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 adsorption/decomposition kinetics/dynamics of thiophene has been studied on silica-supported Mo and MoSx clusters. Two-dimensional cluster formation at small Mo exposures and three-dimensional cluster growth at larger exposures would be consistent with the Auger electron spectroscopy (AES) data. Thermal desorption spectroscopy (TDS) indicates two reaction pathways. H4C4S desorbs molecularly at 190–400 K. Two TDS features were evident and could be assigned to molecularly on Mo sites, and S sites adsorbed thiophene. Assuming a standard preexponential factor (ν = 1 × 1013/s) for first-order kinetics, the binding energies for adsorption on Mo (sulfur) sites amount to 90 (65) kJ/mol for 0.4 ML Mo exposure and 76 (63) kJ/mol for 2 ML Mo. Thus, smaller clusters are more reactive than larger clusters for molecular adsorption of H4C4S. The second reaction pathway, the decomposition of thiophene, starts at 250 K. Utilizing multimass TDS, H2, H2S, and mostly alkynes are detected in the gas phase as decomposition products. H4C4S bond activation results in partially sulfided Mo clusters as well as S and C residuals on the surface. S and C poison the catalyst. As a result, with an increasing number of H4C4S adsorption/desorption cycles, the uptake of molecular thiophene decreases as well as the H2 and H2S production ceases. Thus, silica-supported sulfided Mo clusters are less reactive than metallic clusters. The poisoned catalyst can be partially reactivated by annealing in O2. However, Mo oxides also appear to form, which passivate the catalyst further. On the other hand, while annealing a used catalyst in H/H2, it is poisoned even more (i.e., the S AES signal increases). By means of adsorption transients, the initial adsorption probability, S0, of C4H4S has been determined. At thermal impact energies (Ei = 0.04 eV), S0 for molecular adsorption amounts to 0.43 ± 0.03 for a surface temperature of 200 K. S0 increases with Mo cluster size, obeying the capture zone model. The temperature dependence of S0(Ts) consists of two regions consistent with molecular adsorption of thiophene at low temperatures and its decomposition above 250 K. Fitting S0(Ts) curves allows one to determine the bond activation energy for the first elementary decomposition step of C4H4S, which amounts to (79 ± 2) kJ/mol and (52 ± 4) kJ/mol for small and large Mo clusters, respectively. Thus, larger clusters are more active for decomposing C4H4S than are smaller clusters.  相似文献   

2.
We have studied desorption of 13CO and H2O and desorption and reaction of coadsorbed, 13CO and H2O on Au(310). From the clean surface, CO desorbs mainly in, two peaks centered near 140 and 200 K. A complete analysis of desorption spectra, yields average binding energies of 21 ± 2 and 37 ± 4 kJ/mol, respectively. Additional desorption states are observed near 95 K and 110 K. Post-adsorption of H2O displaces part of CO pre-adsorbed at step sites, but does not lead to CO oxidation or significant shifts in binding energies. However, in combination with electron irradiation, 13CO2 is formed during H2O desorption. Results suggest that electron-induced decomposition products of H2O are sheltered by hydration from direct reaction with CO.  相似文献   

3.
In this paper we review the preparation and reaction properties of ordered SmRh surface alloys and SmOx/Rh(1 0 0) model catalyst which have been systematically investigated by low energy electron diffraction (LEED), Auger electron spectroscopy (AES), X-ray photoelectron spectroscopy (XPS), high-resolution electron energy loss spectroscopy (HREELS) and temperature desorption spectroscopy (TDS). The growth of Sm on Rh(1 0 0) at room temperature follows the Stranski-Krastanov mode. Thermal treatment of the Sm films on Rh(1 0 0) leads to the formation of ordered SmRh surface alloys. An “inverse” SmOx/Rh(1 0 0) model catalyst is produced under controlled oxidation of the SmRh surface alloy. CO adsorption on the SmRh alloy and SmOx/Rh(1 0 0) surfaces gives rise to five TDS characteristic features originating from different adsorption sites. Both the site blocking of SmOx and the electron transfer between SmOx and Rh substrate significantly affect the CO adsorption. Acetate decomposition on both Rh(1 0 0) and the SmOx/Rh(1 0 0) surfaces are found to undergo two competitive pathways that yields either (i) CO(a) and O(a) or (ii) CO2(g) and H2(g) at high temperature. The reactive desorption of acetic acid on SmOx/Rh(1 0 0) is dramatically different from that on Rh(1 0 0). A rapid decomposition of acetic acid to produce CO(g) and CO2(g) can be observed only on SmOx/Rh(1 0 0), where CO(g) becomes the predominant product at 225 K, indicating a strong promotional effect of SmOx in the selective decomposition of acetate. Finally, we briefly describe the future outlook of research on rare earth oxide/metal model catalysts.  相似文献   

4.
S.J. Alas  L. Vicente 《Surface science》2010,604(11-12):957-964
The kinetics of NO desorption and its decomposition on Rh(111) surfaces have been simulated by using a dynamic Monte Carlo method. During the simulations, we used a triangular lattice that mimics the Rh(111) phase. NO decomposition was studied at low pressure and temperatures ranging from 120 to 1000 K. The present analysis incorporates recent experimental evidence showing that N2 production occurs either from the classical N + N recombination step or by the formation and successive decay of an (N–NO)* intermediate species. Moreover, N2 and NO desorption rates are enhanced and the NO dissociation rate is inhibited by coadsorbed NO, N, and O species as nearest neighbors. These effects are taken into account in this study, along with the experimental adsorption, desorption, dissociation, and diffusion rates of the reactants. Our simulations are consistent with the experimental results of TPD spectra and can explain the formation of two peaks, δ-N2 and β-N2, as a natural consequence of the reaction mechanism herein proposed. Comparisons with different mechanisms used by other authors are also made.  相似文献   

5.
A sharp change in the N2 emission channel from N2O(a)  N2(g) + O(a) to N(a) + N(a)  N2(g) has been found at around 500 K in a steady-state NO + D2 reaction over stepped Pd(211) = [(S)3(111) × (100)] by means of angle-resolved desorption. The desorbing N2 is highly collimated at around 30° off normal toward the step-down direction below about 500 K due to the intermediate N2O decomposition, whereas, above 500 K, the near normally directed desorption due to the recombination of N(a) is relatively enhanced. The N2O decomposition channel is promoted when the reaction is carried out with hydrogen (deuterium) and the channel change is accelerated by quick changes of the amounts of surface hydrogen and oxygen (or NO(a)) into the opposite directions, and enhanced nitrogen removal as ammonia on the resultant hydrogen-rich surface. In the steady-state NO + CO reaction, the N2 emission channel gradually changes above 500 K toward recombination. A model for the off-normal N2 emission is briefly described.  相似文献   

6.
The adsorption structure of nitric oxide (NO) on Ir(111) was studied by thermal desorption spectroscopy (TDS) and dynamical analyses of low-energy electron diffraction (LEED). At the saturation coverage at about 100 K, a 2 × 2 pattern was observed by LEED and two peaks appeared at 365 and 415 K in TDS. No change in the LEED IV curves was observed by annealing at 280 K, which means that the NO-saturated surface was retained at this temperature. On the contrary, partial desorption and changes of the LEED IV curves were observed by annealing at 360 K. Combined with previous vibrational studies, it is suggested that one adsorption species is not affected, while another species is partially desorbed and the rest of them are dissociated by annealing at 360 K. Dynamical analyses of LEED were performed for the 280 K-annealed and the 360 K-annealed surfaces, which correspond to the NO-saturated and the NO-dissociated Ir(111) surfaces, respectively. These revealed that NO occupies the atop, fcc-hollow and hcp-hollow sites (atop-NO + fcc-NO + hcp-NO) for the NO-saturated Ir(111) surface with the saturation coverage of 0.75 ML. For the 360 K-annealed surface, the atop-NO is not affected but the fcc-NO and the hcp-NO are partially desorbed as NO and partially dissociated to N and O, both of which occupy the fcc-hollow site on the surface.  相似文献   

7.
《Physica B: Condensed Matter》2005,355(1-4):202-206
Specific heat (SH) measurements on TbMn2(H,D)2 powders have been performed in the temperature range from 2 to 350 K, in zero magnetic field and in 9 T. Due to the low heat conductivity of the samples, the measurements were carried out on a mixed Cu- and sample-powder pellet. For TbMn2, the anti-ferromagnetic phase transition was manifest by a single SH peak at TN=47 K, whereas a double SH peak at 281 and 288 K and an upturn below 5 K were observed for the hydride sample. Upon applying the magnetic field of 9 T, the SH upturn was suppressed, whereas no visible influence was found on the specific heat in the whole temperature range above 10 K as well as on the double peak.  相似文献   

8.
The decomposition of N2O was studied on Pd(110) through analysis of the angular and velocity distributions of desorbing products by means of angle-resolved thermal desorption combined with time-of-flight techniques. Both desorption and decomposition of N2O(a) were completed below 170 K, simultaneously emitting N2. N2 showed two desorption peaks, β1-N2 at 152 K and β2-N2 at 134–140 K. The former revealed an inclined emission at ±43° off the normal in a plane along the (001) direction and a hyper-thermal translational energy, whereas the latter showed a cosine angular distribution without an excess translational energy. Different desorption channels were proposed to yield these desorption.  相似文献   

9.
Temperature-programmed-desorption (TPD) spectra and isothermal desorption rates of D2 molecules from a Si(100) surface have been calculated to reproduce experimental β1, A-TPD spectra and isothermal desorption rate curves. In the diffusion-promoted-desorption (DPD) mechanism, hydrogen desorption from the Si(100) (2 × 1) surfaces takes place via D atom diffusion from doubly-occupied Si dimers (DODs) to their adjacent unoccupied Si dimers (UODs). Taking a clustering interaction among DODs into consideration, coverages θDU of desorption sites consisting of a pair of a DOD and UOD are evaluated by a Monte Carlo (MC) method. The TPD spectra for the β1, A peak are obtained by numerically integrating the desorption rate equation R = νA exp(? Ed, A / kBT)θDU, where νA is the pre-exponential factor and Ed, A is the desorption barrier. The TPD spectra calculated for Ed, A = 1. 6 eV and νA = 2.7 × 109 /s are found to be in good agreement with the experimental TPD data for a wide coverage range from 0.01 to 0.74 ML. Namely, the deviation from first-order kinetics observed in the coverage dependent TPD spectra as well as in the isothermal desorption rate curves can be reproduced by the model simulations. This success in reproducing both the experimental TPD data and the very low desorption barrier validates the proposed DPD mechanism.  相似文献   

10.
《Solid State Ionics》2006,177(7-8):803-811
The purpose of this study was to synthesize highly dispersed Ni/Al2O3 catalysts and to develop a suitable hydrogen-temperature programmed desorption (H2-TPD) method for the determination of nickel metal surface area, dispersion, and crystallite sizes. Several highly dispersed Ni/Al2O3 catalysts with a Ni loading between 15 and 25 wt.% were synthesized. The reducibility of catalysts was determined by temperature programmed reduction (TPR) experiments. All catalysts exhibited a single reduction peak with a maximum rate of H2 consumption (Tmax in TPR) occurring below 450 °C. Three different H2-TPD methods were employed to determine the amount of H2 chemisorbed. In TPD-1, a 10% H2/Ar mixture was used for catalyst pre-reduction and surface saturation by cooling down from Tmax in TPR to room temperature. In TPD-2, the catalyst surface after pre-reduction was flushed with Ar at Tmax in TPR + 10 °C. The TPD-3 was similar to the TPD-2, but used 100% H2 instead of 10% H2/Ar mixture. In all three TPD methods, the profiles exhibited 2 domains of H2 desorption peaks, one below 450 °C, referred to as type-1 peaks, and attributed to H2 desorbed from exposed fraction of Ni atoms, and the other above 450 °C, denoted as type-2 peaks, and assigned to the desorption of H2 located in the subsurface layers and/or to spillover H2. Flushing the reduced catalyst surface in Ar at Tmax in TPR + 10 °C in TPD-2 and TPD-3 removed most of the H2 located in the subsurface layers/ spillover H2. The amount of H2 chemisorbed to form a monolayer on the reduced Ni/Al2O3 catalysts was determined quantitatively from the TPD peak areas of type-1 peaks in TPD-1, and from both type-1 and type-2 peaks in TPD-2 and TPD-3. The Ni metal surface area, dispersions and crystallite sizes were calculated from the chemisorption data and the values were compared with those obtained using the static chemisorption method. Both TPD-2 and TPD-3 gave chemisorption results similar to that obtained from the static method.  相似文献   

11.
Phase transformations in squaric acid (H2C4O4) have been investigated by thermogravimetry and differential scanning calorimetry with different heating rates β. The mass loss in TG apparently begins at onset temperatures Tdi=245±5 °C (β=5 °C min?1), 262±5 °C (β=10 °C min?1), and 275±5 °C (β=20 °C min?1). A polymorphic phase transition was recognized as a weak endothermic peak in DSC around 101 °C (Tc+). Further heating with β=10 °C min?1 in DSC revealed deviation of the baseline around 310 °C (Ti), and a large unusual exothermic peak around 355 °C (Tp), which are interpreted as an onset and a peak temperature of thermal decomposition, respectively. The activation energy of the thermal decomposition was obtained by employing relevant models. Thermal decomposition was recognized as a carbonization process, resulting in amorphous carbon.  相似文献   

12.
E. Demirci  A. Winkler 《Surface science》2010,604(5-6):609-616
Co-adsorption of hydrogen and CO on Cu(1 1 0) and on a bimetallic Ni/Cu(1 1 0) surface was studied by thermal desorption spectroscopy. Hydrogen was exposed in atomic form as generated in a hot tungsten tube. The Ni/Cu surface alloy was prepared by physical vapor deposition of nickel. It turned out that extended exposure of atomic hydrogen leads not only to adsorption at surface and sub-surface sites, but also to a roughening of the Cu(1 1 0) surface, which results in a decrease of the desorption temperature for surface hydrogen. Exposure of a CO saturated Cu(1 1 0) surface to atomic H leads to a removal of the more strongly bonded on-top CO (α1 peak) only, whereas the more weakly adsorbed CO molecules in the pseudo threefold hollow sites (α2 peak) are hardly influenced. No reaction between CO and H could be observed. The modification of the Cu(1 1 0) surface with Ni has a strong influence on CO adsorption, leading to three new, distinct desorption peaks, but has little influence on hydrogen desorption. Co-adsorption of H and CO on the Ni/Cu(1 1 0) bimetallic surface leads to desorption of CO and H2 in the same temperature regime, but again no reaction between the two species is observed.  相似文献   

13.
An interstitial Pr3(Fe,Ti)29 hydride was synthesised by gas-phase hydrogenation on Pr3(Fe,Ti)29 powder using H2. The reaction kinetics between Pr3(Fe,Ti)29 and H2 gases was studied in a constant-volume reactor. The sample starts to rapidly absorb hydrogen, interstitially, at about 533 K. Absorption passes through a maximum at about 598 K (1.6 H/f.u) and then interstitial hydrogen desorption takes place up to the temperature of 673 K. By cooling to room temperature, the sample absorbs more hydrogen, interstitially, reaching the value of 3.6 H/f.u. By remaining at room temperature, the sample absorbs even more hydrogen reaching the value of 5.2 H/f.u. The lattice expansion observed is 2.1% and the Curie temperature, TC, increased from 392 to 518 K. The hydride exhibits saturation magnetisation, MS, of 145.4 and 157.5 Am2/kg at room temperature (RT) and at 5 K, respectively, anisotropy field, HA, of 2.1 T (RT) and 4.5 T (5 K) and average hyperfine field, Heff, of 23.3 T (RT). The magnetic anisotropy of Pr3(Fe,Ti)29 hydride is the same as that of the parent compounds, easy-cone-like, changing only in the cone angle (from 34° to 26°).  相似文献   

14.
The formation of complex species of dioxouranium(VI) ion with EDTA was studied in the pH range of 1–3.5 and at 25 °C using a combination of potentiometric and spectrophotometric techniques. Results showed evidence for formation of the following species: [UO2H4EDTA]2+, [UO2H3EDTA]+, and [UO2H2EDTA]. Investigations were performed in sodium perchlorate as background electrolyte at 0.1, 0.3, 0.5, 0.7, and 1.0 mol dm? 3. The parameters based on the formation constants were calculated, and the dependences of protonation and the stability constants on ionic strength are described. The dependence on ionic strength of the formation constants was analyzed using the specific ion interaction theory (SIT) model. The stability constant values at infinite dilution, obtained using SIT model, are log β°141 = 6.77, log β°131 = 5.99 and log β°121 = 9.29, where indexes for the overall stability constant, βpqr, refer to the equilibrium pUO22+ + qH+ + rL4? ? MpHqLr(2p + q ? 4r). The specific interaction coefficients are also reported.  相似文献   

15.
CO adsorption on clean and oxidized Pt3Ti(111) surfaces has been investigated by means of Auger Electron Spectroscopy (AES), Thermal Desorption Spectroscopy (TDS), Low Energy Electron Diffraction (LEED) and High Resolution Electron Energy Loss Spectroscopy (HREELS). On clean Pt3Ti(111) the LEED patterns after CO adsorption exhibit either a diffuse or a sharp c(4 × 2) structure (stable up to 300 K) depending on the adsorption temperature. Remarkably, the adsorption/desorption behavior of CO on clean Pt3Ti(111) is similar to that on Pt(111) except that partial CO decomposition on Ti sites and partial CO oxidation have also been evidenced. Therefore, the clean surface cannot be terminated by a pure Pt plane. Partially oxidized Pt3Ti(111) surfaces (< 135 L O2 exposure at 1000 K) exhibit a CO adsorption/desorption behavior rather similar to that of the clean surface, showing again a c(4 × 2) structure (stable up to 250 K). Only the oxidation of CO is not detectable any more. These results indicate that some areas of the substrate remain non-oxidized upon low oxygen exposures. Heavily oxidized Pt3Ti(111) surfaces (> 220 L O2 exposure at 1000 K) allow no CO adsorption indicating that the titanium oxide film prepared under these conditions is completely closed.  相似文献   

16.
(Gd,Y)Ba2Cu3Ox tapes have been fabricated by metal organic chemical vapor deposition (MOCVD) with Zr-doping levels of 0–15 mol.% and Ce doping levels of 0–10 mol.% in 0.4 μm thick films. The critical current density (Jc) of Zr-doped samples at 77 K, 1 T applied in the orientation of H 6 c is found to increase with Zr content and shows a maximum at 7.5% Zr doping. The 7.5% Zr-doped sample exhibits a critical current density (Jc) of 0.95 MA/cm2 at H 6 c which is more than 70% higher than the Jc of the undoped sample. The peak in Jc at H 6 c is 83% of that at H 6 ab in the 7.5% Zr-doped sample which is more than twice as that in the undoped sample. Superconducting transition temperature (Tc) values as high as about 89 K have been achieved in samples even with 15% Zr and 10% Ce. Ce-doped samples with and without Ba compensation are found to exhibit substantially different Jc values as well as angular dependence characteristics.  相似文献   

17.
Karl Jacobi  Yuemin Wang 《Surface science》2009,603(10-12):1600-1604
The interaction of NO with the O-rich RuO2(1 1 0) surface, exposing coordinatively unsaturated O-bridge, O-cus, and Ru-cus atoms, was studied at 300 K by thermal desorption spectroscopy (TDS) and high-resolution electron energy-loss spectroscopy (HREELS). The conclusions are validated by isotope substitution experiments with 18O. During exposure to NO an O···N–O surface group (NO2-cus) is formed with O-cus. Additionally, a smaller number of empty Ru-cus sites are filled by NO-cus. If one warms the sample to 400 K, NO2-cus does not desorb but decomposes into O and NO again, the latter being either released into gas phase or adsorbed as NO-cus. With O-bridge such a surface group is not stable at 300 K. Our experiments further prove that O-cus is more reactive than O-bridge.  相似文献   

18.
The adsorption and desorption of the system CO/Pt(111) and C6H6/Pt(111) at 300 K has been investigated with a pulsed molecular beam method in combination with a microcalorimeter. For benzene the sticking probability has been measured in dependence of the coverage θ. For coverages θ > 0.8 transient adsorption is observed. From an analysis of the time-dependence of the molecular beam pulses the rate constant for desorption is determined to be 5.6 s? 1. With a precursor-mediated kinetic adsorption model this allows to obtain also the hopping rate constant of 95.5 s? 1. The measured adsorption enthalpies could be best described by (199 ? 77θ ? 51θ2) kJ/mol, in good agreement with the literature values. For CO on Pt(111) also transient adsorption has been observed for θ > 0.95 at 300 K. The kinetic analysis yields rate constants for desorption and hopping of 20 s?1 and 51 s?1, respectively. The heats of adsorption show a linear dependence on coverage (131 ? 38θ) kJ/mol between 0  θ  0.3, which is consistent with the desorption data from the literature. For higher coverage (up to θ = 0.9ML) a slope of ?63 kJ/mol describes the decrease of the differential heat of adsorption best. This result is only compatible with desorption experiments, if the pre-exponential factor decreases strongly at higher coverage. We found good agreement with recent quantum chemical calculations made for (θ = 0.5ML).  相似文献   

19.
Adsorption and decomposition of NO on Pt (1 1 2) have been studied by temperature programmed desorption (TPD), ultraviolet photoelectron spectroscopy (UPS) and X-ray photoelectron spectroscopy (XPS). NO adsorbs molecularly on Pt (1 1 2) at 95 K. About half amount of NO molecules adsorbs at the terrace sites and remaining half amount adsorbs at the step sites at a full monolayer coverage. Then about half of NO molecules adsorbed at step sites decomposes at around 483 K desorbing N2, promptly.  相似文献   

20.
The structural and chemical characterization of Rh, Mo and Rh–Mo nanosized clusters formed by physical vapor deposition on TiO2 single crystal was performed by Auger Electron Spectroscopy (AES), Thermal Desorption Spectroscopy (TDS) and Reflection Absorption Infrared Spectroscopy (RAIRS), applying CO as test molecule. On a slightly reduced titania surface 2D-like growth of Rh was revealed at 300 K up to 0.23 ML coverage by AES and CO-desorption experiments. For CO-saturated Rh particles TDS showed molecular CO desorption in a broad temperature range with Tp = 400, 440, 490 and 540 K (α-states), the latter state appearing only on the smallest Rh particles. The population of γ-state (Tp = 780–820 K) originating from the recombination of C and O atoms on the support began at ΘRh = 0.23ML and was maximized at around 1–2 ML Rh coverage, corresponding to 30% dissociation of CO. A possible dissociation precursor on Rh particles is identified as linearly bonded CO on step sites characterized by ν(C–O) of 2017 cm? 1. Deliberation of CO2 could not be detected between 170 and 900 K, showing the absence of disproportionation reaction. Instead of oxidizing CO molecules, oxygen atoms stemming from the dissociation of CO attached to the reduced centers of titania, indicating the role of adsorption sites at the perimeter of Rh particles in the decomposition process. 2 ML of predeposited Mo enhanced markedly the dispersion of Rh particles as a result of strong Rh–Mo interaction, but it slightly reduced the molecular α-CO desorption possibly due to enhanced dissociation. The formation of γ-CO was suppressed considerably through elimination of adsorption centers by Mo on the TiO2 substrate. The reactivity of Rh layers deposited on Mo-covered surface towards CO was reduced after repeated annealing to 600 K due to partial encapsulation of Rh by titania, manifesting in the suppression of the more strongly bonded α-state. Mo-deposits (up to 0.5ML) on Rh particles decreased the saturation coverage of α-CO through a site-blocking mechanism without detectable influence on the binding energy of CO to Rh, indicating Mo island formation. The carbon arising from the decomposition of CO dissolved in the Mo-containing particles formed a solid solution stable even at 900 K, suggesting a possible role of molybdenum carbide regarding the enhanced catalytic activity of Rh clusters.  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