scispace - formally typeset
Search or ask a question

Showing papers on "Chemisorption published in 1972"


Journal ArticleDOI
TL;DR: In this article, the chemisorption of hydrogen, oxygen, carbon, carbon monoxide and ethylene was studied by low-energy electron diffraction on ordered stepped surfaces of platinum which were cut at angles less than 10° from the (111) face.

309 citations


Journal ArticleDOI
TL;DR: In this article, a flow method was used to measure hydrogen chemisorption on supported platinum catalysts at ambient temperature, the data were in accord with results obtained by the conventional static method.

172 citations


Journal ArticleDOI
TL;DR: In this paper, the potential of the oxygen desorption peak on a voltammogram was found to vary linearly with surface composition of homogeneous platinum-rhodium, palladium rhodium and palladium-gold alloy surfaces.

157 citations


Journal ArticleDOI
TL;DR: In this paper, the interaction of O 2 with a single crystal W(100) surface at 300K has been examined using a combination of techniques, where a precision gas dosing system was used to deposit O 2 onto the crystal; the absolute accuracy in the average flux F (molecules cm −2 s −1 ) is ± 6 %.

155 citations


Journal ArticleDOI
TL;DR: In this paper, an Extended Huckel MO (EHMO) method was used to calculate the adsorption energy of a hydrogen atom over a surface nickel atom, which was shown to be more favorable than adsorbing in some surface holes.

114 citations


Journal ArticleDOI
TL;DR: In this paper, the chemisorption of hydrogen, oxygen and carbon monoxide on supported platinum was studied by a pulse technique, and the resulting adsorption stoichiometries were compared with literature data which either include measurements of size or which compare the adsorptions of two or more gases on the same solid.

110 citations


Journal ArticleDOI
TL;DR: In an extension of previous work on Ni(100), the method of ion-neutralization spectroscopy has been used to determine orbital energy spectra of electrons in the chemisorption bonds holding O, S, or Se in ordered arrays on the Ni(110) and Ni(111) faces.

109 citations


Journal ArticleDOI
TL;DR: In this article, surface potentials, diffraction patterns, and vibrational frequencies have been compared in relation to the chemisorption bond and the relationship between the surface potential and the vibrational frequency has been discussed.
Abstract: Although it is much less active than the transition metals copper does catalyse reactions of hydrogen and carbon monoxide. The nature of the weak adsorptions of these gases has been investigated by surface potentials, low energy electron diffraction, and infrared spectroscopy. Hydrogen is dissociatively adsorbed by copper but the adsorption is activated. Surface potential measurements have provided isotherms and isosteric heats of adsorption. Carbon monoxide is adsorbed in two distinct stages giving positive and negative surface potential changes. The positive stage is accompanied by a sharp infrared absorption band near 2100 cm−1 which is sufficiently strong to permit reflection spectroscopic studies on single crystal surfaces. Frequency shifts occur during the negative stage. Direct comparisons of surface potentials, diffraction patterns, and vibrational frequencies have been made and will be discussed in relation to the chemisorption bond.

104 citations


Journal ArticleDOI
TL;DR: In this article, an oxygen uptake model based on an incorporation-chemisorption transition is proposed to explain the "stable layer" associated with the aluminum oxygen reaction at room temperature.

104 citations


Journal ArticleDOI
TL;DR: In this paper, the adsorption energy, Arrhenius preexponential factor and sticking coefficient have been determined for surface hydrogen populations less than 1013 atoms/cm2, the values obtained being 22.7 kcal/mole, 2×10−1 cm2 · sec−1, and 0.1, respectively.
Abstract: The adsorption of hydrogen on the (111) plane of nickel at room temperature has been investigated using the technique of flash desorption spectroscopy. The residual CO partial pressure has been reduced below 1 × 10−11 torr to minimize the effect of H2 and CO coadsorption. The experimental apparatus and crystal cleaning procedure are described. Hydrogen was observed to desorb in a single pressure burst for all the experimental conditions examined. Four techniques have been used to analyze the experimental results and a coherent picture emerges for low surface populations. The adsorption energy, Arrhenius preexponential factor and sticking coefficient have been determined for surface hydrogen populations less than 1013 atoms/cm2, the values obtained being 22.7 kcal/mole, 2×10−1 cm2 atoms−1 · sec−1, and 0.1, respectively. Furthermore it is shown that the adsorption energy is constant at low coverages. For surface populations greater than 1013 atoms/cm2, we have been able to reproduce the desorption spectra with a model in which either the adsorption energy decreased with adsorbed population or the pre‐exponential factor increased with population. It is not possible to distinguish between the two possibilities. Such a model does not however describe very well the observed adsorption kinetics. We have proposed that this discrepancy may be due to diffusion of hydrogen into the nickel when the crystal is heated.

94 citations


Journal ArticleDOI
TL;DR: In this article, the adsorption of carbon monoxide on epitaxial (100) and (111) planes of Ag/Pd alloys with definite surface compositions has been studied by means of LEED, Auger electron spectroscopy and work function measurements.

Journal ArticleDOI
TL;DR: In this paper, the Pt particle size determinations were made from electron micrographs and from X-ray line broadening experiments following the chemisorption measurements on samples sintered in H 2 to various temperatures.

Journal ArticleDOI
TL;DR: The surface potentials (s.p.) of atomic hydrogen adsorbed at low temperature on clean copper films show wide variations from one film to another as discussed by the authors, which is believed to reflect varying proportions of different crystal planes.
Abstract: Surface potentials (s.p.) of atomic hydrogen adsorbed at low temperature on clean copper films show wide variations from one film to another. This is believed to reflect varying proportions of different crystal planes. Initial adsorption with a negative s.p. may be followed by a weaker adsorption with a positive effect. The negative s.p. corresponds to hydrogen which can be in equilibrium with gaseous molecular hydrogen. This spontaneous dissociative chemisorption has been studied between 242 and 337 K at pressures up to 60 Torr, using vibrating capacitor s.p. measurements to follow the equilibrium concentration of adatoms. Isosteric heats (typically 40–50 kJ mol–1) are almost independent of coverage on some films, and a major part of the adsorption can then be described approximately by the Langmuir isotherm for dissociative adsorption.

Journal ArticleDOI
TL;DR: In this paper, the authors used the flash desorption technique to study adsorption of hydrogen on very clean, reproducible surfaces of polycrystalline iron and showed that the initial rates of adsorment of hydrogen are consistent with an activated complex that requires about 0.5 kcal mole−1 for its formation, and this energy remains nearly constant up to a surface coverage of 0.2 × 1015 H-atoms cm−2.

Journal ArticleDOI
TL;DR: In this paper, the thermal desorption method was used in conjunction with mass spectrometry and work function changes to examine the adaption and displacement processes involving CH4, H2, and O2 on the surface of tungsten.

Journal ArticleDOI
TL;DR: In this paper, the chemisorption of CO on a tungsten (100) single crystal surface has been studied at adsorption temperatures of 295 K and 195 K. The role of surface species' structure in determining ionic fragmentation probabilities is discussed.

Journal ArticleDOI
TL;DR: In this article, the interaction of water vapor with clean bulk gold surfaces was studied by means of measurement of work function changes, and a work function decrease of as much as one volt occurred on exposure of gold to water vapor.

Journal ArticleDOI
TL;DR: In this paper, a nomenclature for surface defects and adsorbed species was proposed which is compatible with a recently suggested notation for defects in bulk lattice, and the optical transition energy calculated from the e.s.r. data with the observed transition was analyzed.
Abstract: Nitrous oxide reacts with Fs+(H) centres (surface oxygen ion vacancies with a single trapped electron) to give (O–)s+ ions adsorbed at the anion vacancy. The reaction, Fs+(H)+ N2O →(O–)s++ N2 is quantitative and the reflectance spectrum indicates a band centred at 2 eV which is ascribed to the (O–)s+ centre. Comparison of the optical transition energy calculated from the e.s.r. data with the observed transition shows that the normally accepted value of 135 cm–1 for the spin-orbit coupling of the O– ion is too low and that a value of >200 cm–1 is more appropriate for the solid state. Nitrous oxide also reacts with other surface centres to give an adsorbed species which may be (O2–2)s or (O–2)s. The reaction of (O–)s+ with hydrogen reforms the Fs+(H) centre which then reacts with more nitrous oxide; the reaction can be cycled several times. The reactions of (O–)s+ with O2, CO2, alcohols and olefins are also investigated. To facilitate reference to various types of surface defects and adsorbed species a nomenclature is proposed which is compatible with a recently suggested notation for defects in bulk lattice.

Journal ArticleDOI
TL;DR: In this paper, a method for the determination of palladium surface areas using hydrogen chemisorption is described, which is independent of the extent to which oxygen is preadsorbed on the metal surface.

Journal ArticleDOI
TL;DR: In this article, the effect of exposure to oxygen at pressures up to 1 Torr and temperatures up to 500 C was investigated using a crystal isolation valve, and the chemisorbed oxygen caused mainly a strong faceting of the (100) and (110) faces, and a 4×4 superstructure on the (111) face.
Abstract: The chemisorption of oxygen on the (100), (110), and (111) faces of silver single crystals was studied with LEED, secondary electron spectroscopy, and other auxiliary techniques. The effect of exposures to oxygen at pressures up to 1 Torr and temperatures up to 500 °C was investigated using a crystal isolation valve. The chemisorbed oxygen causes mainly a strong faceting of the (100) and (110) faces, and a 4×4 superstructure on the (111) face. The effect of the internally diffused oxygen on the surface behavior was also investigated.

Journal ArticleDOI
TL;DR: In this paper, the vibrational frequency of carbon monoxide chemisorbed on an atomically clean polycrystalline tungsten ribbon has been studied using reflectance-infrared spectroscopy and ultrahigh vacuum techniques.

Journal ArticleDOI
TL;DR: In this article, a comparison of measured conductivities with those computed from different surface state models (discrete or distributed), chosen a priori, reveals that the oxygen chemisorption states of natural surfaces of CdS (single crystals and evaporated films) are distributed in energy.

Journal ArticleDOI
TL;DR: In this paper, high Miller index crystal surfaces of platinum have been shown to consist of low index (111) or (100) terraces of constant width, linked by steps of monatomic height.
Abstract: Several high Miller index crystal surfaces of platinum have been shown to consist of low index (111) or (100) terraces of constant width, linked by steps of monatomic height. The surface structures that form in the presence of diatomic molecules (H 2 , O 2 , CO, NO), aliphatic and aromatic hydrocarbons on the (111), (100) and on these stepped platinum surfaces were studied by low energy diffraction. Several catalytic reactions (H 2 + D 2 ; H 2 + O 2 ; dehydrocyclization of n -heptane) that take place on the various platinum crystal surfaces at low pressures were monitored by means of a mass spectrometer and the surface composition by Auger electron spectroscopy. The chemisorption characteristics of diatomic molecules on stepped platinum surfaces are markedly different from those on low index [(111) and (100)] platinum surfaces. Organic molecules of different types form ordered surface structures on low index faces of platinum, but decompose rapidly on stepped surfaces under identical experimental conditions. Surface chemical reactions of diatomic molecules that were studied take place only on stepped surfaces at a detectable rate. The dehydrocyclization of n -heptane to toluene occurs slowly on the Pt(lll) face. The rate of decomposition of n -heptane on stepped platinum surfaces is rapid and the carbon deposit that forms prevents dehydrocyclization. In the presence of hydrogen, however, dehydrocyclization occurs at a rapid rate on stepped surfaces with atomic terraces of (111) orientation, while at a slower rate on stepped surfaces with atomic terraces of (100) orientation. It appears that catalytic reactions can readily be studied using one face of a single crystal of small surface area. Dissociation of diatomic molecules and breaking of C—C and C—H bonds occur preferentially at atomic steps, while the structure of the atomic terraces plays an important role in the more complex dehydrocyclization reaction.

Journal ArticleDOI
TL;DR: In this article, the size distribution of platinum particles is controlled by the calcination temperature of the zeolite 13Y, loaded with platinum cations by ion exchange, was calcined in the air and subsequently reduced in a hydrogen stream.
Abstract: Zeolite 13Y, loaded with platinum cations by ion exchange, was calcined in the air and subsequently reduced in a hydrogen stream. Hydrogen and carbon monoxide adsorption studies, together with an electron microscopic observation of the catalyst, revealed that the size distribution of platinum particles is controlled by the calcination temperature. The hydrogen chemisorption method gave the average particle diameters, which were in fairly good agreement with the observations made by means of an electron microscope on the catalyst with a platinum particle diameter larger than 20 A, while trace amounts of hydrogen chemisorption were observed for platinum particles less than 20 A.

Journal ArticleDOI
TL;DR: In this paper, the exact sticking probability of oxygen adsorption on a clean platinum (111) surface was derived using the absolute rate theory of gas phase kinetics together with the Crystal Field Surface Orbital-Bond Energy Bond Order (CFSO-BEBO) model.

Journal ArticleDOI
TL;DR: In this paper, the authors measured the isotherms and rates for NO and CO chemisorption on supported and unsupported nickel oxide samples in the 0-140 °C temperature range.

Journal ArticleDOI
TL;DR: In this paper, the dependence of the isosteric heat of adsorption on the coverage σ is given by E = 16.3 − 1.4 × 10 − 14 σ (kcal/mole) with σ 14 molecules/cm 2.

Journal ArticleDOI
TL;DR: In this article, the adsorption of boron trifluoride on silica has been studied using infrared-red spectroscopy and it is shown that BF3 reacts with a new reactive site, probably a siloxane bridge site, which is formed as a result of the removal of the surface silanol groups during heating.
Abstract: Infra-red spectroscopy has been used to study the adsorption of boron trifluoride on silica. On highly dehydroxylated samples chemisorption is rapid and the major product initially appears to be a surface SiOBF2 species. This species is also formed when BF3 interacts with totally dehydroxylated silicas which have been prepared by heating silica in vacuum at about 1250°C. Spectroscopic evidence indicates that on highly or totally dehydroxylated silica, BF3 preferably reacts with a new reactive site, probably a siloxane bridge site, which is formed as a result of the removal of the surface silanol groups during heating. Only after the reaction is complete on the “siloxane” sites does chemisorption take place with silanol groups if these are present. During prolonged evacuation the surface SiOBF2 groups react further with trace amounts of water to produce (SiO)2BF groups; the latter are also produced during the initial stages of chemisorption when the silanol concentration is high. When water or air is admitted after chemisorption of BF3, a spectrum characteristic of boric acid is observed.

Journal ArticleDOI
TL;DR: In this paper, a new cutting procedure inside an ultra-high vacuum system was used to stabilize the reference electrode of a vibrating capacitor by equilibration with the gas, which led to the conclusion that the water chemisorbs by a precursor mechanism.

Journal ArticleDOI
TL;DR: In this article, Anderson's Hamiltonian, which describes chemisorption in terms of the formation of a virtual level and restricts electron interactions (Coulomb plus exchange) to the adatom, is used as the starting point.