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.
J.M.R. Muir  H. Idriss 《Surface science》2009,603(19):2986-2990
The reaction of formamide over the (0 1 1) faceted TiO2(0 0 1) surface has been studied by Temperature Programmed Desorption (TPD) and X-ray Photoelectron Spectroscopy (XPS). Two main reactions were observed: dehydration to HCN and H2O and decomposition to NH3 and CO. The dehydration reaction was found to be three to four times larger than the decomposition at all coverages. Each of these reactions is found to occur in two temperature domains which are dependent upon surface coverage. The low temperature pathway (at about 400 K) is largely insensitive to surface coverage while the high temperature pathway (at about 500 K) shifts to lower temperatures with increasing surface coverage. These two temperature pathways may indicate two adsorption modes of formamide: molecular (via an η1(O) mode of adsorption) and dissociative (via an η2(O,N) mode of adsorption). C1s and N1s XPS scans indicated the presence of multiple species after formamide absorption at 300 K. These occurred at ca. 288.5 eV (-CONH-) and 285 eV (sp3/sp2 C) for the C1s and 400 eV-(NH2), 398 eV (-NH) and 396 eV (N) for the N1s and result from further reaction of formamide with the surface.  相似文献   

2.
Yuhai Hu 《Surface science》2007,601(21):5002-5009
The influence of pre-dosed O2 on the catalytic reduction of NO with 13C2H5OH on the surface of stepped Pt(3 3 2) was investigated using Fourier transform infra red reflection-absorption spectroscopy (FTIR-RAS) and thermal desorption spectroscopy (TDS). We show that the oxidation of 13C2H5OH with O2 is a very effective reaction, occurring at 150 K and giving rise to acetate. The presence of NO does not lead to any evident oxidation of 13C2H5OH irrespective of the annealing temperature. For the case of O2 + 13C2H5OH + NO co-adlayers, oxidation of 13C2H5OH also takes place at 150 K. However, no new surface species that are supposed to be an intermediate for the production of N2 are detected.The influence of O2 on the production and desorption of N2 is intimately related to both O2 and 13C2H5OH coverage. The presence of pre-dosed O2 does not greatly promote N2 desorption. In fact, N2 desorption is suppressed quantitatively with increasing O2 coverage, after which unreacted, or left-over O atoms appear and remain on steps. It is concluded that the presence of pre-dosed O2 does not play a role of activating reactants in the catalytic reduction of NO with 13C2H5OH on the surface of Pt(3 3 2).  相似文献   

3.
By means of density functional theory calculations we have investigated the role of adsorbed atomic oxygen and adsorbed OH in the oxidation of ammonia on Pt{1 1 1}. We have investigated the dissociation of NH3,ads, NH2,ads and NHads on Pt{1 1 1} and the oxidation of these species by Oads and OHads. We have done normal mode frequency analysis and work function calculations to characterise reactant, product and transition states. We have determined reaction energies, activation entropies, kinetic parameters and corrected total energies with the zero point energy. We have shown that Oads only activates the dehydrogenation of NH3,ads and that OHads activates the dehydrogenation of all NHx,ads species and have reasoned this difference in activation by a bond order conservation principle. We have pointed out the importance of a zero point energy correction to the reaction energies and barriers. We have compared the calculated vibrational modes of the adsorbates with corresponding experimental EELS data. This has led to a revise of the frequency assignment of ν(Pt-OH2), a revise in the identification of a NH2 species on the Pt{1 1 1} surface after electron bombardment of pre-adsorbed NH3 and the confirmation of an ammonia dimer binding model at the expense of a hollow site occupation by ammonia on the Pt{1 1 1} surface.  相似文献   

4.
The oxidation of the Pd(1 1 1) surface was studied by in situ XPS during heating and cooling in 3 × 10−3 mbar O2. A number of adsorbed/dissolved oxygen species were identified by in situ XPS, such as the two dimensional surface oxide (Pd5O4), the supersaturated Oads layer, dissolved oxygen and the R 12.2° surface structure.Exposure of the Pd(1 1 1) single crystal to 3 × 10−3 mbar O2 at 425 K led to formation of the 2D oxide phase, which was in equilibrium with a supersaturated Oads layer. The supersaturated Oads layer was characterized by the O 1s core level peak at 530.37 eV. The 2D oxide, Pd5O4, was characterized by two O 1s components at 528.92 eV and 529.52 eV and by two oxygen-induced Pd 3d5/2 components at 335.5 eV and 336.24 eV. During heating in 3 × 10−3 mbar O2 the supersaturated Oads layer disappeared whereas the fraction of the surface covered with the 2D oxide grew. The surface was completely covered with the 2D oxide between 600 K and 655 K. Depth profiling by photon energy variation confirmed the surface nature of the 2D oxide. The 2D oxide decomposed completely above 717 K. Diffusion of oxygen in the palladium bulk occurred at these temperatures. A substantial oxygen signal assigned to the dissolved species was detected even at 923 K. The dissolved oxygen was characterised by the O 1s core level peak at 528.98 eV. The “bulk” nature of the dissolved oxygen species was verified by depth profiling.During cooling in 3 × 10−3 mbar O2, the oxidised Pd2+ species appeared at 788 K whereas the 2D oxide decomposed at 717 K during heating. The surface oxidised states exhibited an inverse hysteresis. The oxidised palladium state observed during cooling was assigned to a new oxide phase, probably the R 12.2° structure.  相似文献   

5.
Eldad Herceg 《Surface science》2006,600(19):4563-4571
The formation of a well-ordered p(2 × 2) overlayer of atomic nitrogen on the Pt(1 1 1) surface and its reaction with hydrogen were characterized with reflection absorption infrared spectroscopy (RAIRS), temperature programmed desorption (TPD), low energy electron diffraction (LEED), Auger electron spectroscopy (AES), and X-ray photoelectron spectroscopy (XPS). The p(2 × 2)-N overlayer is formed by exposure of ammonia to a surface at 85 K that is covered with 0.44 monolayer (ML) of molecular oxygen and then heating to 400 K. The reaction between ammonia and oxygen produces water, which desorbs below 400 K. The only desorption product observed above 400 K is molecular nitrogen, which has a peak desorption temperature of 453 K. The absence of oxygen after the 400 K anneal is confirmed with AES. Although atomic nitrogen can also be produced on the surface through the reaction of ammonia with an atomic, rather than molecular, oxygen overlayer at a saturation coverage of 0.25 ML, the yield of surface nitrogen is significantly less, as indicated by the N2 TPD peak area. Atomic nitrogen readily reacts with hydrogen to produce the NH species, which is characterized with RAIRS by an intense and narrow (FWHM ∼ 4 cm−1) peak at 3322 cm−1. The areas of the H2 TPD peak associated with NH dissociation and the XPS N 1s peak associated with the NH species indicate that not all of the surface N atoms can be converted to NH by the methods used here.  相似文献   

6.
In H2 and H2/CO oxidation, the H + O2 + M termination step is one of the most important reactions at elevated pressures. With the recent, increased interest in synthetic fuels, an accurate assessment of its rate coefficient becomes increasingly important, especially for real fuel/air mixtures. Ignition delay times in shock-tube experiments at the conditions selected in this study are only sensitive to the rates of the title reaction and the branching reaction H + O2 = OH + O, the rate of which is known to a high level of accuracy. The rate coefficient of the title reaction for M = N2, Ar, and H2O was determined by adjusting its value in a detailed chemical kinetics model to match ignition delay times for H2/CO/O2/N2, H2/CO/O2/Ar, and H2/CO/O2/N2/H2O mixtures with fuel/air equivalence ratios of ? = 0.5, 0.9, and 1.0. The rate of H + O2 + N2 = HO2 + N2 was measured to be 2.7 (−0.7/+0.8) × 1015 cm6/mol2 s for T = 916-1265 K and P = 1-17 atm. The present determination agrees well with the recent study of Bates et al. [R.W. Bates, D.M. Golden, R.K. Hanson, C.T. Bowman, Phys. Chem. Chem. Phys. 3 (2001) 2337-2342], whose rate expressions are suggested herein for modeling the falloff regime. The rate of H + O2 + Ar = HO2 + Ar was measured to be 1.9 × 1015 cm6/mol2 s for T = 932-965 K and P = 1.4 atm. The rate of H + O2 + H2O = HO2 + H2O was measured to be 3.3 × 1016 cm6/mol2 s for T = 1071-1161 K and P = 1.3 atm. These are the first experimental measurements of the rates of the title reactions in practical combustion fuel/air mixtures.  相似文献   

7.
A.P. Farkas  F. Solymosi 《Surface science》2006,600(11):2355-2363
The adsorption and surface reactions of propyl iodide on clean and potassium-modified Mo2C/Mo(1 0 0) surfaces have been investigated by thermal desorption spectroscopy (TPD), X-ray photoelectron spectroscopy (XPS) and high resolution electron energy loss spectroscopy (HREELS) in the 100-1200 K temperature range. This work is strongly related to the better understanding of the catalytic effect of Mo2C in the conversion of hydrocarbons. Potassium was found to be an effective promoter: it induced the rupture of C-I bond in the adsorbed C3H7I even at 100 K. The extent of C-I bond scission varied approximately linearly with the concentration of K coverage at the adsorption temperature of 100 K. As revealed by HREELS and TPD measurements the primary products of the dissociation are C3H7 and I. The former one was stabilized by potassium and underwent dehydrogenation and hydrogenation to give propene and propane. The desorption of both compounds is reaction-limited process. A fraction of propyl groups was converted into di-σ-bonded propene, which was stable up to ∼380 K. The coupling reaction of propyl species was also facilitated by potassium and resulted in the formation of hexane and hexene with Tp ∼ 230-250 K. Hydrogen was released with Tp = 390 K, indicative of a desorption limited process. The effect of potassium was explained by the extended electron donation to adsorbed propyl iodide in one hand, and by the direct interaction between potassium and I on the other hand. This was reflected by the shift of the desorption of potassium from the coadsorbed layer at and above 1.0 ML to higher temperature, and by the coincidal Tp values (∼700 K) of potassium and iodine. The formation of KI was also supported by the appearance of a loss feature at 650 cm−1 in the HREEL spectra attributed to a phonon mode of KI.  相似文献   

8.
The effect of coadsorbed oxygen on the thermal chemistry of diiodomethane on Ni(1 1 0) single-crystal surfaces was studied by temperature-programmed desorption (TPD) and X-ray photoelectron spectroscopy (XPS). I 3d and C 1s XPS data indicated that adsorbed diiodomethane undergoes two sequential C-I bond scission steps to ultimately produce methylene surface species, the same as on clean Ni(1 1 0). Moreover, significant amounts of methane and other heavier hydrocarbons are produced after further thermal activation of those chemisorbed methylene groups. The production of alkanes and alkenes, which is accounted for by a chain growth mechanism where the initial hydrogenation of some adsorbed methylene to methyl moieties is followed by a rate-limiting methylene insertion step to yield ethyl intermediates, is inhibited but not fully blocked by the coadsorbed oxygen. New reaction pathways are also opened up by the presence of oxygen in this system, including a direct coupling of two methylene groups to ethene, the insertion of an oxygen atom into a nickel-methylene group to produce formaldehyde, and a parallel methylene insertion chain growth sequence starting from a CH2Iads intermediate to ultimately yield C3H5 and C4H7 unsaturated gas-phase radicals.  相似文献   

9.
The oxidation of the Pd(1 1 1) surface was studied by in situ XPS during heating and cooling in 0.4 mbar O2. The in situ XPS data were complemented by ex situ TPD results. A number of oxygen species and oxidation states of palladium were observed in situ and ex situ. At 430 K, the Pd(1 1 1) surface was covered by a 2D oxide and by a supersaturated Oads layer. The supersaturated Oads layer transforms into the Pd5O4 phase upon heating and disappears completely at approximately 470 K. Simultaneously, small clusters of PdO, PdO seeds, are formed. Above 655 K, the bulk PdO phase appears and this phase decomposes completely at 815 K. Decomposition of the bulk oxide is followed by oxygen dissolution in the near-surface region and in the bulk. The oxygen species dissolved in the bulk is more favoured at high temperatures because oxygen cannot accumulate in the near-surface region and diffusion shifts the equilibrium towards the bulk species. The saturation of the bulk “reservoir” with oxygen leads to increasing the uptake of the near-surface region species. Surprisingly, the bulk PdO phase does not form during cooling in 0.4 mbar O2, but the Pd5O4 phase appears below 745 K. This is proposed to be due to a kinetic limitation of PdO formation because at high temperature the rate of PdO seed formation is compatible with the rate of decomposition.  相似文献   

10.
Infrared reflection absorption spectroscopy together with mass spectrometry has been used to investigate the interaction of NO and CO on Pt{1 0 0}, initially prepared in the reconstructed ‘hex’ phase, under ambient pressures of these gases, in the temperature range 300-500 K. The results allow the local and total coverages of adsorbed CO and NO to be related to the rate of reaction to produce gas phase CO2, and provide insight into the species present on the surface during the so-called low temperature oscillatory reaction regime of this process. At temperatures below that at which NO dissociation occurs (approximately 390-400 K) adsorption is controlled by the non-reactive displacement of NO by CO and results in a CO-poisoned surface. Above 400 K when significant CO2 production occurs, the NO coverage increases to produce a surface with NO and CO fully intermixed; the increase in NO coverage is attributed to the higher rate of NO arrival from the gas phase (with a partial pressure ratio of PNO:PCO>1) at free surface sites created by NO dissociation and subsequent reaction with CO. The competition between these two processes of non-reactive NO displacement by CO and reactive displacement of CO by NO is proposed to determine the parameter space of the low temperature oscillatory regime. Rapid equilibration between bridged and atop CO species leads to them appearing to exhibit identical reaction behaviour. Particularly at the lowest reaction temperatures (around 400 K), islands of pure CO may coexist on the surface but not participate in the reaction. Under conditions corresponding to the high temperature oscillatory regime, small quantities of absorbed CO, but no NO, are seen on the surface.  相似文献   

11.
The catalytic reduction of NO in the presence of benzene on the surface of Pt(3 3 2) has been studied using Fourier transform infra red reflection-absorption spectroscopy (FTIR-RAS) and thermal desorption spectroscopy (TDS). IR spectra show that while the presence of benzene molecules at low coverage (e.g., following an exposure of just 0.25 L) promotes NO-Pt interaction, the adsorption of NO on Pt(3 3 2) at higher benzene coverages is suppressed. It is also shown that there are no strong interactions between the adsorbed NO molecules and the benzene itself or benzene-derived hydrocarbons, which can lead to the formation of intermediate species that are essential for N2 production.TDS results show that the adsorbed benzene molecules undergo dehydrogenation accompanied by hydrogen desorption starting at 300 K and achieving a maximum at 394 K. Subsequent dehydrogenation of the benzene-derived hydrocarbons then begins with hydrogen desorption starting at 500 K. N2 desorption from NO adlayers on clean Pt(3 3 2) surface becomes significant at temperatures higher than 400 K, giving rise to a peak at 465 K. This peak corresponds to N2 desorption from NO dissociation on step sites. The presence of benzene promotes N2 desorption, depending on the benzene coverage. When the benzene exposure is 0.25 L, the N2 desorption peak at 459 K is dramatically increased. Increasing benzene coverage also results in the intensification of N2 desorption at ∼410 K. At benzene exposures of 2.4 L, N2 desorption develops as a broad peak with a maximum at ∼439 K.It is concluded that the catalytic reduction of NO by platinum in the presence of benzene proceeds by NO decomposition and subsequent oxygen removal at temperatures lower than 500 K, and NO dissociation is a rate-limiting step. The contribution of benzene to N2 desorption is mainly attributed to providing a source of H, which quickly reacts with NO-derived atomic O, leaving the surface with more vacant sites for further NO dissociation.  相似文献   

12.
Xueing Zhao 《Surface science》2007,601(12):2445-2452
This article reports photoemission and STM studies for the adsorption and dissociation of water on Ce-Au(1 1 1) alloys and CeOx/Au(1 1 1) surfaces. In general, the adsorption of water at 300 K on disordered Ce-Au(1 1 1) alloys led to O-H bond breaking and the formation of Ce(OH)x species. Heating to 500-600 K induced the decomposition or disproportionation of the adsorbed OH groups, with the evolution of H2 and H2O into gas phase and the formation of Ce2O3 islands on the gold substrate. The intrinsic Ce ↔ H2O interactions were explored by depositing Ce atoms on water multilayers supported on Au(1 1 1). After adsorbing Ce on ice layers at 100 K, the admetal was oxidized immediately to yield Ce3+. Heating to room temperature produced finger-like islands of Ce(OH)x on the gold substrate. The hydroxyl groups dissociated upon additional heating to 500-600 K, leaving Ce2O3 particles over the surface. On these systems, water was not able to fully oxidize Ce into CeO2 under UHV conditions. A complete Ce2O3 → CeO2 transformation was seen upon reaction with O2. The particles of CeO2 dispersed on Au(1 1 1) did not interact with water at 300 K or higher temperatures. In this respect, they exhibited the same reactivity as does a periodic CeO2(1 1 1) surface. On the other hand, the Ce2O3/Au(1 1 1) and CeO2−x/Au(1 1 1) surfaces readily dissociated H2O at 300-500 K. These systems showed an interesting reactivity for H2O decomposition. Water decomposed into OH groups on Ce2O3/Au(1 1 1) or CeO2−x/Au(1 1 1) without completely oxidizing Ce3+ into Ce4+. Annealing over 500 K removed the hydroxyl groups leaving behind CeO2−x/Au(1 1 1) surfaces. In other words, the activity of CeOx/Au(1 1 1) for water dissociation can be easily recovered. The behavior of gold-ceria catalysts during the water-gas shift reaction is discussed in light of these results.  相似文献   

13.
The surface chemistry of NO and NO2 on clean and oxygen-precovered Pt(1 1 0)-(1 × 2) surfaces were investigated by means of high resolution electron energy loss spectroscopy (HREELS), X-ray photoelectron spectroscopy (XPS) and thermal desorption spectroscopy (TDS). At room temperature, NO molecularly adsorbs on Pt(1 1 0), forming linear NO(a) and bridged NO(a). Coverage-dependent repulsive interactions within NO(a) drive the reversible transformation between linear and bridged NO(a). Some NO(a) decomposes upon heating, producing both N2 and N2O. For NO adsorption on the oxygen-precovered surface, repulsive interactions exist between precovered oxygen adatoms and NO(a), resulting in more NO(a) desorbing from the surface in the form of linear NO(a). Bridged NO(a) experiences stronger repulsive interactions with precovered oxygen than linear NO(a). The desorption activation energy of bridged NO(a) from oxygen-precovered Pt(1 1 0) is lower than that from clean Pt(1 1 0), but the desorption activation energy of linear NO(a) is not affected by the precovered oxygen. NO2 decomposes on Pt(1 1 0)-(1 × 2) surface at room temperature. The resulted NO(a) (both linear NO(a) and bridged NO(a)) and O(a) repulsively interact each other. Comparing with NO/Pt(1 1 0), more NO(a) desorbs from NO2/Pt(1 1 0) as linear NO(a), and both linear NO(a) and bridged NO(a) exhibit lower desorption activation energies. The reaction pathways of NO(a) on Pt(1 1 0), desorption or decomposition, are affected by their repulsive interactions with coexisting oxygen adatoms.  相似文献   

14.
Spontaneous reaction rate oscillations and spatio-temporal patterns have been observed by mass spectrometry and photoemission electron microscopy (PEEM) during the reduction of NO by NH3 on polycrystalline platinum at 1 × 10−4 Torr and temperatures from 460-520 K. The appearance of both oscillations and patterns was found to be strongly dependent on the gas phase composition and the temperature. In addition, the overall dynamics of the catalyst were found to be dominated by the nonlinear behavior of Pt(1 0 0) type grains, while other types of grains did not participate. In contrast to previous studies, a large number of complex multimodal oscillations were observed, particularly as the coupling between the surface and the gas phase was increased. The appearance of these complex oscillations demonstrates the importance of gas phase coupling to understanding catalytic reactions, even in high vacuum systems.  相似文献   

15.
Yuhai Hu 《Surface science》2009,603(2):336-2840
Interactions between S18O2 and NO on the surface of stepped Pt(3 3 2) were studied using Fourier transform infra red reflection-absorption spectroscopy (FTIR-RAS) combined with thermal desorption spectroscopy (TDS). Adsorbed S18O2 does not seem to have a preference for step sites on Pt(3 3 2). As such, the presence of S18O2 molecules following exposures of ?1.6 L does not significantly block the subsequent adsorption of NO (?0.8 L) on these step sites. Adsorbed S18O2 molecules undergo dissociation (S18O2(a) → S18O(a) + 18O(a)) as the surface temperature is increased to 250 K and above, but the resultant 18O(a) further reacts with sulfur oxides (S18O2(a) and S18O(a)) to form S18Ox (x > 2) species at ∼400 K and above. The S18Ox species desorb as S18O2. Even though the presence of co-adsorbed S18O2 suppresses NO dissociation and subsequent N2 production, this effect is not significantly enhanced with increasing the exposures of S18O2 in the range ?1.6 L; N2 desorption is still detectable at an exposure of 1.6 L S18O2, at which a considerable amount of S18O2 desorption is detected.  相似文献   

16.
Our recent studies of the steady-state kinetics of the N2O-CO reaction on Rh(1 1 0) indicate that at CO excess, the reaction rate increases with increasing temperature. At N2O excess, the reaction rate is nearly independent of temperature at T < 520 K and rapidly decreases with increasing temperature at T > 520 K. Our present analysis of the relevant data indicates that the latter feature seems to be related to surface-oxide formation. Following this line, we propose a mean-field kinetic model making it possible to describe and clarify the experiment.  相似文献   

17.
The surface reaction and desorption of sulfur on Rh(1 0 0) induced by O2 and H2O are investigated with X-ray photoelectron spectroscopy (XPS) technique. The Rh(1 0 0) sample covered with atomic sulfur is prepared by means of the exposure to H2S gas, and subsequently the sample is annealed under O2 or H2O atmosphere. The XPS results show that atomic sulfur adsorbed on Rh(1 0 0) reacts with O2 and desorbs from the surface at 473 K or more. On the other hand, atomic sulfur can not be removed from Rh(1 0 0) surface by H2O at any temperature.  相似文献   

18.
The effect of the irradiation with Al Kα X-rays during an XPS measurement upon the surface vanadium oxidation state of a fresh in vacuum cleaved V2O5(0 0 1) crystal was examined. Afterwards, the surface reduction of the V2O5(0 0 1) surface under Ar+ bombardment was studied. The degree of reduction of the vanadium oxide was determined by means of a combined analysis of the O1s and V2p photoelectron lines. Asymmetric line shapes were needed to fit the V3+2p photolines, due to the metallic character of V2O3 at ambient temperature. Under Ar+ bombardment, the V2O5(0 0 1) crystal surface reduces rather fast towards the V2O3 stoichiometry, after which a much slower reduction of the vanadium oxide occurs.  相似文献   

19.
STM, STS, LEED and XPS data for crystalline θ-Al2O3 and non-crystalline Al2O3 ultra-thin films grown on NiAl(0 0 1) at 1025 K and exposed to water vapour at low pressure (1 × 10−7-1 × 10−5 mbar) and room temperature are reported. Water dissociation is observed at low pressure. This reactivity is assigned to the presence of a high density of coordinatively unsaturated cationic sites at the surface of the oxide film. The hydroxyl/hydroxide groups cannot be directly identify by their XPS binding energy, which is interpreted as resulting from the high BE positions of the oxide anions (O1s signal at 532.5-532.8 eV). However the XPS intensities give evidence of an uptake of oxygen accompanied by an increase of the surface coverage by Al3+ cations, and a decrease of the concentration in metallic Al at the alloy interface. A value of ∼2 for the oxygen to aluminium ions surface concentration ratio indicates the formation of an oxy-hydroxide (AlOxOHy with x + y ∼ 2) hydroxylation product. STM and LEED show the amorphisation and roughening of the oxide film. At P(H2O) = 1 × 10−7 mbar, only the surface of the oxide film is modified, with formation of nodules of ∼2 nm lateral size covering homogeneously the surface. STS shows that essentially the valence band is modified with an increase of the density of states at the band edge. With increasing pressure, hydroxylation is amplified, leading to an increased coverage of the alloy by oxy-hydroxide products and to the formation of larger nodules (∼7 nm) of amorphous oxy-hydroxide. Roughening and loss of the nanostructure indicate a propagation of the reaction that modifies the bulk structure of the oxide film. Amorphisation can be reverted to crystallization by annealing under UHV at 1025 K when the surface of the oxide film has been modified, but not when the bulk structure has been modified.  相似文献   

20.
A tin layer 0.8 nm thick was deposited onto the CeO2(1 1 1) surface by molecular beam deposition at a temperature of 520 K. The interaction of tin with cerium oxide (ceria) was investigated by X-ray photoelectron spectroscopy (XPS), ultra-violet photoelectron spectroscopy (UPS) and resonant photoelectron spectroscopy (RPES). The strong tin-ceria interaction led to the formation of a homogeneous bulk Ce-Sn-O mixed oxide system. The bulk compound formation is accompanied by partial Ce4+ → Ce3+ reduction, observed as a giant 4f resonance enhancement of the Ce3+ species. CeO2 and SnO2 oxides were formed after oxygen treatment at 520 K. The study proved the existence of strong Ce-Sn interaction and charge transfer from Sn to the Ce-O complex that lead to a weakening of the cerium-oxygen bond, and consequently, to the formation of oxygen deficient active sites on the ceria surface. This behavior can be a key for understanding the higher catalytic activity of the SnOx/CeOx mixed oxide catalysts as compared with the individual pure oxides.  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