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.
The coverages of adsorbed oxygen and CO on an Ir(111) surface have been determined using X-ray photoelectron spectroscopy (XPS) during the steady-state catalytic production of CO2. Correlating the coverages of the reacting adsorbates with the rate of CO2 production allows the kinetics of the CO oxidation reaction to be determined. The reaction is found to obey a Langmuir-Hinshelwood rate expression of the form RCO22 = k0[CO][O]exp(?EakT), where RCO2 is the rate of CO2 production, k0 is the pre-exponential factor of the reaction rate coefficient, [CO] and [O] are the surface coverages of CO and oxygen, respectively, and Ea is the activation energy for the oxidation reaction. The activation energy for this catalytic oxidation reaction is found to be approximately 9 kcalmole.  相似文献   

2.
Self-broadened, air-broadened and CO2-broadened half-widths of lines R(0) through R(0) in the CO fundamental have been measured at 100°K (self-broadening only), 200°K, 250°K and 300°K using the Ladenburg-Reiche curve-of-growth. The relation γ°m(T)γ°m(300°K)=(300T)0.75, which we found previously for the nitrogen-broadened half-widths of R(0), R(8) and R(16), is shown to be valid for all of the line widths measured in the present study.  相似文献   

3.
Reflection absorption infrared spectroscopy has been used in conjunction with LEED and surface potential measurements to study low temperature CO adsorption on the oxidised Cu surfaces Cu(111)O|32?2|, Cu(110)O(2 × 1) and Cu(110)Oc(6 × 2). On all three surfaces adsorption at 80 K yields surface potential changes in excess of 0.6 V and does not lead to the formation of an ordered overlayer. At high coverages the adsorption enthalpy is lower than on the clean surfaces. Infrared spectra show the growth of a doublet band with components initially at 2100 and 2117 cm?1 on the oxidised Cu(111) surface. Similar features seen on the oxidised Cu(110) surfaces are accompanied by a band at 2140 cm?1: a very weak band at the same frequency on oxidised Cu(111) is attributed to defect sites. Studies of the temperature dependence of the spectrum from oxidised Cu(111) lead to the conclusion that two different binding sites are occupied. Spectra of 12CO13CO mixtures show that the molecules occupying these sites are in close proximity to each other, and that the spectrum is subject to large but opposing coverage-dependent frequency shifts.  相似文献   

4.
The catalytic efficiency, E, of rhenium at high temperatures for the equilibration of a mixture of carbon monoxide isotopes (12C18O + 13C16O) is reduced by pre-adsorbed oxygen; E at 1300 K declines linearly to zero at an oxygen uptake of about 5 × 1014 atoms cm?2. The replacement of one pre-adsorbed carbon monoxide isotope by another can be correlated with the characteristic desorption temperatures of the two main states (α and β) of CO on Re. The observation that a considerable fraction of CO is non-replaceable at filament temperatures below 700 K suggests a high activation energy for migration of some adsorbed CO. The probability of exchange of 16O between an oxygenated rhenium filament and gaseous 12C18O for oxygen coverages ?4 × 1014 O atoms cm?2 is 0.012 per 1014 O atoms cm?2 per collision with the filament at 900 K. The surface reaction Re-16O + 12C18O(g) = Re-18O + 12C16O(g) is completely reversible. However, in the presence of nitrous oxide no reaction is observed until the filament temperature exceeds 1600 K, when continuous decomposition of N2O is appreciable. Possible transition states for isotope exchange are discussed.  相似文献   

5.
A Study of electronic conductivity using the d.c. polarization technique has been carried out in α and β-AgI which shows the former is a hole and the latter an electron conductor. Activation energies of undoped and Cu-doped single crystals and polycrystalline β-AgI were found to be 0.46 eV, 0.34 eV and 0.44 eV respectively and can be related to electron trap depths. The electron transference number (σθσt) for polycrystalline β-AgI was found to be 0.008 at 306 K. The activation energy for hole conduction in α-AgI was determined to be 0.97 eV in agreement with previous XPS studies.Transient measurements have also been conducted using the charge transfer technique in double cells of polycrystalline β-AgI. The carrier concentration Cθ and electron mobility μθ, have thus been estimated to be 1.8 × 1015cm3 and 5.14 × 10?5cm2V?sec. respectively at 306 K, while the double layer capacitance was 0.496 μFcm2.  相似文献   

6.
CO adsorption/desorption on clean and sulfur covered Pt(S)-[9(111) × (100)] surfaces was studied using AES, TPD, and modulated beam experiments. CO desorption occurred from two states on the clean surface — a low temperature state associated with the (111) terraces and a high temperature state associated with the steps/defects. Thermal desorption results indicated that above small CO coverages conversion from the low temperature state into the high temperature state was activated and that back conversion was slow. Sulfur preferentially adsorbed at step/defect sites and decreased the population of the high temperature desorption state. Modulated beam experiments were performed in order to determine CO adsorption/desorption parameters as a function of sulfur coverage on the Pt crystal. The sticking coefficient and binding energy of CO decreased as the sulfur concentration increased. Sulfur adsorption at step/defect sites decreased the CO sticking coefficient only slightly but increased the effective rate constant for CO desorption significantly. Sulfur adsorption on the terraces affected CO adosrption more than sulfur at step sites. On the clean surface the effective rate constant for CO desorption was
1 × 1015 s?1 exp (?36.2 kcal/moleRT)
Desorption occurred from both terrace and step/defect sites, but the kinetics were characteristic of the step/defect sites. For the surface on which step/defect sites were blocked by sulfur the effective desorption rate constant was
keff = 1 × 1013 s?1 exp (?27.5 kcal/moleRT)
indicating an appreciable decrease in CO binding on the terraces, though sulfur-CO repulsive interactions had probably made keff larger than the true rate constant for desorption from clean (111) planes. The results showed clearly a compensation effect in activation energy and preexponential factor.  相似文献   

7.
The kinetics of the desorption of CO from a Pt(111) crystal between 419 and 505 K is reported using a Low-Energy Molecular-Beam-Scattering (LEMS) technique with a helium probe beam and a CO dosing beam. The resulting first-order Arrhenius rate constant is k = 2.7 × 1013exp(?31.1 kcalmole · RT) s?1. We also report a study of the equilibriumadsorbed CO between 400 and 600 K using LEMS. These results, fitted to a Temkin isotherm model, indicate that the adsorption energy decreases linearly with surface coverage with the average value equal to 31.1 + 1.2 kcalmole over the coverage range 0 < θ ? 0.5. The average harmonic oscillator frequency of the adsorbed CO molecules is 191 ± 76 cm?1.  相似文献   

8.
Clean Pt(100) surfaces with bulk-like 1×1 structure, or the stable, reconstructed 5×20 structure and held at 200 or 330 K were exposed to ethylene. Ultraviolet photoemission spectroscopy identified the nature of the adsorbed species which depends on the structure and temperature of the clean surface and the amount adsorbed. It is ethylene on the 5 × 20 structure at 200 K, a vinyl radical on the same surface at 300 K up to half a monolayer, the remainder being added as acetylene; it is acetylene on the 1 × 1 surface at 330 K and a mixture of acetylene, vinyl and ethylene on the 1 × 1 surface at 200 K. Whatever the nature of the adsorbate, the surface coverage θ increased with exposure ? as (1 ? θ = C??13). By contrast, on a surface covered with any C2 hydrocarbon acetylene adsorbs with Langmuir kinetics. The kinetics are explained in terms of the relationship between the attraction an approaching molecule experiences from the bare surface and its Van der Waals repulsion from preadsorbed molecules.  相似文献   

9.
Angular distributions of six polarization transfer coefficients Kxx′(θ), kxz′(θ), Kzx?(θ), Kzz?(θ), and Kyyy?(θ); of the four analyzing powers Ay(θ), Axx(θ), Ayy(θ), and Azz(θ); and of the polarization function Pý(θ), have been measured atEd = 10.00 MeV for the reaction 2H(d, n)3He. Measurements were made for neutron lab angles between 0° and 80° in 10° steps. Additionally the y-axis associated quantities were measured at θ1ab = 99°. Most of the measured coefficients are large at some angles and all show considerable variation with angle.  相似文献   

10.
The adsorption of K on Pt(100) has been followed by thermal desorption spectroscopy (TDS) and Auger electron spectroscopy (AES); carbon monoxide was used as a probe for the modification of the chemical properties of K promoted surfaces. The role of subsequent adsorption of oxygen on the K modified surfaces has also been measured. For low potassium coverage (θK = 0 to 0.35), the mass-28 TDS peak temperature of adsorbed CO increases continuously with the K coverage, indicating an increase of the adsorption energy of CO which has been explained by a substantial charge donation from K into the 1 orbitals of CO via long range interactions through the platinum substrate. No oxygen uptake was detected after oxygen exposure at room temperature. For high potassium content (θK = 0.45 to 1), the mass-28 TDS peak temperature of coadsorbed CO is very narrow and remains constant at 680 K. We propose the formation of a COKPt surface complex which decomposes at 680 K, since K desorption is detected concomitantly to CO. On such K covered surfaces, the oxygen uptake is promoted, and it cancels the modifications of the surface properties induced by potassium.  相似文献   

11.
The quantum efficiency of fluorescence, Y, of the 4607.33 Å Sr line (1P1?1S0transition) was measured in four pre-mixed, laminar, shielded CO/N2O flames of about 2700 K, with different quantitative compositions at 1 atm. From these data, the specific quenching cross sections for O2, CO2, CO and N2 were found to be (152±20 Å2), (30±5 Å2), (49±8 Å2) and (16±3 Å2), respectively.  相似文献   

12.
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   

13.
14.
The catalytic decomposition of formic acid by a polycrystalline platinum surface was studied by use of modulated molecular beam techniques with mass spectrometric phasesensitive detection. Kinetic information about elementary surface reaction steps was obtained. The formation of CO2 was found to be a monomolecular, whereas that of D2 was a bimolecular process. The resulting reaction mechanism may be described as follows:
The rate constants in dependence from the surface temperature t0 are η = 7.1 × 103exp(?9.9RT0kcal/mole),kd1 = 6.3 × 107exp(?10.2RT0kcal/mole) (sec?1),kd2 = 3.2 × 10?8exp(?9.3RT0kcal/mole) (particles?1sec?1cm2). The sticking probability η is provided by the temperature dependence of the intensity of the nonreactive scattered formic acid molecules; the rate constants kd1 and kd2 are derived from the measured phase shift between reactive and nonreactive scattered particles. From the phase angle ?, the average surface residence time τ of the intermediates is computed: 3.7 ? τDCOO ? 0.41 msec (418 ? T0 ? 505 K), 31.8 ? τD ? 11.6 msec (418 ? T0 ? 460 K). The difference between τD and τDCOO is because of the different molecularity of desorption.  相似文献   

15.
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   

16.
Electron spin resonance relaxation times were measured for the radiation induced radical ion SeO43? in selenium doped KDP single crystals. The spin-lattice relaxation time was found to obey the relation T1R?1 = AT + BT50θ2T x4csch2x dx from 7 K to 200 K except in the neighborhood of the transition temperature where the data fit the expression T1?1 = T1R?1b±T ? Tcm± where θ is the Debye temperature and the plus and minus signs refer to data at temperatures above and below Tc respectively.  相似文献   

17.
Measurements of the molar magnetic susceptibility (Xm) of a powdered sample of Nd2(WO4)3 in the temperature range 300–900 K, and the electrical conductivity (σ) and dielectric constant (?)? of pressed pellets of the compound in the temperature range 4.2–1180 K are reported. Xm obeys the Curie-Weiss law with a Curie constant C= 3.13 K/mole, a paramagnetic Curie temperature θ= ?60 K and a moment of Bohr magnetons, p= 3.49 for the Nd3+ ion. The electrical conductivity data can be explained in terms of the usual band model and impurity levels. Both the σ and ?$?data indicate some sort of phase transition round 1025 K. The conductivity follows Mott's law σ = A exp (?B/T14) in the temperature range 200 < T < 3000 K with B = 45.00 (K)14and A = 1.38 × 10?5 Ω?1cm?1. The dielectric constant increases slowly up to 600 K, as is usual for ionic solids. The increase becomes much faster above 600 K, which is attributed to space-charge polarization of thermally generated charge carriers.  相似文献   

18.
The chemisorption of CO on Cu, Ni and CuNi alloy surfaces was examined by SIMS, work function measurements and desorption spectroscopy. Using a dynamic SIMS technique the M+, M+2, MCO+ and M2CO+ emission at different temperatures (100–400 K) was measured as a function of CO exposure. In agreement with the work function and desorption experiments an increase of M+ and MCO+ emission due to the CO adsorption on Cu was found only at low temperatures (100–190 K). On the Ni surface an increase of Ni+, NiCO+ and Ni2CO+ was measured up to 400 K. The adsorption of CO on CuNi alloy surfaces — as derived from the work function measurements — can be described by the assumption of two different states of adsorbed carbon monoxide. They can be characterized by different binding energies and from sign and magnitude different work function changes. These states were interpreted as adsorption at Ni or Cu sites of the alloy surfaces, respectively. To a certain extent the SIMS results from the alloy surfaces are incompatible with the work function measurements and desorption spectroscopy and the SIMS studies on the pure metals. A Cu+ emission with comparable intensity to the Ni+ emission was found for alloys with bulk concentrations of 60 and 40 at% Cu at 300 K. The ratio Ni+Cu+ was nearly independent of CO pressure and temperature. The measured ratios of Cu+2(Cu+ + Ni+), Ni+2(Cu+ + Ni+) and CuNi+(Cu+ + Ni+) with values about 10?2 can be explained the basis of a statistical arrangement of Cu and Ni atoms in the alloy surface. The intensities of the MCO+ emissions are 102 times smaller than the corresponding values of the pure metals. No emission of M2CO+ was found on CuNi during CO adsorption.  相似文献   

19.
The adsorption of 12CO on Ir films evaporated under ultrahigh vacuum (UHV) conditions was studied using infrared reflection-absorption spectroscopy (IRAS). Only a single absorption band was observed at 300 K, shifting continuously from the “singleton” value ~2010 cm?1 at very low coverages to 2093 cm?1 at saturation coverage. This band is attributed to CO adsorbed on top of the surface atoms. Synchronously with this shift the bandwidth at half maximum intensity Δv12 decreases from ~30 to 8 cm?1. The integrated peak area increases linearly with coverage up to a relative coverage (θr) of approximately 0.4, then the increase levels off and a maximum is observed. Upon continuing adsorption the intensity decreases slightly. In addition results are presented on adsorption at 300 K of 12CO?13CO isotopic mixtures. The coverage induced frequency shift is discussed in terms of a dipole-dipole coupling mechanism and it is concluded that intermolecular coupling can explain the shift (~83 cm?1) observed. The decrease in intensity at coverages > 0.4 is attributed to the formation of a compressed overlayer with part of the CO molecules adsorbed in a multicentre position with different spectral properties. No infrared bands of nitrogen adsorbed at 78 K could be detected at pressures up to 6.7 kPa (1 Pa = 0.0075 Torr, 1 Torr = 133.32 Pa).  相似文献   

20.
In lattice gauge theory, many computations such as the strong coupling expansions, mean field theory, or the few plaquette models require the evaluation of the one-link integral in the presence of an arbitrary N × N complex matrix source (J). For SU(N) gauge theories, we express our general solution to the external field problem as an integral over the maximal abelian subgroup [U(1)]N?1
dUe?r(J+U+U+J)=∫i=1Ndøp(Σøi)eS0+InG
where S0 = 2Σkzk cos(φk ? θ), zj are eigenvalues of √JJ+, e2iNθ=detJ/detJ+, and G is an appropriate jacobian determinant. Our explicit solution follows from differential Schwinger-Dyson equations cast in a separable form by using fermionic variables, and the special cases of N = 2, 3 and ∞ agree with earlier derivations.  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