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 859 毫秒
1.
C2H4在清洁和有Cs覆盖的Ru(0001)表面吸附的TDS研究   总被引:1,自引:0,他引:1       下载免费PDF全文
用热脱附谱(TDS)方法研究了乙烯(C2H4)在Ru(0001)表面上的吸附.在低温下(200K以下)乙烯可以在清洁及有Cs的Ru(0001)表面上以分子状态稳定吸附,在衬底温度升高至200K以上时,乙烯发生了脱氢分解反应,乙烯分解后的主要产物为乙炔(C2H2).在清洁的Ru(0001)表面,乙烯有两种吸附状态,脱附温度分别为275K和360K.而乙炔的脱附温度为350K.在Ru(0001)表面有Cs的存在时,乙烯分解 关键词: 乙烯 钌(0001)表面 铯钌(0001)表面乙烯 钌(0001)表面 铯钌(0001)表面  相似文献   

2.
《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   

3.
The behavior of dimethyl methylphosphonate (DMMP), dosed at 100 K with and without coadsorbed water on oxidized iron has been examined by temperature programmed desorption (TPD) and Auger electron spectroscopy (AES). Molecular and dissociated states of DMMP are readily distinguished by the P(LMM) Auger lineshape. At low coverages DMMP undergoes complete decomposition during heating, leaving carbon, phosphorus and oxygen residues on the surface. The major low temperature decomposition products are CH3OH, H2O, CO, H2 and a surface phosphate species. The DMMP decomposition is limited and large exposures lead to molecular DMMP desorption characteristics of multilayers (200–210 K). Pre-exposure to H2O increases the extent of DMMP decomposition.  相似文献   

4.
The chemisorption of C2H2 and C2H4 on a clean or partly C- or O-covered Fe(111) surface was investigated with AES, TDS and HREELS. On the clean surface, both molecules adsorb under strong rehybridization close to sp3. Above 230 K, C2H2 reacts to form CH and presumably CH2 as the main products, which on further heating decompose to yield H2 desorption maxima at 580 and 490 K, leaving two carbon species on the surface which correspond to two loss peaks at 400 and 1290 cm?1 in the HREELS spectrum. C2H4 undergoes very rapid decomposition above 250 K; no intermediates have been detected. The presence of coadsorbed oxygen or carbon atoms only reduced the maximum uptake of C2H2, but led to the appearance of new molecular adsorption states of C2H4 and inhibited C2H4 decomposition.  相似文献   

5.
The adsorption of ethylene-oxide (Et-O) on Ni(111) was studied with high resolution electron energy loss spectroscopy and angular resolved UV-light induced photoelectron spectroscopy (ARUPS) at 140K; these measurements were complemented by thermal desorption spectroscopy (TDS) and workfunction change measurements ( δφ ).For fractional Et-O monolayer coverages five loss peaks were observed with HREELS at 835, 1155, 1270, 1495 and 3150 cm−1 which are attributed to the C2O ring deformation, CH2 wagging and twisting modes, to the C2O ring breathing, to the CH2 scissor modes and C-H stretching modes of molecular adsorbed Et-O. At low coverage, the HREELS is dominated by the 835 and 3150 cm−1 losses, whereas the 1155, 1270 and 1495 cm−1show only weak intensities. The latter loss peaks increase significantly in intensity for Et-O coverage near the saturation of the first adsorption layer, θ (Et-O)~0.3.UPS measurements confirm the molecular adsorption of Et-O on Ni(111) at 140 K. Compared to the Et-O gas phase UPS, a considerable shift to lower binding energy is observed for the 6a1 oxygen lone pair orbital and also for the 2b1 (n, σCO, σCC) which has some lone pair character. These chemical shifts suggest a bonding of Et-O to Ni(111) through the oxygen atom.  相似文献   

6.
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   

7.
The chemisorption of CO on W(100) at ~ 100K has been studied using a combination of flash desorption and electron stimulated desorption (ESD) techniques. This is an extension of a similar study made for CO adsorption on W(100) at temperatures in the range 200–300K. As in the 200–300 K CO layer, both α1-CO and α2-CO are formed in addition to more strongly bound CO species upon adsorption at ~ 100K; the α-CO states yield CO+ and O+ respectively upon ESD. At low CO coverages, the α1 and α2-CO states are postulated to convert to β-CO or other strongly bound CO species upon heating. At higher CO coverages, α1-CO converts to α2-CO upon thermal desorption or electron stimulated desorption. There is evidence for the presence of other weakly-bound states in the low temperature CO layer having low surface concentration at saturation. The ESD behavior of the CO layer coadsorbed with hydrogen on W(100) is reported, and it is found that H(ads) suppresses either the concentration or the ionic cross section for α1 and α2-CO states.  相似文献   

8.
Adsorption of NO and O2 on Rh(111) has been studied by TPD and XPS. Both gases adsorb molecularly at 120 K. At low coverages (θNO < 0.3) NO dissociates completely upon heating to form N2 and O2 which have peak desorption temperatures at 710 and 1310 K., respectively. At higher NO coverages NO desorbs at 455 K and a new N2 state obeying first order kinetics appears at 470 K. At saturation, 55% of the adsorbed NO decomposes. Preadsorbed oxygen inhibits NO decomposition and produces new N2 and NO desorption states, both at 400 K. The saturation coverage of NO on Rh(111) is approximately 0.67 of the surface atom density. Oxygen on Rh(111) has two strongly bound states with peak temperatures of 840 and 1125 K with a saturation coverage ratio of 1:2. Desorption parameters for the 1125 peak vary strongly with coverage and, assuming second-order kinetics, yield an activation energy of 85 ± 5 kcalmol and a pre-exponential factor of 2.0 cm2 s?1 in the limit of zero coverage. A molecular state desorbing at 150 K and the 840 K state fill concurrently. The saturation coverage of atomic oxygen on Rh(111) is approximately 0.83 times the surface atom density. The behavior of NO on Rh and Pt low index planes is compared.  相似文献   

9.
The chemisorption of H2, O2, CO, CO2, NO, C2H4, C2H2 and C has been studied on the clean Rh(111) and (100) surfaces. LEED, AES and thermal desorption were used to determine the surface structures, disordering and desorption temperatures, displacement and decomposition characteristics for each species. All of the molecules studied readily chemisorbed on both surfaces. A large variety of ordered structures was observed, especially on the (111) surface. The disordering temperatures of most ordered surface structures on the (111) surface were below 100°C. It was necessary to adsorb the gases at 25° C or below in order to obtain well-ordered surface structures. Chemisorbed oxygen was readily removed from the surface by H2 or CO gas at crystal temperatures above 50°C. CO2 appears to dissociate to CO upon adsorption on both rhodium surfaces as indicated by the identical ordering and desorption characteristics of these two molecules. C2H4 and C2H2 also had very similar ordering and desorption characteristics and it is likely that the adsorbed species formed by both molecules is the same. Decomposition of ethylene produced a sequence of ordered carbon surface structures on the (111) face as a result of a bulk-surface carbon equilibrium. The chemisorption properties of rhodium appear to be generally similar to those of iridium, nickel and palladium.  相似文献   

10.
Yuhai Hu  Keith Griffiths   《Surface science》2008,602(17):2949-2954
Fourier transform infra red reflection–absorption spectroscopy (FTIR-RAS), thermal desorption spectroscopy (TDS), and auger electron spectroscopy (AES), were employed to explore the mechanism of NO reduction in the presence of C2H4 on the surface of stepped Pt(3 3 2). Both NO–Pt and C2H4–Pt interactions are enhanced when NO and C2H4 are co-adsorbed on Pt(3 3 2). As a result, C2H4 is dissociated at surface temperatures as low as 150 K, and the N–O stretch band is weakened. The presence of post-exposed C2H4 leads NO desorption from steps to decrease significantly, but the same effect on NO desorption from terraces becomes appreciable only at higher post-exposures of C2H4, e.g., 0.6 L and 1.2 L, and proceeds to a much slighter extent. Auger spectra indicate that as a result of the reaction with O from NO dissociation, the amount of surface C species is greatly reduced when NO is post-exposed to a C2H4 adlayer. It is concluded that reduction of NO in the presence of C2H4 proceeds very effectively on the surface of the Pt(3 3 2), through a mechanism of NO dissociation and subsequent O removal. Following this mechanism, the significant dissociation of adsorbed NO molecules on steps at surface temperatures below 400 K, and subsequent rapid reaction between the resultant O and C-related species, accounts for the considerable amount of N2 desorption at temperatures below 400 K.  相似文献   

11.
The adsorption and reaction of C2H4 on oxygen covered Pd(100) was studied with high resolution electron energy loss spectroscopy (EELS) and temperature programmed reaction spectroscopy (TPRS). The clean Pd(100) surface at 300 K was exposed to O2 to produce atomic oxygen in the p(2×2) structure for coverages between 0.05 and 0.25. The EELS and TPRS measurements were conducted following saturation coverage of the oxygen covered surface by C2H4 at 80 K. Both the di-σ- and π-bonded forms of C2H4 were stable on the surface for θO less than 0.25. The π-bonded form desorbed without reaction between 100 and 300 K, but the di-σ-bonded form underwent dehydrogenation above 250 K. The C2H4 dehydrogenation products were reactive towards atomic oxygen and produced H2, H2O, CO, CO2, and adsorbed C. Oxygen preadsorption inhibited C2H4 Oxidation by limiting the formation of di-σ-bonded C2H4, and the fully developed p(2×2)O overlayer, corresponding to θO = 0.25, was sufficient to block completely the reaction of ethylene. The extent of reaction decreased in a 2:1 ratio to the increase in oxygen coverage, and indicated that oxygen islands blocked C2H4 dissociation. Only the π-bonded form of C2H4 was stable on the surface for θO greater than 0.25; the saturation coverage of π-bonded C2H4 of 0.25 was the same as for clean Pd(100).  相似文献   

12.
The adsorption of HNO3/H2O mixtures on Ag(110) was investigated to learn more about the chemistry of the metal/electrolyte interface. The experiments were performed in ultrahigh vacuum (UHV) using thermal desorption spectroscopy (TDS), low energy electron diffraction (LEED), and electron stimulated desorption ion angular distribution (ESDIAD) over temperatures of 80–650 K and coverages of 0–10 monolayers (ML). As this is the first known study of HNO3 in UHV, the mass spectrometer cracking pattern for HNO3 is here reported. HNO3 adsorbs irreversibly on the clean surface at 80 K and loses its acidic proton to form an adsorbed surface nitrate (NO3) below 150 K. The saturation amount of adsorbed NO3 is 0.4 ± 0.1 ML for which adsorption occurs in either a normal or split c(2 × 2) structure. N03 is stable on the surface up to 450 K beyond which it decomposes directly to gaseous NO2 and NO and adsorbed atomic oxygen. NO3 decomposition is first order with an activation energy Ea = 151±4 kJ mol−1 and a pre-exponential factor of A = 1015.4±0.4s−1. NO3 stabilizes adsorbed H2O by about 8 kJ mol−1 and is hydrated by as many as three H2O molecules. Multilayers of HNO3/H2O desorb at 150–220 K and show evidence of extensive hydrogen bonding and hydration interactions. No evidence for HNO3-induced corrosion or other surface damage was detected in any of these experiments.  相似文献   

13.
Raman spectroscopy using a hot stage was used to characterise layered double hydroxides (LDHs) of the formula (Cu,Zn)6Al2(OH)16(CO3)·4H2O. The spectra were used to assess the molecular assembly of the cations in the LDH structure. The sharp band at 1058 cm−1for the Zn6Al2(OH)16(CO3)·4H2O is assigned to the ν1CO32− symmetric stretching mode. This band shifts to higher wavenumbers and is observed at 1103 cm−1at 600 °C. It is proposed that metal carbonate species formed during the decomposition of the hydrotalcite structure is responsible for the increase in the band position. The Cu–Al hydrotalcite did not show the same trend. The symmetric stretching mode of carbonate is observed at around 1110 cm−1, and at temperatures above 200 °C a shoulder appears at around 1210 cm−1, suggested to be due to CuCO3. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

14.
CH3NC adsorption and thermal decomposition on a Pt(111) surface has been studied by high resolution EEL and TD spectroscopies. At 90 K, CH3NC adsorbs initially in a terminal-bonded configuration characterized by a blue-shifted iso-cyanide stretch at 2265-2240 cm?1. At higher coverages this form co-exists with a second form characterized by an imine-like stretch at 1600–1770 cm?1, increasing with coverage. This form is associated with bridge bonding to adjacent surface platinum atoms. Adsorption is irreversible and. except for multilayer desorption at 135 K, only reaction-limited H2 (Tp = 440–460 K) andHCN (Tp = 420–610 K) desorption and. at high coverages, isomerization to CH3CN (Tp = 430 K) was seen. EEL spectra recorded after heating the adsorbed layer indicated that at lower coverages, the molecular integrity of the adsorbed CH3NC was completely lost before dehydrogenation occurred. On the other hand, at saturation structural changes in the adsorbed layer corresponded firstly to the onset of dehydrogenation and then. at higher temperatures, to HCN evolution. No spectroscopic evidence for an η2-bonding configuration was found either at low temperatures or during thermal decomposition. The terminal- and bridged-bonded configurations adopted by CH3NC have been compared and contrasted with those found with the isoelectronic CO and the isomeric CH3CN by reference to the chemically important frontier orbitals of these ligand molecules.  相似文献   

15.
《Solid State Ionics》1987,23(3):183-188
The enthalpies of formation of seven hydrogen vanadium bronzes, HxV2O5(0<x⩽3.77), prepared at ambient temperature by “hydrogen spillover”, have been determined by solution calorimetry. The enthalpy values obtained for their formation from H2(g) and V2O5(s) at 298.15 K are (in kJ mol−1): H0.22V2O5, −1569.95 ± 1.75; H0.46V2O5, −1589.96 ± 1.69; H1.43V2O5, − 1670.53 ± 1.88; H1.87V2O5, −1700.56 ± 1.58; H2.79V2O5, −1744.23 ± 1.72; H3.53V2O5, −1772.98 ± 2.38; H3.77V2O5, −1781.10 ± 2.27. The stabilities of the compounds towards decomposition, disproportion and oxidation are discussed.  相似文献   

16.
Carbon-13 frequency shifts for C2H4, C2D4, and as-C2H2D2 have been measured in isotopic solid solutions in crystalline films at 60 K. All but two of the shifts (for as-C2H2D2) are compatible with recently determined ζ data for C2H4, with 13C frequency shifts for C2H4 and C2D4 in the gas phase and with conventional frequency data. Together, these data completely determine with precision all 18 parameters of the GHFF for ethylene, the previous ambiguity in choice between two sets of Ag species force constants being removed. The force field reproduces closely the observed centrifugal distortion constants for C2H4, a ζ constant observed for trans-C2H2D2, and the inertia defects for C2H4, C2D4, and as-C2H2D2. Vibration and rotation constants for all isotopically deuterated ethylenes are calculated.Possible explanations for the two anomalous crystal shifts in as-C2H2D2 involve the effects of the crystal field, and failure of the use of Dennison's rule for making anharmonic corrections to the shifts. The former explanation is preferred as a result of thorough analysis of the anharmonicity constants for as-C2H2D2 determined from many overtone and combination bands in the gas and crystal spectra.  相似文献   

17.
The adsorption/desorption characteristics of CO, O2, and H2 on the Pt(100)-(5 × 20) surface were examined using flash desorption spectroscopy. Subsequent to adsorption at 300 K, CO desorbed from the (5×20) surface in three peaks with binding energies of 28, 31.6 and 33 kcal gmol?1. These states formed differently from those following adsorption on the Pt(100)-(1 × 1) surface, suggesting structural effects on adsorption. Oxygen could be readily adsorbed on the (5×20) surface at temperatures above 500 K and high O2 fluxes up to coverages of 23 of a monolayer with a net sticking probability to ssaturation of ? 10?3. Oxygen adsorption reconstructed the (5 × 20) surface, and several ordered LEED patterns were observed. Upon heating, oxygen desorbed from the surface in two peaks at 676 and 709 K; the lower temperature peak exhibited atrractive lateral interactions evidenced by autocatalytic desorption kinetics. Hydrogen was also found to reconstruct the (5 × 20) surface to the (1 × 1) structure, provided adsorption was performed at 200 K. For all three species, CO, O2, and H2, the surface returned to the (5 × 20) structure only after the adsorbates were completely desorbed from the surface.  相似文献   

18.
C2H4在Ru(1010)表面吸附与分解的研究   总被引:2,自引:0,他引:2       下载免费PDF全文
用X射线电子能谱(XPS)、热脱附谱(TDS)和紫外光电子能谱(UPS)方法研究了乙烯(C2H4)在Ru(1010)表面的吸附,在低温下(200K以下)乙稀(C24)可以在Ru(1010)表面上以分子状态稳定吸附,在200K以上乙烯(C2H 4)则发生了脱氢分解反应.TDS结果表明乙烯(C2H4)分 解后的主要产物为乙炔(C< 关键词: 乙烯 钌(1010)表面 吸附与分解  相似文献   

19.
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   

20.
C.S. Ko  R.J. Gorte 《Surface science》1985,155(1):296-312
The interactions between oxide support materials and Pt have been studied by incorporating silica, alumina, titania, and niobia into the surface of a clean Pt foil. Auger electron spectroscopy (AES) and temperature-programmed desorption (TPD) of CO and H2 were used for surface characterization. For all of these oxides, TPD indicated no change in the adsorption properties of CO and H2. Peak temperatures were unaffected by the presence of oxide impurities. For silica and alumina, AES results indicated that suboxides could be formed after oxidation at 400 and 800 K respectively. Al2O3 and SiO2 were formed at higher temperatures. Relatively large quantities of these oxides were required to substantially decrease the saturation coverages of CO and H2, indicating that these oxides probably form clusters on the metal surface. For titania and niobia, AES indicated that these oxides dissolved into the Pt above 1300 K, but segregated back to the surface below 500 K. These segregated layers cover the Pt evenly and both oxides completely suppress H2 and CO adsorption at an oxygen coverage of 1 × 1015/cm2. These results are used to discuss the possible reasons for differences in the catalytic properties of Pt on these four oxide supports.  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