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.
Nitric oxide, strongly chemisorbed on ruthenium, is desorbed almost completely as oxygen and nitrogen. Oxygen, nitrogen, and nitric oxide were observed singly on ruthenium with field emission microscopy. Thermal desorption spectroscopy from Ru(101̄0) showed that molecular nitrogen is only physisorbed but nitrogen from NO decomposition is strongly chemisorbed. Nitrogen from NO shows three binding states, the most strongly bound being present to only a small extent. NO shows three and two binding states when adsorbed at 120 K and 295 K respectively. Work function measurements gave Δφ = 1.3 eV for a monolayer of NO. NO is dissociatively adsorbed above 250 K but a lower temperature limit was not established. The decomposition of NO on Ru(1010) under high vacuum conditions is catalytic in that no oxides of ruthenium were observed to form in the process.  相似文献   

2.
The adsorption of cyclohexane on Ru(001) at 90 K has been investigated by thermal desorption mass spectrometry, EELS, UV photoemission and LEED. Thermal desorption indicates the adsorption of the undissociated molecule first in a chemisorbed monolayer (Td = 200 K) with subsequent formation of multilayers (Td = 165 K) at higher exposures. The vibrational spectrum obtained by EELS is characterized by a frequency shift of the C-H stretching mode from 2920 cm?1 (multilayer) to 2560 cm?1 for the chemisorbed monolayer. Off-specular EELS data indicate two different electron scattering mechanisms for the C-H stretching mode. Whereas for the C-H stretching mode of the multilayer, large angle electron impact scattering is observed, the C-H soft-mode of the monolayer is largely due to small angle dipolar scattering. The He I photoelectron spectra of cyclohexane multilayers are characteristic of the undissociated molecule. A new assignment of C(2s) and the lowest C(2p) level, based on a comparison with benzene, shows that the chemisorbed monolayer is characterized by the absence of emission or broadening of the 2a1u level. This is attributed to C3v symmetry of the chemisorbed layer and to a possible interaction of the 2aIu orbital with the metal surface.  相似文献   

3.
The strong OH stretch at 3400 cm?1 (fwhm ~ 230 cm?1) in the IR reflection absorption spectra of the system H2ORu(001) at 85 K indicates the presence of hydrogen-bonded clusters. These clusters appear to form even at the lowest coverages. On the clean surface there is a linear relationship between integrated absorption intensity and coverage. The presence of small quantities of preadsorbed oxygen delays, however, the onset of absorption. It is thought that the oxygen atoms “bind” the water molecules, thus preventing cluster formation and in turn eliminating the intensity enhancement due to hydrogen bonding. Flash desorption spectra also indicate a second binding state when oxygen is coadsorbed. The relevance of these results to models of the electric double layer at the metal-electrolyte interface is discussed.  相似文献   

4.
The adsorption of alkali metals on transition metals can produce several technologically important effects, but only limited results have been reported on the geometrical structure of such adlayers, especially for adsorption temperatures below 300 K. We have examined the adsorption of Na on Ru(001) as a function of coverage and temperature using LEED to determine the adlayer structure and thermal desorption spectroscopy to characterize binding kinetics and relative Na coverages. The only Na LEED pattern observed following adsorption at 300 K was that of (32 × 32) structure which occurred near saturation of the first layer. However, Na adsorbed at 80 K produces a progression of distinct, ordered LEED patterns with increasing coverage which does not include the (32 × 32) pattern. These patterns result from increasingly compressed, hexagonal arrangements of adsorbate atoms which are uniformly spaced due to mutually repulsive interactions. The order-disorder transition temperature for each structure was also determined by LEED and used to develop a 2D phase diagram for Na on Ru(001). Ordered structures were observed only when Na thermally induced motion was sufficiently limited and the repulsive Na-Na interaction could force the uniform spacing of Na atoms. Thus, low coverage structures only developed where Na mobility was limited by low temperature. High coverage structures were stable to much higher temperatures since motion was inhibited by the high Na density.  相似文献   

5.
The interaction of oxygen with clean Bi(0001) was studied for adsorption between 118 and 296 K using LEED, Auger, electron energy loss, and work function measurements. Oxygen adsorption kinetics show an activated process with a dissociative sticking probability (<10?4) which smoothly decreases over two orders of magnitude up to saturation coverage. Changes in surface electronic properties indicate that an oxide-like bond is formed from the onset of adsorption. There is no evidence for a stable chemisorbed phase. LEED shows simultaneous growth of epitaxial BiO(0001) and a 3 overlayer. At 296 K the adsorption terminates after about three equivalent monolayers of BiO(0001). Periodic trends extended from the transition metal series suggest that local and atomic characteristic rather than long-range electronic properties determine the low reactivity of this surface toward O2.  相似文献   

6.
Chlorine adsorbs on Ru(100) with high efficiency; UPS, Δφ, TD, AES and LEED data suggest that the low pressure interaction ceases with the formation of a dense chemisorbed layer of atomic Cl. This overlayer desorbs exclusively as Cl(g) with an activation energy of ~ 220 kJmol. The failure of the system to grow bulk halide and the absence of molecular chlorine and ruthenium-chlorine species in the desorption products are discussed in thermochemical terms. A series of four ordered surface structures evolves with increasing coverage, and structural models are proposed which are in good quantitative accord with the results from Auger spectroscopy. As coverage deceases, it seems likely that the overlayer relaxes by the growth of ordered vacancies rather than by uniform expansion along a preferred direction.  相似文献   

7.
LEED and AES have been used to study the structural changes and kinetics of the initial interaction between Fe(001) and water vapor at temperatures from 298 to 473 K. A disordered c(2 × 2) structure was formed at all temperatures, and only 80% of the total number of sites were filled at saturation. The initial sticking coefficient was 0.56 ± 0.03, and the reaction rate increased with increasing temperature. A model was proposed that successfully accounted for these experimental observations. Irreversible chemisorption of water is proposed to take place via a precursor of physically adsorbed water molecules. The precursor, which is adsorbed on both bare surface and surface covered by chemisorbed species, is mobile and retains most of its degrees of rotational freedom. Water molecules in the precursor state can either desorb or dissociate, and the difference in activation energies for these reactions was found to be 5.7 ± 0.5 kcalmol. Only 80% of the available c(2 × 2) sites are filled and the surface layer is disordered because the chemisorbed species are immobile, and because each one blocks four nearest neighbor sites for further adsorption. The chemisorbed species occupy the fourfold symmetric sites either above the iron atoms or above the interstitial “holes” betweeh iron atoms.  相似文献   

8.
The adsorption of oxygen and the interaction of carbon monoxide with oxygen on Ru(101) have been studied by LEED, Auger spectroscopy and thermal desorption. Oxygen chemisorbs at 300 K via a precursor state and with an initial sticking probability of ~0.004, the enthalpy of adsorption being ~300 kJ mol?1. As coverage increases a well ordered ¦11,30¦ phase is formed which at higher coverages undergoes compression along [010] to form a ¦21,50¦ structure, and the surface eventually saturates at 0 ~ 89. Incorporation of oxygen into the subsurface region of the crystal leads to drastic changes in the surface chemistry of CO. A new high; temperature peak (γ CO, Ed ~ 800 kJ mol?1) appears in the desorption spectra, in addition to the α and β CO peaks which are characteristic of the clean surface. Coadsorption experiments using 18O2 indicate that γ CO is not dissociatively adsorbed, and this species is also shown to be in competition with β CO for a common adsorption site. The unusual temperature dependence of the LEED intensities of the ¦11,30¦-O phase and the nature of α, β, and β CO are discussed. Oxygen does not displace adsorbed CO at 300 K and the converse is also true, neither do any Eley-Rideal or Langmuir-Hinshelwood reactions occur under these conditions. Such processes do occur at higher temperatures, and in particular the reaction CO(g) + O(a) → CO2(g) appears to occur with much greater collisional efficiency than on Ru(001). The oxidation of CO has been examined under steady state conditions, and the reaction was found to proceed with an apparent activation energy of 39 kJ mol?. This result rules out the commonly accepted explanation that CO desorption is rate determining, and is compared with the findings of other authors.  相似文献   

9.
D. Weaire 《Surface science》1981,103(1):L115-L120
The qualitative variation of the IR absorption spectrum of CO adsorbed at 200 K on Ru(001) in the range of coverage 0 < θ ? 13 is reproduced by a calculation based on a random configuration of adsorbed molecules and nearest neighbour interactions.  相似文献   

10.
A combination of modern surface measurement techniques such as LEED, AES and Thermal Desorption Spectroscopy were used to study the chemisorptive behavior of NO and CO on a (1010)Ru surface. The experimental evidence strongly favors a model in which NO adsorbs and rapidly dissociates into separate nitrogen and oxygen adsorbed phases, each exhibiting ordered structures: the C(2 × 4) and (2 × 1) structures at one-half and full saturation coveilage, respectively. At temperatures as low as 200°C, the nitrogen phase begins to desorb, and continuous exposure to NO in this temperature range results in an increasing oxygen coverage until the surface is saturated with oxygen and no further NO dissociation can take place. The nitrogen desorption spectrum depends strongly on coverage and exhibits several peaks which are related to structure of the adsorbed phase. There is evidence that once the surface is saturated with the dissociated NO phase further NO adsorption occurs in a molecular state. Carbon monoxide adsorbs in a molecular state and does not exhibit an ordered structure. The implications of the results with respect to the catalytic reduction of NO by H2 and CO and the N2 selectivity of Ru catalysts are discussed.  相似文献   

11.
The chemisorption of nitric oxide on (110) nickel has been investigated by Auger electron spectroscopy, LEED and thermal desorption. The NO adsorbed irreversibly at 300 K and a faint (2 × 3) structure was observed. At 500 K this pattern intensified, the nitrogen Auger signal increased and the oxygen signal decreased. This is interpreted as the dissociation of NO which had been bound via nitrogen to the surface. By measuring the rate of the decomposition as a function of temperature the dissociation energy is calculated at 125 kJ mol?1. At ~860 K nitrogen desorbs. The rate of this desorption has been measured by AES and by quantitative thermal desorption. It is shown that the desorption of N2 is first order and that the binding energy is 213 kJ mol?1. The small increase in desorption temperature with increasing coverage is interpreted as due to an attractive interaction between adsorbed molecules of ~14 kJ mol?1 for a monolayer. The (2 × 3) LEED pattern which persists from 500–800 K is shown to be associated with nitrogen only. The same pattern is obtained on a carbon contaminated crystal from which oxygen has desorbed as CO and CO2. The (2 × 3) pattern has spots split along the (0.1) direction as (m, n3) and (m2, n). This is interpreted as domains of (2 × 3) structures separated by boundaries which give phase differences of 3 and π. The split spots coalesce as the nitrogen starts to desorb. A (2 × 1) pattern due to adsorbed oxygen was then observed to 1100 K when the oxygen dissolved in the crystal leaving the nickel (110) pattern.  相似文献   

12.
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   

13.
The reaction of carbon monoxide with oxygen chemisorbcd on polycrystalline platinum has been studied using Auger spectroscopy. Two types of chemisorbed oxygen are distinguished on the basis of Auger electron chemical shifts and reactivity towards carbon monoxide. When the substrate is below 800 K, a single very reactive type of chemisorbed oxygen is formed. Above 800 K a new species begins to form which is characterized by an Auger chemical shift of about 6 eV and by low reactivity. The decay of the oxygen Auger signal using several fixed pressures of carbon monoxide was measured. The reaction is first order in carbon monoxide pressure but no clear decision can be made about the order with respect to oxygen coverage. With the reaction CO + 1202 → CO2 operating at steady-state, the oxygen coverage was measured as a function of CO pressure. In the region 363–600 K, the steady state oxygen coverage began to decline measurably when pCOpO2 reached 0.1. When pCO>pO2the oxygen coverage became immeasurably small. A simple model is used to relate these phenomena to observed carbon dioxide production rates.  相似文献   

14.
In the early stages of reaction with sulfur, a clean Fe{001} surface develops a c(2 × 2) superstructure. A low-energy electron diffraction analysis of this structure leads to a model in which the S atoms lie in the four-fold symmetrical sites on the Fe{001} surface, the S-Fe interplanar spacing being 1.09 ± 0.05 A? and corresponding to an effective radius of 1.06 Å for the chemisorbed S atoms. In contrast to Fe{001} 1 × 1-O, the first interlayer spacing of the substrate here is not significantly expanded.  相似文献   

15.
An extensive photoemission and LEED study of K and CO+K on Ru(001) has been carried out. In this paper the LEED and some XPS results together with TPD and HREELS data are presented in terms of adsorption, desorption. and structural properties, and their compatibility is discussed. Potassium forms (2 × 2) and (3 × 3)R30° overlayers below and near monolayer coverage, and multilayer bonding and desorption is similar to that of bulk K. The initial sticking coefficients for CO adsorption on K predosed surfaces are correlated with the initial K structure, and s0 and CO saturation coverages decrease with increasing K coverage. Two well-characterized mixed CO+K layers have been found which are correlated with predosed (2 × 2) K and (3 × 3)R30° K. They have CO to K ratios of 3:2 and 1:1, and lead to LEED patterns with (2 × 2) and (3 × 3) symmetry, respectively. The molecule is believed to be sp2 rehybridized under the influence of coadsorbed K, leading to stronger CO-Ru and weaker C-O bonds as indicated by the TPD and HREELS results, and to stand upright in essentially twofold bridges.  相似文献   

16.
The structures of Ru(001) and of the √3 × √3 R30° overlayer of CO on Ru(001) have been determined by LEED IV measurements and comparison to calculations. Special attention was paid to accurate angular alignment, selection of a well-ordered portion of the surface, and avoidance of beam-induced changes of the CO layer. Five orders of reflexes over a range of 300 eV each were used for the clean surface and 7 orders over 200 eV each for the CO superstructure. For the clean surface, a slight contraction of the first layer spacing (by 2%) was found which gave r-factors of 0.04 (Zanazzi-Jona) and 0.16 (Pendry) for 5 non-degenerate beams. For the CO structure the most probable geometry is the on-top site with spacings d(RuC) = 2.0 ± 0.1 A?andd(CO) = 1.10 ± 0.1 A? (rZJ = 0.21; rP = 0.51). The two threefold hollow and the bridge sites can be clearly excluded.  相似文献   

17.
High resolution electron energy loss spectroscopy (EELS), thermal desorption mass spectrometry (TDMS) and low energy electron diffraction (LEED) have been used to investigate the molecular chemisorption of N2 on Ru(001) at 75 K and 95 K. Adsorption at 95 K produces a single chemisorbed state, and, at saturation, a (√3x√3) R30° LEED pattern is observed. Adsorption at 75 K produces an additional chemisorbed state of lower binding energy, and the probability of adsorption increases by a factor of two from its zero coverage value when the second chemisorbed state begins to populate. EEL spectra recorded for all coverages at 75 K show only two dipolar modes — ν(RuN2) at 280–300 cm?1 and ν(NN) at 2200–2250 cm?1 — indicating adsorption at on-top sites with the axis of the molecular standing perpendicular to the surface. The intensities of these loss features increase and ν(NN) decreases with increasing surface coverage of both chemisorbed states.  相似文献   

18.
The evolution of a polycrystalline magnesium surface during oxidation at room temperature has been studied by Metastable Impact Electron Spectroscopy (MIES). This technique allowed us to follow the metal-to-insulator transformation of the top layer of the surface. An electronic signal corresponding to a metallic behavior of the surface evidences the presence of under-stoichiometric MgO species on the surface. The total covering by oxygen of the Mg surface uppermost layer, obtained at around 10 L of oxygen deposition, does not correspond to a fully insulating surface. An insulating surface is obtained after 30 L of oxygen deposition. Depositions of CO2 on a clean and a preoxidized polycrystalline Mg surface have been analyzed to give information about the composition of the surface and its evolution. CO2 adsorption in the form of CO32− compounds on preoxidized Mg is more efficient than on clean Mg. Oxygen species, corresponding to chemisorbed oxygen less bounded than oxygen in the MgO lattice, allows the formation of CO32−. Therefore, it is concluded that during oxygen deposition at room temperature, MgO islands and chemisorbed oxygen species coexist on the surface. Moreover, the larger the oxygen predeposition is, the less CO32− compounds are formed, meaning a decrease of available chemisorbed oxygen sites. From oxidation measurements at high temperature (420 K), we show that MgO islands and uncovered Mg domain coexist. Further, no under-stoichiometric compound features have been observed. The high temperature allows the direct formation of oxide MgO species in islands.  相似文献   

19.
Isotope labelling experiments have established that the adsorption of O2 on the W(110) plane at 20 K leads first to the formation of a dissociated atomic layer. A weakly bound molecular species, α-O2, forms only when the atomic layer is essentially complete (O/W = 0.6). The desorption of α-O2 was found to be first order with an activation energy of E = 1.9 kcalmole and a frequency factor γ = 3 × 109 s?1. The activation energy is shown to be less than the enthalpy of desorption and the meaning of this result is discussed.  相似文献   

20.
The polar GaAs(1&#x0304;1&#x0304;1&#x0304;)As surface can be prepared in three stable and ordered states: two by molecular beam epitaxy (MBE), namely the As-stabilized and the Ga-stabilized states and one simply by ion bombardment and annealing at 770 K. The respective LEED structures are (2 × 2), (19 × 19)R23.4°, and (1 × 1) with a diffuse faint (3 × 3) superstructure. Auger measurements and the comparison with the stoichiometric cleaved (110) surface show that there are different As concentrations in the first atomic layer associated with each of these three surfaces. Whereas about 10 to 15% of the first As layer appears to be missing on the (2 × 2) surface, about 50% is missing on the 19 surface. On the (1 × 1) surface the first As layer is removed completely. The intensity of emission from the surface sensitive states between 1 and 4 eV below the valence band edge, as seen by angular resolved UPS, roughly corresponds to the amount of As at the surface thus confirming their interpretation as As-derived surface states. The inital sticking coefficent for oxygen depends strongly on the surface structure: ~10?8 for the (2 × 2), ~10?7 for the 19, and ~10?4 for the (1 × 1) surface. The sticking coefficient does not depend on the surface concentration of As but rather on the degree of saturation of dangling bonds on Ga atoms.  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