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 734 毫秒
1.
The chemisorption of C2H4 and C2D4 on Pd(111) at 150 K has been studied by high resolution electron energy loss spectroscopy. Analysis of the vibrational spectra indicates that (i) C2H4 is more weakly bound on Pd(111) than on Ni(111) and Pt(111) and (ii) softened and broadened CH stretching frequencies suggest hydrogen bond-like interactions between the molecule and the metal surface.  相似文献   

2.
Despite the application of a variety of surface sensitive techniques to the adsorption of simple hydrocarbons on well characterized metallic surfaces, no consistent picture has appeared. We review briefly the published spectroscopic results of ultraviolet photoelectron spectroscopy (UPS) and electron energy loss spectroscopy (EELS) which probe, respectively, the electronic and vibrational structure of the surface-molecular complex, and we consider appropriate free molecular analogues, not only in their ground state but also in their first excited states. A simplified approach to determine the chemisorption geometry from UPS level shifts and EELS is presented. The technique allows an isolation of distortion induced shifts from the total relaxation shift, and we find that the true relaxation shift is rather constant, approximately 2.1 eV for the cases considered. These shifts can then be used to estimate the distance of the molecule to the surface. We concentrate primarily on four systems, C2H2 and C2H4 on Ni(111) and Pt(111), adsorbed at low temperature (below the onset of dissociation). Depending on the metal, the hydrocarbon can adsorb in a di-σ arrangement or with a distortion resembling the lowest energy configuration of the first excited state of the free molecule. We also consider briefly C2H4 on Ag and Cu in which no distortion occurs. The distortions that resemble the first excited states might occur as a consequence of donation of bonding (backbonding) electrons from (to) the normally filled π (empty π1) to (from) the empty (filled) d-band states of the metal. The net effect on the hydrocarbon to partially empty the π level and fill the π1 level, is analogous to a low excitation of the free molecule, π → π1. For C2H4 (planar in the ground state), the lowest excitation is the triplet T-state (3–4 eV) of minimal energy for a 90° twisted configuration with a lengthened C-C bond. Acetylene is a linear molecule in the ground state, but cis- or trans-bent for the triplet excitations, ~a (5.2 eV) or ~b (6.0 eV), respectively. Chemisorbed geometries derived from these configurations seem possible for C2H4 on Ni(111) and C2H2 on Pt(111), while interchanging the adsorbates and substrates gives di-σ bonding, (sp3 hybridization), as proposed previously in the literature. For C2H4 on Ni(111), two of the hydrogens are twisted into the surface which leads to a softening of the CH vibrational frequency. For the four systems considered, the data are consistent with the C-C bond essentially parallel to the surface, but tilted orientations are not ruled out. While the models are clearly oversimplified, they suggest an interesting point of departure for likely chemisorption geometries. Also, some intriguing correspondences to the (presumed) location of the normally empty π1 level and the d-band are noted.  相似文献   

3.
Acetonitrile (CH3CN) coordination to a Pt(111) surface has been studied with electron energy loss vibrational spectroscopy (EELS), XPS, thermal desorption and work function measurements. We compare data for the surface states with known acetonitrile coordination complexes. For CH3CN adsorbed on Pt(111) at 100 K, the molecule is rehybridized and adsorbs with the CN bond parallel or slightly inclined to the surface plane in an η2(C, N) configuration. The ν(CN) frequency is 1615 cm?1 and the C ls and N ls binding energies are 284.6 eV and 397.2 eV respectively. By contrast, weakly adsorbed multilayer acetonitrile exhibits a ν(CN) vibrational frequency of 2270 cm?1, and C ls and N ls binding energies of 286.9 eV and 400.1 eV respectively. Both the EELS and XPS results are consistent with rehybridization of the CN triple bond to a double bond with both C and N atoms of the CN group attached to the surface. In addition to this majority η2(C, N) monolayer state, evidence is found for a second, more strongly bound minority molecular state in thermal desorption spectra. As a result of the low coverage of this state, EELS was unable to spectroscopically identify it and we tentatively assign it as an η4(C, N) species associated with accidental step sites. By contrast to the surface complexes, almost all of the known platinum-nitrile coordination complexes are end-bonded via the N lone-pair orbital. Several cases of side-on bonding are known, however, and we compare the results with the known complex Fe32-NCCH3)(CO)9. The difference in the coordinative properties of a Pt(111) surface versus a single Pt atom must be due to the increased ability of multi-atom arrays to back-donate electrons into the π1 system of acetonitrile. Previously published EELS and XPS results for monolayer acetonitrile on Ni(111) and polycrystalline films are almost identical to the present results on Pt(111). We believe that the monolayer of CH3CNNi(111) is also an η2(C, N) species, not an end-bonded species previously proposed by Friend, Muetterties and Gland.  相似文献   

4.
The adsorption of ammonia on the Ni(110) and Ni(111) surfaces has been studied with high resolution (≤ 65 cm?1) electron energy loss spectroscopy (EELS) combined with thermal desorption spectroscopy. The EELS spectra of the initial chemisorbed layer or α state on each surface are very different. Ammonia chemisorbed on the Ni(110) surface exhibits a strong Ni-N stretching mode at 570 cm?1 which is absent on the Ni(111) surface. The Ammonia adsorption site appears to be different on the Ni(110) and Ni(111) surfaces. We suggest that the absence of the M-N stretching mode on the Ni(111) surface is a general characteristic of the ammonia adsorption site on the (111) surfaces of fcc Group VIII metals.  相似文献   

5.
Clean and oxygen covered {111} recrystallized Pt surfaces were studied by EELS after surface preparation at 150≤T≤165OK. The clean surface shows Stokes as well as anti-Stokes lines of surface phonons at ±195?1. Adsorption of small amounts of (<10?2 monolayers) of O2 or H2 leads to substrate-derived phonon losses at ±380cm?1. Oxygen exposure at different pressures, times and temperatures leads to atomic and/or molecular adsorption as well as oxide-related features which have been identified by EELS.  相似文献   

6.
《Surface science》1986,175(2):241-248
EELS spectra of CH4, CD4 and CH2D2 physisorbed on NaCl(100) at 40 K are presented. For the clean NaCl surface, increasing the incident beam energy to the 30 to 50 eV range is sufficient to overcome charging effects. However, when adsorbate is present charging prevents extended signal averaging. All methane modes appear to contribute to the EELS spectra. Loss peaks corresponding to adsorbate vibrations are very broad (400 to 600 cm−1, FWHH), the low resolution being probably due to a combination of effects including surface disorder, scattering of electrons by gas phase molecules and charging effects.  相似文献   

7.
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   

8.
CO adsorption on the (111) face of a Pt10Ni90 alloy single crystal has been investigated at room temperature by vibrational electron energy loss spectroscopy (EELS) and photoelectron spectroscopy (XPS and UPS). Two well separated CO stretching modes develop at 2070 and 1820 ± 10 cm?1, with their intensities reaching 64 and 36% respectively of the total intensity at saturation coverage. They are attributed to CO adspecies in terminal and bridge bonded configuration respectively. The UPS spectra of 4σ, 5σ and 1π molecular orbitais of adsorbed CO show complex features which may be resolved into two components having the main characteristics of CO adsorbed on pure Pt(111) and Ni(111) respectively. Such behaviour is also observed by XPS on C 1s on O 1s peaks. Their respective contributions, in both XPS and UPS spectra are about 64 and 36% of the whole spectrum. Finally compared to Ni(111) — on which CO adsorbs mainly in bridge configuration — the alloying with 10% Pt has generated the appearance of a large number of new sites for CO chemisorption associated with the presence of Pt atoms at the surface. The large amount of terminal CO adspecies is interpreted in terms of considerable surface enrichment of the alloy in platinum.  相似文献   

9.
On the surface of NaF the adsorption isotherms of H2O, D2O, CH3OH, C3H3OH and 1-C3H7OH as well as the infrared spectra of H3O, D2O, dilute HDO, CH3OH and CH3OD were measured. The adsorption temperatures of H3O (253–308 K) were within the phase transition region where two phases of low and high density coexist, while those of CH3OH, C2H5OH and 1-C3H3OH were yet within a super-critical region. The entropy of the 2D condensed H2O on NaF was found to be 14.0 cal K?1 mol?1, which suggests that the condensed phase of water on NaF is liquid-like. The OD stretching band of dilute HDO in the 2D condensed water gives a maximum adsorption at ca. 2530 cm?1 with a half width of ca. 150 cm?1, being in good agreement with that in liquid water. Comparison of the integrated absorbance of the D2O bending mode with that of the OD stretching mode suggests that the cluster size of the 2D condensed water on NaF decreases with increasing temperature. The 2D critical temperature and the occupied areas of these adsorbates enable us to conclude that the compatibility of the molecular size with the surface lattice is not important in the occurrence of the 2D condensation of the hydrogen-bonding molecules on NaF and that adsorbed molecules are randomly oriented on the surface to the extent similar to that in 3D liquid state.  相似文献   

10.
The inelastic neutron scattering spectra of C2H2 and C2D2 adsorbed on a Ag+ exchanged 13X zeolite (0–800 cm?1) and of C2H2 on the Na+ form (0–300 cm?1) have been obtained. For the Na-13X system no distinct vibrational modes were observed, however for the Ag-13X systems the low frequency intramolecular modes of the adsorbed gas and some of the vibrations of the adsorbed gas relative to the surface have been assigned. From the deuteration shifts it appears that C2H2 and C2D2, adsorbed on Ag-13X, are non-linear.  相似文献   

11.
《Surface science》1986,167(1):101-126
The kinetics and mechanism of the decomposition of methanol (CH3OD) on oxygen-covered Pt(111) were studied using static secondary ion mass spectrometry (SIMS) and temperature programmed desorption (TPD). The initial sticking coefficient and the saturation first layer coverage of CH3OD are unity and 0.36 ML, respectively. The maximum amounts decomposed in TPD on O/Pt(111) and clean Pt(111) are 0.19 and 0.047 ML, respectively. At low methanol coverages (< 0.05 ML) on O/Pt(111) the only reaction products were CO2, H2O and D2O, whereas at saturation CO, H2O, D2O and H2 were observed. The decomposed amount did not saturate at or before the onset of molecular methanol desorption, but increeased monotonically until saturation of the first layer. No oxygen exchange was observed between CH3OD and preadsorbed 18O. A decomposition mechanism involving methoxy and hydroxyl type species is proposed. Methanol coverages as low as 0.002 ML could be detected with SIMS. The CH3+ ion was the most intense ion in the positive SIMS spectrum of both methanol and methoxy. Larger ion clusters such as (CH3OD)n+ (n = 2, 3) developed successively at specific multilayer coverages. At low coverages on O/Pt(111), methoxy formation occurs above 125 K and its decomposition becomes detectable above 150 K during temperature programming. In the isothermal mode, the SIMS CH3+ ion was used to follow the kinetics. Over the temperature range 120–145 K, the second order Arrhenius rate parameters for methoxy formation are E = 5.5±0.7 kcal/mol and A = 1.5×10−7±0.6 cm2/s·molecule for an initial methanol coverage of 0.05 ML. Methoxy decomposition was studied in the temperature range 150–165 K and at an initial coverage of 0.015 ML. The first order kinetic parameters, E = 11.4±0.5 kcal/mol and A = 5.3×1013±1 s−1 were derived. Advantages and limitations of using SIMS as a tool for kinetic studies are discussed.  相似文献   

12.
K.E. Lu  R.R. Rye 《Surface science》1974,45(2):677-695
The adsorption and flash desorption of hydrogen and the equilibration of H2 and D2 has been studied on the (110), (211), (111) and (100) planes of platinum. Desorption from Pt (211), a stepped surface composed of (111) and (100) ledges, yields a desorption spectrum which apparently is a composite of desorption from the individual ledges. Pt (110) is quite similar to the tungsten structural analog, W (211), in that both yield two-peak desorption spectra, and on both planes adsorption kinetics are dramatically different for filling of the two states. On all four planes adsorption kinetics are apparently proportional to (1 ? θ)2, and estimates of the initial sticking probabilities show them to decrease in the order: (110) > (211) > (100) > (111). Equilibration activity follows approximately the same order [(110) > (211) > (111) > (100)] with a factor of ~ 5 difference between the most and least active planes; no extraordinary activity is observed for the stepped surface, Pt(211). Below ~ 570 K equilibration of H2 and D2 is activated by less than 2 kcal/mole with the magnitude dependent on the specific face, and above this temperature the reaction is nonactivated. The non-activated case apparently results from absorption followed by statistical mixing on the surface. Calculated rates for HD production per cm2 based on this model are in excellent agreement with the experimental values for Pt(110) and Pt(211), and in somewhat poorer agreement in the case of Pt (111) and Pt (100). This latter is probably due to the greater inaccuracy in the values of the sticking coefficients on these planes.  相似文献   

13.
The decomposition of D2CO, CH3OD and HCOOH on Pt(110) and of D2CO on Pt(S)-[9(111) × (100)] was studied by molecular beam relaxation spectroscopy. D2CO and CH3OD evolved CO and H2 via a desorption limited sequence of elementary steps. The rate constant for CO desorption from Pt(110) was 6 × 1014exp(? 35.5 kcalgmol · RT) s?1, and from Pt(S)-[9(111) × (100)] it was 1 × 1015 exp(?36.2 kcalgmol·RT) s?1. On Pt(110) the rate constant for hydrogen formation was 100 ± 1exp(?24 kcalgmol·RT) m?2atom · s. On Pt(S)-[9(111) × (100)] two pathways for H2 formation existed with rate constants of 8.7 × 10?2exp( ?24.9 kcalgmol· RT) cm2atom· s and 3.2 × 10?3 exp(?19.5 kcalgmol·RT) cm2atom· s. These pre-exponential factors are in order of magnitude agreement with values typical of hydrogen recombination on other metals. When a small amount of sulfur ( ~ 0.1 ML) was adsorbed on the stepped Pt surface, only one pathway for H2 formation existed due to blockage of stepped sites. A similar result was obtained when a beam of CO was impinged on the surface. Formic acid decomposed via a branched process to form primarily CO2 and H2.  相似文献   

14.
The diffusion constants for C and O adsorbates on Pt(111) surfaces have been calculated with Monte-Carlo/Molecular Dynamics techniques. The diffusion constants are determined to be DC(T)=(3.4 × 10?3e?13156T)cm2s?1 for carbon and DO(T) = (1.5×10?3 e?9089T) cm2 s?1 for oxygen. Using a recently developed diffusion model for surface recombination kinetics an approximate upper bound to the recombination rate constant of C and O on Pt(111) to produce CO(g) is found to be (9.4×10?3 e?9089T) cm2 s?1.  相似文献   

15.
The molecule styrene-β-D2 has been prepared. The liquid-phase infrared spectrum in the region 400 to 3500 cm?1 and the laser Raman spectrum have been recorded. Vibrational assignments for this molecule have been made largely by comparison with those of Condirston and Laposa (2) for C6H5CHCH2, C6H5CDCD2, C6D5CHCH2, and C6D5CDCD2.  相似文献   

16.
We employed tunable diode laser absorption spectroscopy to measure the line strength, the methane (CH4), ethane (C2H6) and the propane (C3H8) broadening coefficients for the 523–422 H2O transition at 3619.61 cm?1. Water amount fractions generated by a stable and accurate humidity transfer standard, traceable to the SI units via the German national humidity standard, were used to calibrate the spectroscopic line strength measurements. We focus on the traceability of the measured line data to the SI and on uncertainty assessments following the guidelines of the Guide to the Expression of Uncertainty in Measurement. We determined the line strength to be (8.42 ± 0.07)×10?20 cm?1/(cm?2 molecule) corresponding to a relative uncertainty of ±0.8%. To the best of our knowledge, we report the first methane, ethane and propane broadening coefficients of (8.037 ± 0.056)×10?5 cm?1/hPa, (9.077 ± 0.064)×10?5 cm?1/hPa and (10.469 ± 0.073)×10?5 cm?1/hPa for the 523–422 H2O transition at 3619.61 cm?1, respectively. The relative combined uncertainties of the stated CH4, C2H6 and C3H8 broadening coefficients are in the ±0.7% range.  相似文献   

17.
Diazirine is one of the seven possible isomers of diazomethane. The medium-resolution infrared spectrum of this cyclic compound, which seems to be very stable in the gas phase, was reported by Ettinger. The mid-infrared spectra of diazirine and of the substituted isotopic species D2CN2, H213CN2, and H2C15N2 were recorded with a resolution of 0.08 cm?1. The overall assignment of these spectra is reported here. The ν3 fundamental of the main species at 1459.15 cm?1 (CH2 deformation) is an A-type parallel band but presents a complicated band structure since diazirine is an asymmetric rotor with κ = ?0.427. The rovibrational assignment and the analysis of this band, together with the determination of the molecular constants, is given.  相似文献   

18.
We have measured the intensities of the 992 cm?11) and 1032 cm?112) lines in the Raman spectra of pyridine multilayers adsorbed on a Ni(111) surface, as a function of multilayer thickness. These experiments have been carried out in a standard ultra high vacuum chamber with Auger electron spectroscopy and mass spectroscopy capabilities. We find that the intensity of the 992 cm?1 line decreases linearly with multilayer thickness down to ~ 5 layers. From these data and spectra of submonolayer quantities of Nitrobenzene on Ni(111) we predict an observable intensity for unenhanced spectra for monolayer amounts of pyridine on Ni(111). Attempts at observation of monolayer pyridine on this surface have been unsuccessful. We conclude that the scattering intensity of chemisorbed pyridine on Ni(111) is not enhanced and is likely less than would be predicted from the gas phase cross section.  相似文献   

19.
Molecular vibrations of C2H2 and C2D2 adsorbed on Pt(111) at 140 K and ∼300K have been measured by high resolution electron energy loss spectroscopy. The comparison of C2H2 and C2D2 spectra allows an unambiguous assignment of the observed losses to the excitation of C−H bending, C−H stretching, and C−C stretching modes of nondissociatively adsorbed acetylene. From the relative intensities of losses the hybridisation state is determined to be nearsp 2. The C−C stretching frequency indicates a C−C bond order of ∼1.8.  相似文献   

20.
Two potential models for acetylene are developed and tested by comparison between variational calculations for the stretching vibrational term values and available spectroscopic data. The first model based on local bond potentials with harmonic interbond coupling gives root mean square deviations of 6 cm?1 for C2H2 and 3 cm?1 for C2D2. The second model is more ambitious, being designed to reproduce the dissociation characteristics of the molecules, and the calculated root mean square deviations from the experimental vibrational term values are larger, 32 cm?1 for C2H2 and 24 cm?1 for C2D2. The eigenvalue spectrum of C2H2 is shown to differ from that of C2D2 in showingmarked local mode features and this difference in behaviour is underlined by means of a correlation diagram. Finally it is shown how the known normal mode frequencies and anharmonic constants may be introduced into a simple model in order to predict the excited term values of C2H2, again with a root mean square deviation of 6 cm?1.  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