3–5) × 1014atoms/cm2, it has a maximum. For a monolayer coverage (ca. 7.2 × 1014atoms/cm2) the values of Δσ are not much different from those of a clean Ge surface. The surface recombination velocity is a three-valued function of the surface potential, US (calculated from the Δσ values), depending on the Na overlayer coverage and heat treatment of the sample. Three different surface structures (LEED data) were found to correspond to the three S versus US curves reported here. Thermal desorption studies show that Na is desorbed in a wide temperature interval. Two peaks have been isolated, studied and discussed. At low coverages a single peak is found to exist, which obeys the first-order desorption kinetics, with a desorption energy of (52 ± 3)kcal/mol. This peak is attributed to the surface defects. For coverages close to14 monolayer a new peak was observed in the spectrum. The desorption energy of this binding state exceeds that of all the other states. When the overlayer coverage is increased, this peak is shifted to higher temperatures, as predicted for a half-order desorption kinetics. By comparing also with LEED data, it may be concluded that this most tightly bound sodium has formed on the Ge(111) surface patches of an ordered structure in which one Na atom is bonded to three Ge atoms.  相似文献   

6.
The adsorption and desorption of nitrogen on a platinum filament have been studied by thermal desorption techniques. Nitrogen adsorption becomes significant only after any carbon contamination is removed from the surface by heating the platinum filament in oxygen, and after the CO content in the background gas is reduced substantially. At room temperature nitrogen populates an atomic tightly bound β-state, E = 19 kcal mole?1. The saturation coverage of the (3-state is 4.5 × 1014 atoms cm?2. Formation of the (β-state is a zero order process in the pressure range studied. At 90 K two additional α1- and α2-desorption peaks are observed. The activation energy for desorption for the α2-state is 7.4 kcal mole?1 at low coverage decreasing to 3 kcal mole?1 at saturation of this state, 6 × 10 molecules cm?2. The maximum total coverage in the α-states was 1.2 × 1015 molecules cm?2. A replacement process between the β- and α-states has been observed where each atom in the (β-state excludes two molecules from the α-state.  相似文献   

7.
The adsorption and desorption of O2 on a Pt(111) surface have been studied using molecular beam/surface scattering techniques, in combination with AES and LEED for surface characterization. Dissociative adsorption occurs with an initial sticking probability which decreases from 0.06 at 300 K to 0.025 at 600 K. These results indicate that adsorption occurs through a weakly-held state, which is also supported by a diffuse fraction seen in the angular distribution of scattered O2 flux. Predominately specular scattering, however, indicates that failure to stick is largely related to failure to accommodate in the molecular adsorption state. Thermal desorption results can be fit by a desorption rate constant with pre-exponential νd = 2.4 × 10?2 cm2 s?1 and activation energy ED which decreases from 51 to 42 kcal/mole?1 with increasing coverage. A forward peaking of the angular distribution of desorbing O2 flux suggests that part of the adsorbed oxygen atoms combine and are ejected from the surface without fully accomodating in the molecular adsorption state. A slight dependance of the dissociative sticking probability upon the angle of beam incidence further supports this contention.  相似文献   

8.
The decomposition of HCOOD was studied on Ni(100). Low temperature adsorption of HCOOD resulted in the desorption of D2O, CO2, CO, and H2. The D2O was evolved below room temperature. CO2 and H2 were evolved in coincident peaks at a temperature above that at which h2 desorbed following H2 adsorption and well above that for CO2 desorption from CO2 adsorption; CO desorbed primarily in a desorption limited step. The decomposition of formic acid on the clean surface was found to yield equal amounts of H2, CO, and CO2 within experimental error. The kinetics and mechanism of the decomposition of formic acid on Ni (110) and Ni(100) single crystal surfaces were compared. The reaction proceeded by the dehydration of formic acid to formic anhydride on both surfaces. The anhydride intermediate condensed into islands due to attractive dipole-dipole interactions. Within the islands the rate of the decomposition reaction to form CO2 was given by:
Rate = 6 × 1015 exp{?[25,500 + ω(ccsat)]/RT} × c
, where c is the local surface concentration, csat is the saturation coverage for the particular crystal plane, and ω is the interaction potential. The interaction potential was determined to be 2.7 kcal/mole on Ni(110) and 1.4 kcal/mole on Ni(100); the difference observed was due to structural differences of the surfaces relating to the alignment of the dipole moments within the islands. These attractive interactions resulted in an autocatalytic reaction on Ni(110), whereas the interaction was not strong enough on Ni(100) to sustain the autocatalytic behavior. Formic acid decomposition oxidized the Ni(100) surface resulting in the formation of a stable surface oxide. The buildup of the oxide resulted in a change in the selectivity reducing the amount of CO formed. This trend indicated that on the oxide surface the decomposition proceeded via a formate intermediate as on Ni(110) O.  相似文献   

9.
《Surface science》1988,195(3):L199-L207
The desorption of deuterium from the low temperature (α2) state on a Pd(110) surface was studied by TDS, LEED and Δφ. Simultaneous measurements of the isothermal desorption of the α2 state, intensity of a half-order beam of the (1 × 2) phase and work function change show clearly that the desorption of this low temperature state is associated with the phase transition (1 × 2) to (2 × 1). This transition correlation with a work function change of ∼ 70 mV. The desorption follows nearly zero-order kinetics with an activation energy of ∼ 71 kJ mol−1. The desorption kinetics are thus closely analogous to those on Ni(110).  相似文献   

10.
The binding states and sticking coefficients of CO and H2 on clean and oxide covered (111)Pt are examined using flash desorption mass spectrometry and Auger electron spectroscopy (AES). On the clean surface at 78 K there is one major binding state of CO with a desorption activation energy which decreases with coverage plus a second smaller state, while H2 exhibits three binding states with peak temperatures of 140, 230 and 310 K and saturation density ratios of 0.5 : 1 : 1. Desorption kinetics of CO are consistent with a first order state with a normal pre-exponential factor of 1013 ± 1 sec?1, while all three peaks of H2 are broader than expected. Interpretations in terms of anomalous pre-exponential factors, coverage dependent desorption activation energies, and desorption orders are considered. On the oxidized surface saturation densities of both gases are nearly identical to those on the clean surface, but desorption temperatures are increased significantly and the initial sticking coefficient on the oxide decreases slightly for CO and increases slightly for H2.  相似文献   

11.
The quasiequilibrium treatment of heterogeneous reactions suggested by Batty and Stickney1) and first order desorption kinetics are combined to formulate a model of the equilibrium coverage of oxygen on tungsten at high temperatures (1200 ? T ? 2500°K) and low oxygen partial pressures (10?9 ? PO2 ? 10?5 Torr). The results are compared with existing data and the agreement is fairly good in view of the extreme simplicity of the theoretical model.  相似文献   

12.
The adsorption of CO on Rh(111) has been studied by thermal desorption mass spectrometry and low-energy electron diffraction (LEED). At temperatures below 180 K, CO adsorbs via a mobile precursor mechanism with sticking coefficient near unity. The activation energy for first-order CO desorption is 31.6 kcal/mole (νd = 1013.6s?1) in the limit of zero coverage.As CO coverage increases, a (√3 ×√3)R30u overlayer is produced and then destroyed with subsequent formation of an overlayer yielding a (2 × 2) LEED pattern in the full coverage limit. These LEED observations allow the absolute assignment of the full CO coverage as 0.75 CO molecules per surface Rh atom. The limiting LEED behavior suggests that at full CO coverage two CO binding states are present together.  相似文献   

13.
The physisorption of Xe on W(111) and of Xe on partial layers of oxygen chemisorbed on W(111) has been studied using flash desorption and work function methods. It has been found that xenon adsorbs up to monolayer coverages at 104K. Xenon desorbs from W(111) as a single binding state following first order kinetics. At low coverages (θXe < 0.07) the binding energy decreases with increasing coverage possibly because of the presence of high energy adsorption sites due to crystal imperfections and edge effects. For θXe > 0.07 the desorption data fit a first order rate expression with a desorption energy of 9.3 kcal/mol and preexponential ν = 1015s?1. The observed work function change of ?1.1 ± 0.1 eV is consistent with monolayer estimates reported in field emission studies of physisorbed xenon on tungsten. The effect of preadsorbed oxygen layers on the physisorption of xenon on this surface is very striking. The energy of desorption shifts as much as 50% higher for a moderate exposure of oxygen. Several physisorption models are explored along with estimates of dispersion and electrostatic interaction contributions.  相似文献   

14.
The adsorption, desorption, and surface structural properties of Na and NO on Ag(111), together with their coadsorption and surface reactivity, have been studied by LEED, Auger spectroscopy, and thermal desorption. On the clean surface, non-dissociative adsorption of NO into the a-state occurs at 300 K with an initial sticking probability of ~0.1, saturation occurring at a coverage of ~120. Desorption occurs reversibly without decomposition and is characterised by a desorption energy of Ed ~ 103 kJ mol?1. In the coverage regime 0 < θNa < 1, sodium adsorbs in registry with the Ag surface mesh and the desorption spectra show a single peak corresponding to Ed ~ 228 kJ mol?1. For multilayer coverages (1 < θ Na < 5) a new low temperature peak appears in the desorption spectra with Ed ~ 187 kJ mol?1. This is identified with Na desorption from an essentially Na surface, and the desorption energy indicates that Na atoms beyond the first chemisorbed layer are significantly influenced by the presence of the Ag substrate. The LEED results show that Na multilayers grow as a (√7 × √7) R19.2° overlayer, and are interpreted in a way which is consistent with the above conclusion. Coadsorption of Na and NO leads to the appearance of a more strongly bound and reactive chemisorbed state of NO (β-NO) with Ed ~ 121 kJ mol?1. β-NO appears to undego surface dissociation to yield adsorbed O and N atoms whose subsequent reactions lead to the formation of N2, N2O, and O2 as gaseous products. The reactive behaviour of the system is complicated by the effects of Na and O diffusion into the bulk of the specimen, but certain invariant features permit us to postulate an overall reaction mechanism, and the results obtained here are compared with other relevant work.  相似文献   

15.
The molecular chemisorption of N2 on the reconstructed Ir(110)-(1 × 2) surface has been studied with thermal desorption mass spectrometry, XPS, UPS, AES, LEED and the co-adsorption of N2 with hydrogen. Photoelectron spectroscopy shows molecular levels of N2 at 8.0 (5σ + 1π) and 11.8 (4σ) eV in the valence band and at 399.2 eV with a satellite at 404.2 eV in the N(1s) region, where the binding energies are referenced to the Ir Fermi level. The kinetics of adsorption and desorption show that both precursor kinetics and interadsorbate interactions are important for this chemisorption system. Adsorption occurs with a constant probability of adsorption of unity up to saturation coverage (4.8 × 1014 cm?2), and the thermal desorption spectra give rise to two peaks. The activation energy for desorption varies between 8.5 and 6.0 kcal mole?1 at low and high coverages, respectively. Results of the co-adsorption of N2 and hydrogen indicate that adsorbed N2 resides in the missing-row troughs on the reconstructed surface. Nitrogen is displaced by hydrogen, and the most tightly bound state of hydrogen blocks virtually all N2 adsorption. A p1g1(2 × 2) LEED pattern is associated with a saturated overlayer of adsorbed N2 on Ir(110)-(1 × 2).  相似文献   

16.
A systematic investigation of the thermal desorption of halogens from well characterized (111), (100) and (110) 4d (Nb, Mo) and 5d (Ta, W) transition metal surfaces has been carried out under low coverage conditions (θ < 10?2 of a monolayer). Characterization of the surfaces was achieved by LEED, AES and work function determinations while the desorption kinetics were recorded in a large temperature range (1700–2300 K) using a pulsed ionic beam method. The new data concerning some Ta and W surfaces are presented and the results of this systematic study are discussed. It is shown that the halogen desorption parameters, e.g., desorption energies and preexponential factors, are independent of both surface structure and d bond filling of the substrate; E(F) ~4.75 eV, E(Cl) ~4.1 eV, E(Br) ~3.7 eV and τ0 ~10?13 ?10?14 S. The halogen behaviour is compared with that of other adsorbates and with the predictions of a general chemisorption model.  相似文献   

17.
At 300 K and in the coverage regime (0<θ<13) bromine chemisorbs rapidly on Pd(111); the sticking probability and dipole moment per adatom remain constant at 0.8 ± 0.2 and 1.2 D, respectively. This stage is marked by the appearance of a √3 structure: desorption occurs exclusively as atomic Br. At higher coverages, desorption of molecular Br2 begins (desorption energy ~130 kJ mol?1) as does the nucleation and growth of PdBr2 on the surface. This latter stage is signalled by the appearance of a √2 LEED pattern and the observation of PdBr2 as a desorption product (desorption energy ~37 kJ mol?1). Some PdBr2 is also lost by surface decomposition and subsequent evaporation of atomic Br. The data indicate that the transition state to Br adatom desorption is localised and that PdBr2(a) ? Br(a) interconversion occurs; these surface species do not appear to be in thermodynamic equilibrium during the desorption process.  相似文献   

18.
NO adsorbs on Pt(111) with a (temperature independent) initial sticking coefficient S0=0.88. The fraction of molecules not being chemisorbed is directly inelastically scattered back due to failure of translational energy accommodation. The nonlinear variation of s with coverage can well be described by a precursor-state model, the precursor state being formed by NO molecules translationally and rotationally accommodated in a physisorbed second layer. Dissociation is essentially restricted to defect sites and is negligible on perfect (111) planes. These defect sites (present in small concentration) are first populated and are also sampled by the modulated beam technique yielding an activation energy for desorption Ed = 33.1 kcal/mole and preexponential factor vd = 1015.5s?1. Isothermal desorption measurements yielded Ed and vd as a function of coverage: Ed rapidly drops from its initial value (at defect sites) to about 27 kcal/mole — which value is considered as representing the adsorption energy on a perfect (111) plane — and then decreases continuously due to effective repulsive interactions. Simultaneously vd is decreasing to about 1012 s?1 at θ = 0.25 which marks the equilibrium coverage to be reached at 300 K. If the surface is precovered with oxygen atoms the NO sticking coefficient is reduced to 0.6, and the desorption parameters are lowered to Ed = 17.1 kcal/mole and vd= 1012.6s?1 (at zero NO coverage).  相似文献   

19.
The flash decomposition of CH3COOH was studied on a clean nickel (110) surface following adsorption at 30° C in order to access the applicability of chemical reaction rate theory to a homologous series of reactants on a well-defined surface. As was observed for formic acid, acetic acid adsorbed at 30° C to yield gaseous H2O and to form islands of adsorbed anhydride intermediates; the decomposition proceeded by a two-dimensional auto-catalytic mechanism to form H2, CO2, Co and surface carbon. The decomposition of the anhydride was rate determining for the formation of CO2 and H2. The rate of decomposition was well described by the equation governing the formic acid decomposition on the same surface. The activation energy for this first order decomposition was determined to be 28.2 kcalgmol and the pre-exponential factor, v, was found to be 6.4 × 1014 s?1 with a fraction of initiation sites of 0.004. These values were nearly the same as those observed for the decomposition of HCOOH, suggesting identical intramolecular mechanisms for the unimolecular decomposition of the adsorbed intermediates. The relative values of v for the decomposition of HCOOH, DCOOH and CH3COOH indicated that the motion of the H, D or CH3 group was involved in the rate-limiting step.  相似文献   

20.
The kinetics of the hydrogen (deuterium) Sorption processes in the α-phase region of a thin electrodeposited Pd layer (thickness 3 × 10?5 cm) are reported. Measurements have been performed with a piezoelectric quartz crystal microbalance using an AT-cut crystal with a resonance frequency f0q = 5.27 MHz and a sensitivity of ~1010 digits g?1, at a constant temperature 80.3 ± 0.05 °C and different pressures. Both the absorption and the desorption of H2(D2) on Pd layers are controlled successively by a chemisorption step (second order kinetics) in the early stage of the processes and by a surface migration step (first order kinetics), in the subsequent stage of the processes. Consistent with the reported piezoelectric quartz crystal microbalance measurements, thermal desorption and electrochemical desorption studies, a mechanism is suggested which takes into account the adsorption ofH(D) atoms on two different surface states: (a) a “pre-dissolved” state which does not depend on the surface nature; (b) and adsorption state, the density of which depends upon the structure and nature of the Pd sample. The kinetic control of the entire processes (sorption and desorption) depends on the ratio of the density of sites (a) to (b).  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