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 idea of an activation complex is popular for explaining reaction rates, but the characteristics of reactions and catalysis may not be explained in this way. A predestined state for each reaction composed of surface atoms and adsorbed species is responsible for these features. Two single Sn atoms trapped in adjacent half-unit cells of an Si(111) 7 × 7 surface is an example of a predestined state. An isolated Sn atom in a half-unit cell does not migrate to other half-unit cells at room temperature, but when two single Sn atoms are in adjacent half-unit cells they undergo rapid combination to form an Sn2 dimer. In addition, these two single Sn atoms replace the center Si adatoms and an Si4 cluster is formed. The spatial distribution of molecules desorbing from surfaces may reflect the predestined states for the desorption processes. The spatial distribution in the temperature-programmed desorption (TPD) of NO on Pd(110) and Pd(211) surfaces and that in the temperature-programmed reaction (TPR) of NO + H2 were studied. N2 desorbing from Pd(110) by the recombination of N atoms obeys cos6 – cos7 but the N2 produced by a catalytic reaction of NO with H2 obeys cos. In contrast, the N2 desorbing with NO at 490 K in the TPD of Pd(110) shows a sharp off-normal distribution expressed by cos46( – 38). The adsorption of NO on Pd(211) predominantly occurs on the (111) terrace but the spatial distribution suggests that the predestined states for the reaction and desorption are formed on both the (111) terrace and (100) step surfaces.  相似文献   

2.
The effect of palladium particle size on its catalytic activity was investigated by the decomposition of chemisorbed nitric oxide and the reaction of nitric oxide with carbon monoxide in flow conditions. Palladium particles (30–500 Å) were prepared on silica thin films (100 Å) which were supported on a Mo(110) surface. The reactivity of the supported palladium varied with the metal particle size. On large palladium particles, nitric oxide (NO) reacts to form nitrous oxide (N2O), dinitrogen (N2) and atomic oxygen during temperature-programmed reaction, whereas on small particles (< 50 Å), nitrous oxide is not formed. Similarly, reactions of NO with CO on large particles, in flow conditions produce N2O, N2 and CO2, whereas N2O is not produced on small particles. In addition, more extensive NO decomposition is observed on the smaller particles.  相似文献   

3.
Reaction rates and product selectivities were measured over the Rh(100) surface as a function of temperature, and CO and NO partial pressures. These results are compared with our prior studies of the NO–CO reaction on the Rh(111) and Rh(110) surfaces. The only products detected for all three surfaces were CO2, N2O, and N2. Furthermore, for the Rh(100) surface we have found a significant change in the apparent activation energy (E a) with reaction temperature. For the Rh(100) surface it was found that the E a can change by a factor of 2.3 in the temperature range investigated here, from 528 to 700 K, with the lower values obtained at higher temperatures. In contrast, E a's were found to remain constant over the same temperature range for the Rh(110) and Rh(111) surfaces. The results observed for the Rh(100) surface suggest that reaction kinetics are dominated by variations in NO coverages. At low temperatures, the surface is fully saturated with NO, and dissociation is limited by the availability of vacancy sites through NO desorption. At high temperatures, the surface is still primarily covered with NO, however, the number of vacancy sites has increased substantially. In this case, we propose that the apparent activation energy is now reflecting NO dissociation kinetics rather than those for NO desorption. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

4.
A temperature programmed reaction/desorption (TPD) study of decomposition pathways of methanol, ethanol, 1-propanol and 2-propanol was conducted on the clean Pd(110) surface under ultra-high vacuum conditions. No alcohol underwent C-O scission. Alcohols appear to react on this clean surface via the same dehydrogenation and decarbonylation steps observed on the Pd(111) surface. In contrast to previous reports noting substantial differences in methanol chemistry on the Pt(110) and (111) surfaces, the reactions of methanol and ethanol were found to be the same on the Pd(110) and (111) surfaces, giving rise to H2 plus CO from methanol, and H2, CO, and CH4 from ethanol. The C3 alcohols, 1- and 2-propanol, did produce somewhat different products on the Pd(110) and (111) surfaces, but these differences can be accounted for by differences in the chemistry of intermediate reaction products, rather than different reaction pathways of the parent alcohols.  相似文献   

5.
The thermal decomposition pathways of isobutene and 1-butene on both Mo(110) and 4 × 4-C/Mo(110) surfaces have been studied using high-resolution electron energy loss spectroscopy (HREELS) in order to highlight the substantially different activities of these two surfaces towards the cleavage of C–H and C–C bonds. On clean Mo(110), the CH2 group of isobutene decomposes upon heating to 150 K, producing either a /-bonded isobutenylidene [(CH3)2CCH] species or a 1,1-di-/-bonded isobutenyl [(CH3)2CC] species. Upon further heating, extensive C–H bond scission occurs to form hydrocarbon fragments which do not contain CH3 or CH2 groups, but appear to have largely intact carbon skeletons. By contrast, isobutene is molecularly adsorbed on the carbide-modified surface at 150 K. Further heating produces isobutylidyne [(CH3)2HCC] by 300 K, which subsequently decomposes via C–C bond scission to generate surface methyl groups. The different activation sequence of the C–H and C–C bonds of isobutene on clean and carbide-modified Mo(110) surfaces is also qualitatively confirmed by comparative studies of 1-butene on the two surfaces.  相似文献   

6.
The reduction of NO by H2 was studied over three different Pt-Rh single crystal surfaces, i.e. Pt-Rh(111), (410) and (100). The adsorption and dissociation of NO was studied by HREELS, LEED, XPS, AES and TDS. It was found that the dissociation of NO and its reaction with H2 is very surface structure sensitive. The selectivity towards nitrogen and the dissociation activity increases in the same order, i.e. Pt-Rh(111) < (100) < 410). Nitrogen atoms were easily hydrogenated at 400 K in hydrogen to NH x (x = 1 or 2) on the surface. A model is proposed in which the selectivity of the NO-H2 reaction over Pt-Rh surfaces is determined by the relative amounts of hydrogen, NO and nitrogen adatoms on the surface.  相似文献   

7.
The results of the LRS study of N-containing chemisorbed species on the promoted iron catalyst for ammonia synthesis have further substantiated the existence of two species of N2(a) as the dominant N-containing chemisorbed species under the functioning catalyst conditions. A model of active site, as 3-Fe cluster on (111) or (211) surface of -Fe, and two modes of multinuclear coordination activation for the observed two species of N2(a) were proposed. It was further illustrated from reaction energetics that the mechanism of the dominant reaction pathway for ammonia synthesis/decomposition may be associative, rather than dissociative.The work was supported by a grant from the National Natural Science Foundation of China; parts of the work were presented at 24th ICCC (Athens, 1986).  相似文献   

8.
A systematic study over Pt/Al2O3 powder and monolith catalysts is carried out using temporal analysis of products (TAP) to elucidate the transient kinetics of NO decomposition and NO reduction with H2. NO pulsing and NO–H2 pump-probe experiments demonstrate the effect of catalyst temperature, NO–H2 pulse delay time and H2/NO ratio on N2, N2O and NH3 selectivity. At lower temperature (150 °C) decomposition of NO is negligible in the absence of H2, indicating that N–O bond scission is rate limiting. At higher temperature NO decomposition occurs readily on reduced Pt but the rate is inhibited by surface oxygen as reaction occurs. The reduction of NO by a limiting amount of H2 at lower temperature indicates the reaction of surface NO with H adatoms to form N adatoms, which react with adsorbed NO to form N2O or recombine to form N2. In excess H2, higher temperatures and longer delay times favor the production of N2. The longer delay enables NO decomposition on reduced Pt with the role of H2 being a scavenger of surface oxygen. Lower temperatures and shorter delay times are favorable for ammonia production. The sensitive dependence on delay time indicates that the fate of adsorbed NO depends on the concentration of vacant sites for NO bond scission, necessary for N2 formation, and of surface hydrogen, necessary for hydrogenation to ammonia. A mechanistic-based microkinetic model is proposed that accounts for the experimental observations. The TAP experiments with the monolith catalyst show an improved signal due to the reduction of transport restrictions caused by the powder. The improved signal holds promise for quantitative TAP studies for kinetic parameters estimation and model discrimination.  相似文献   

9.
The reactions of acetone oxime, a proposed reaction intermediate for the SCR (Selective Catalytic Reduction) of NO with propane on FeZSM-5, have been studied with 13C and 15N solid state MAS NMR (magic angle spinning nuclear magnetic resonance). FeZSM-5 with three different loading levels was prepared by the sublimation method. The thermal reactions of acetone [2-13C] oxime adsorbed on FeZSM-5 samples with different iron loadings were monitored by 13C MAS NMR by heating to the desired temperature and then cooling to room temperature for data acquisition. For the sample with the lowest iron loading (Fe/Al = 0.11), acetic acid and N-methyl-2-propanamine were formed by the decomposition of acetone oxime. For the samples with the higher iron loadings (Fe/Al = 0.69 and 0.91), acetone, N,N-methyl-2,2-propanediamine, and N-methyl-2-propanimine were formed by the decomposition of acetone oxime. 15N MAS NMR was used to investigate reactions of 15NO and acetone oxime on the FeZSM-5 samples. The formation of gas phase N2 and N2O was observed.  相似文献   

10.
By means of spin polarized density functional theory with the GGA + U framework, the reaction mechanism of CH3OH oxidation on the Co3O4 (110)-B and (111)-B surfaces has been investigated. Adsorption situation and a part of reaction cycle for CH3OH oxidation are clarified. Our results indicated that: i) U value can affect the calculated energetic result significantly; ii) CH3OH can adsorb with surface lattice oxygen atom (O2f/O3f) to form CoO bond directly, and the adsorption of CH3OH and its decomposition products on (110)-B is more stable than on (111)-B, which means CH3OH prefers Co3 + better than Co2 +; iii) on the (110)-B surface, CH3OH can form CO2, H2O and adsorbed H atom. But on the (111)-B surface, CH3OH can just form formaldehyde (CH2O) and adsorbed H atom, this means oxidative capacity of (110)-B (Co3 +) is higher than (111)-B (Co2 +). The possible reasons corresponding to the high oxidative of (110)-B come from both Co3 + and O2f: Co3 + tends to bind adsorbed species for further decomposition and O2f tends to bind more hydrogenation atom involved in methanol due to its low-coordinates number compared to that of O3f.  相似文献   

11.
The adsorption of NO on Au 3D hemispherical crystals (field emitter tips) has been studied by means of pulsed field desorption mass spectrometry (PFDMS) under dynamic gas flow conditions and at 300 K. Local chemical probing of ~200 Au sites in the stepped surface region between the central (111) pole and the peripheral (001) plane leads to the detection of NO+, N2O+ and (NO) species. Obviously, molecular NO adsorption on stepped Au surfaces can lead to dimerization. Nitrous oxide formation probably occurs via the dimer, (NO)2.  相似文献   

12.
A series of Mo2C/SBA-15 catalysts with different Mo contents were prepared by temperature-programmed carburization (TPC). The materials obtained and their oxide precursors (MoO3/SBA-15) were characterized by Nitrogen adsorption-desorption isotherms, X-ray diffraction (XRD), and Fourier transform-infrared (FT-IR) spectroscopy. The activities of the catalysts for deep hydrodesulfurization (HDS) of thiophene were evaluated. The results of N2 adsorption-desorption isotherms indicated that the surface area and pore diameter of the oxide precursors increase after carburization. The XRD patterns show that Mo2C particles are highly dispersed in the SBA-15 ordered mesoporous. The test results show that Mo2C/SBA-15 catalysts have an excellent performance for the deep HDS under the lower temperature region.  相似文献   

13.
Steady‐state N2O decomposition reaction on polycrystalline Pt and Rh surfaces has been studied using a supersonic free‐jet molecular beam (2.1 × 1018 molecules/cm2 s). The energy of the incident N2O beam was controlled by a nozzle heating technique in conjunction with a seeding technique. The decomposition rate shows both translational and vibrational energy dependence on the Pt surface. However, there is also the surface temperature dependence of the decomposition rate even varying the incident beam energy, indicating precursor‐mediated dissociation of N2O on the Pt surface. On the other hand, no energy dependence was observed on the Rh surface, suggesting that the decomposition dynamics are different between Pt and Rh surfaces. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

14.
The kinetics of NO decomposition were investigated over alkali metal-doped Co3O4 catalysts. For all the alkali metal-doped Co3O4 catalysts tested, the presence of O2 caused a decrease in the N2 formation rate with reaction orders between −0.26 and −0.40. The reaction orders with respect to NO were between 1.21 and 1.47, which are higher than unity, suggesting that NO decomposition proceeds via a bimolecular reaction. The observation by in situ Fourier transform infrared (FT-IR) spectroscopy confirmed the presence of nitrite (NO2) species on the surface under NO decomposition conditions. Isotopic transient kinetic analysis performed using 14NO and 15NO revealed that a surface-adsorbed species, probably NO2, serves as an intermediate during NO decomposition. We proposed a reaction mechanism in which the reaction is initiated by NO adsorption onto alkali metals to form NO2 species, which migrates to the interface between the alkali metals and Co3O4, the active sites, and then react with the adsorbed NO species to form N2.  相似文献   

15.
We have studied the NO-CO-O2 reaction over a Rh(111) catalyst by monitoring the reaction products (CO2, N2O, and N2) and the infrared (IR) intensity of surface CO and NO at various partial pressures of NO, CO and O2, and sample temperatures. The selectivity for N2O formation, apparent activation energy for product formation, and NO consumption rate during NO-CO-O2 are identical to those measured during the NO-CO reaction. The IR measurements show that during NO-CO-O2 the same two adsorbed species, NO at 1640 cm-1 and linear CO at ~2040 cm-1, are present in the same surface concentrations as during NO-CO. For this reason the NO-CO-O2 kinetics are dominated by the NO-CO kinetics, the NO consumption is rate limited by dissociation of adsorbed NO, and the N2O selectivity is dominated by surface NO coverage. In contrast, O2 consumption is adsorption rate limited with the NO-CO adsorption-desorption equilibrium controlling the vacant sites required to dissociatively adsorb O2. These kinetic and IR data of the CO-NO-O2 reaction and our interpretation of them agree with previous studies over supported Rh catalysts and thus confirm the previously proposed explanation. From RAIRS and kinetic data we estimate the rate constant for the CO+O→CO2 elementary step. The pre-exponential factor for this rate is 2×1010 s-1, a factor of 50 smaller than previous estimates. This rate constant is important to the NO-CO-O2 kinetics because it affects O coverage, which, under certain conditions, inhibits NO consumption. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

16.
The dehydrogenation and decomposition of cyclohexene on the Pt-modified C/W(111) surfaces have been studied by temperature-programmed desorption (TPD), Auger electron spectroscopy (AES) and high-resolution electron energy loss spectroscopy (HREELS). The objective of the current study is to investigate how the surface reactivity of tungsten carbide is modified by the presence of submonolayer Pt. Similar to that observed on Pt(111), Pt(100) and C/W(111) surfaces, the characteristic reaction pathway on Pt/C/W(111) is the selective dehydrogenation of cyclohexene to benzene. At a Pt coverage of 0.52 monolayer, the selectivity to the gas-phase benzene product is 86±7%, which is slightly higher than that on Pt(111) (75%) and on C/W(111) (67±7%). More importantly, the desorption of benzene on Pt/C/W(111) is a reaction-limited process that occurs at 290 K, which is much lower than the benzene desorption temperature of 400 K from Pt(111).  相似文献   

17.
The adsorption and thermal reactivity of NO and CO and the kinetics of the NO reduction with CO on Ir surfaces were studied using X-ray photoelectron spectroscopy, polarization modulation infrared reflection–absorption spectroscopy, and temperature programmed desorption. The NO adsorption and dissociation activity was strongly dependent on the Ir surface structure. The NO dissociation activity of the Ir planes decreased in the order (100) > (211) ? (111). In contrast, the type of the CO adsorption site was independent of the Ir surface structure. The activity of Ir(111) for N2 and CO2 production from the NO + CO reaction was low compared with the activities of Ir(100) and Ir(211). The kinetic data for an Ir/SiO2 powder catalyst were similar to data obtained for Ir(211). The order of the turnover frequencies for N2 and CO2 formation for the Ir planes was in good agreement with the order for NO dissociation activity, and this agreement indicates that the catalytic activity for NO reduction was dependent on NO dissociation. A kinetic study of the elementary steps indicated that the rate-limiting step for NO reduction with CO was the NO dissociation step.  相似文献   

18.
A key intermediate of the methanation reaction on nickel catalyst is a carbidic carbon. Accumulated carbidic intermediates on Ni(100) gives a p(2×2) p4g structure, whereas that on Ni(111) is too complex to be solved. A single domain carbide layer accidentally on Ni(111) allowed us to solve the structure explicitly. Comparison of the carbide layer on Ni(100) and that on Ni(111) showed that the carbon atoms are arranged by forming the same ordered structure. The carbide layers prepared on Ni(100), Ni(111) and Ni(110) have almost equal decomposition temperatures. Consequently, we can conclude that the same overlayer compound is formed on the three surfaces. Furthermore, the hydrogenation of the p4g carbide on Ni(100) occurs at almost equal rate to the turnover frequency of the catalytic methanation reaction. The structure insensitive methanation reaction on Ni(100), Ni(111) and Ni(110) is responsible for the same intermediate compound on these surfaces.The same strategy was applied to the reaction of NO with H2 on Pd(100), Rh(100) and Pt-Rh(100) surfaces, and ac(2×2)-N overlayer was isolated on each of these. The hydrogenation of the isolatedc(2×2)-N produced predominantly NH species, which indicates slow NH species hydrogenation.  相似文献   

19.
Ultra-high temperature ceramic nanocomposites (UHTC-NC) within the Si(HfxTa1?x)(C)N system were synthesized via the polymer-derived ceramics (PDC) synthesis route. The microstructure evolution of the materials was investigated upon pyrolysis and subsequent heat treatment. The crystallization behavior and phase composition were studied utilizing X-ray diffraction, scanning- and transmission electron microscopy. Single-source-precursors were converted into amorphous single-phase ceramics, with the exception of surface crystallization effects, at 1000 °C in NH3. Annealing in N2 at 1600 °C resulted in fully crystalline UHTCs. The powder samples revealed microstructures consisting of two characteristic regions, bulk and surface; displaying intrinsic microstructure and phase composition differences. Instead of the expected nitrides, transition metal carbides (TMC) were detected upon high-temperature anneal. The residual carbon available in the system triggered a decomposition reaction, resulting in the formation of TMCs plus gaseous nitrogen and SiC. Experimental data underline that N-containing PDCs are prone to phase separation accompanied by thermal decomposition and diffusion-controlled coarsening.  相似文献   

20.
The decomposition of acetic acid was studied on a clean Ni/Cu(110) alloy single crystal by means of thermal desorption spectroscopy. The primary alloy surface composition employed in this work was 37% Ni and 63% Cu as measured by Auger electron spectroscopy. Acetate was the predominant surface intermediate observed, giving rise to the CO2 and H2 decomposition products observed, in analogous fashion to the decomposition of formic acid. Surface carbon residue was also detected and could be driven into the bulk by annealing the sample above 900 °K. The rate constant for the decomposition of the acetate intermediate on the alloy was found to be 1013.5 exp(?33(kcal/mol)/RT)sec?1. The autocatalytic decomposition of both carboxylic acids previously observed on a clean Ni(110) surface was totally suppressed. The product distribution from acetic acid observed on both the 37% Ni/65% Cu alloy and the carburized Ni(110) surfaces were very similar, indicating chemical similarities between these two surfaces. At high coverages of the acetate intermediate the activation energy for CO2 formation increased by 14 kcal/gmol. This effect was attributed to strong attractive interactions in the adsorbate layer.  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