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 15 毫秒
1.
The rate of ZnAl2O4 formation was measured for η-, γ-; and α- Al2O3 in order to distinguish the reactivity of them. The reactivity decreased as follows: η- > γ- > α-Al2O3. The reaction rate fitted to Jander's equation and the activation energies calculated were 33, 47 and 113 Kcal/mol for η-, γ- and α-Al2O3 systems, respectively. These differences are explained by an assumption that η- and γ-Al2O3 resulted in a ZnAl2O4 with imperfect spinel structure, but α-Al2O3 gave the perfect spinel structure. This assumption is based on the theoretical consideration of the activation energy needed for the diffusion-controlled reaction and date of lattice constant of each ZnAl2O4 obtained from three aluminas. The fact that η-Al2O3 shows very high reactivity compared with that of γ-Al2O3 was found to be explained on the basis of Jander's equation, a comparison of specific surface area and the defect structures of the aluminas.  相似文献   

2.
High-temperature treatment of γ-Al2O3 can lead to a series of polymorphic transformations, including the formation of δ-Al2O3 and θ-Al2O3. Quantification of the microstructure in the range where δ- and θ-Al2O3 are formed represents a formidable challenge, as both phases accommodate a high degree of structural disorder. In this work, we explore the use of an XRD recursive-stacking formalism for the quantification of high-temperature transition aluminas. We formulate the recursive-stacking methodology for modelling of disorder in δ-Al2O3 and twinning in θ-Al2O3 and show that explicitly accounting for the disorder is necessary to reliably model the XRD patterns of high-temperature transition alumina. We also use the recursive stacking approach to study phase transformation during high-temperature (1050 °C) treatment. We show that the two different intergrowth modes of δ-Al2O3 have different transformation characteristics and that a significant portion of δ-Al2O3 is stabilized with θ-Al2O3 even after prolonged high-temperature exposures.  相似文献   

3.
The nature and stability of surface species of CuCl2 supported on α-Al2O3, γ-Al2O3, and SiO2 were investigated by using X-ray diffraction techniques and reflectance spectroscopy. No specific chemical interaction of CuCl2 is observed on an inert α-Al2O3 support, as opposed to hydrated carriers as SiO2 and γ-Al2O3. On these supports the coordination sphere of Cu2+ consists of surface groups (OH? or O? at drying and activation, resp.), H2O and Cl?, with the H2O ligands decreasing in concentration in the process of impregnation, drying and calcination. γ-Al2O3 samples, calcined at 400°C, show γ-Cu2(OH)3Cl as opposed to CuAl2O4 at higher temperatures. The absence of Cu2(OH)3Cl on SiO2-supported samples is related to the acid-base characteristics of the carriers. The various supports can be arranged in the following order of stability of the complexes formed: γ-Al2O3 > SiO2 ? -Al2O3.  相似文献   

4.
用酸中和法制备了活性γ-Al2O3, 并在其表面负载SO3得到固体酸催化剂SO3/γ-Al2O3, 用XRD, TG-DTA, FT-IR,NMR, NH3-TPD等对其进行了结构和酸性研究. 结果表明: 在SO3/γ-Al2O3的制备过程中形成少量的Al2(SO4)3, 同时SO3与γ-Al2O3表面上的羟基反应, 形成强的Brönsted酸位, 根据1H/27Al 双共振(TRAPDOR)MAS NMR与FT-IR实验结果提出了Brönsted酸结构模型. SO3/γ-Al2O3表面存在两种不同强度的酸中心, 其酸强度大于分子筛HZSM-5, 但弱于传统的固体超强酸 /γ-Al2O3.  相似文献   

5.
Formation of α-Al2O3 in thermal treatment of ultradispersed aluminum oxide powders produced by shock-wave synthesis with addition of Cr2O3-(NH4)2Cr2O7 as a precursor was studied. The acceleration of the conversion of δ-and θ-Al2O3 to α-Al2O3 upon introduction of additives in the stage of synthesis of aluminum oxide was examined.  相似文献   

6.
Using a new experimental technique, “Continuous Elution Method”, the desorption behavior of polystyrene(PS) and polystyrene (PS-X) functionalized by a terminal iminium ion (-X) from α-Al2O3 and γ-Al2O3 surface were investigated, and found that PS-X is forming a terminally adsorbed polymer layer on α-Al2O3, surface. Furthermore, it was found that the adsorption force of terminally adsorbed polymer is balanced with the desorption force which is contributed from the osmotic pressure in the adsorption layer. Based on this concept, the adsorption energy of the end-functionalized polystyrene terminally adsorbed on the α-Al2O3, surface was evaluated to be 4.2 ˜4.3 kT.  相似文献   

7.
We prepared Pd catalysts supported on various metal oxides, viz. γ-Al2O3, α-Al2O3, SiO2–Al2O3, SiO2, CeO2 and TiO2 by an incipient wetness method and applied them to propane combustion. Several techniques: N2 physisorption, inductively coupled plasma-atomic emission spectroscopy (ICP-AES), CO chemisorption, temperature-programmed reduction (TPR) and temperature-programmed oxidation (TPO) were employed to characterize the catalysts. Pd/SiO2–Al2O3 showed the least catalytic activity at high temperatures among Pd catalysts supported on irreducible metal oxides, viz. SiO2, Al2O3 and SiO2–Al2O3. Pd/γ-Al2O3 was much superior for this reaction to Pd/α-Al2O3. The Pd catalyst supported on reducible metal oxides (CeO2 and TiO2) with a less specific surface area showed the higher catalytic activity compared with that supported on reducible metal oxides with a higher specific surface area, even though the former had a less Pd dispersion than the latter. In the case of Pd/SiO2–Al2O3, the initially reduced Pd catalyst was superior to the fully oxidized one. The oxidation of metallic Pd occurred in the presence of O2 with increasing reaction temperature, which resulted in the change in the catalytic activity.  相似文献   

8.
Olefin oligomerization by γ-Al2O3 has recently been reported, and it was suggested that Lewis acid sites are catalytic. The goal of this study is to determine the number of active sites per gram of alumina to confirm that Lewis acid sites are indeed catalytic. Addition of an inorganic Sr oxide base resulted in a linear decrease in the propylene oligomerization conversion at loadings up to 0.3 wt %; while, there is a >95 % loss in conversion above 1 wt % Sr. Additionally, there was a linear decrease in the intensity of the Lewis acid peaks of absorbed pyridine in the IR spectra with an increase in Sr loading, which correlates with the loss in propylene conversion, suggesting that Lewis acid sites are catalytic. Characterization of the Sr structure by XAS and STEM indicates that single Sr2+ ions are bound to the γ-Al2O3 surface and poison one catalytic site per Sr ion. The maximum loading needed to poison all catalytic sites, assuming uniform surface coverage, was ∼0.4 wt % Sr, giving an acid site density of ∼0.2 sites per nm2 of γ-Al2O3, or approximately 3 % of the alumina surface.  相似文献   

9.
The surface properties of supported gallium oxide catalysts prepared by impregnation of various supports (γ-Al2O3, SiO2, TiO2, ZrO2) were investigated by adsorption microcalorimetry, using ammonia and water as probe molecules. In the case of acidic supports (γ-Al2O3, ZrO2, TiO2), the acidic character of supported gallium catalysts always decreased in comparison with gallium-free supports; on very weakly acidic SiO2, new acidic centers were created when depositing Ga2O3. The addition of gallium oxide decreased the hydrophilic properties of alumina, titania and zirconia, but increased the amount of water adsorbed on silica. The catalytic performances in the selective catalytic reduction of NO by C2H4 in excess oxygenwere in the order Ga/Al2O3>Ga/TiO2>Ga/ZrO2>>Ga/SiO2. This order is more related to the quality of the dispersion of Ga2O3 on the support than to the global acidity of the solids. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

10.
γ-Al2O3, dried alumina gel, as well as their phosphated forms using (NH4)2HPO4 were prepared by wet impregnation and calcined at 870 K. Resulted samples were subjected to investigate the consequent bulk [X-ray powder diffractometry (XRD), diffuse reflectance spectroscopy (DRS) and infrared spectroscopy (IR)] and surface (texture, N2-adsorption and surface acid properties, pyridine adsorption). Results indicated no detectable bulk phase changes due to phosphation. However, the phosphated gel sample reveals the highest SBET. Surface stabilization of phosphate species by γ-Al2O3 or gel is indicated, leading to modifications on surface hydroxyl and hence surface acidity. The phosphated gel sample exhibits the strongest acidity (both Bronsted and Lewis).  相似文献   

11.
The effective utilization of various biomolecules for creating a series of mesoporous boehmite (γ-AlOOH) and gamma-alumina (γ-Al2O3) nanosheets with unique hierarchical multilayered structures is demonstrated. The nature and concentration of the biomolecules strongly influence the degree of the crystallinity, the morphology, and the textural properties of the resulting γ-AlOOH and γ-Al2O3 nanosheets, allowing for easy tuning. The hierarchical γ-AlOOH and γ-Al2O3 multilayered nanosheets synthesized by using biomolecules exhibit enhanced crystallinity, improved particle separation, and well-defined multilayered structures compared to those obtained without biomolecules. More impressively, these γ-AlOOH and γ-Al2O3 nanosheets possess high surface areas up to 425 and 371 m2 g−1, respectively, due to their mesoporous nature and hierarchical multilayered structure. When employed for molybdenum adsorption toward medical radioisotope production, the hierarchical γ-Al2O3 multilayered nanosheets exhibit Mo adsorption capacities of 33.1–40.8 mg g−1. The Mo adsorption performance of these materials is influenced by the synergistic combination of the crystallinity, the surface area, and the pore volume. It is expected that the proposed biomolecule-assisted strategy may be expanded for the creation of other 3D mesoporous oxides in the future.  相似文献   

12.
The formation of Pd–Ag nanoparticles deposited from the heterobimetallic acetate complex PdAg2(OAc)4(HOAc)4 on α-Al2O3, γ-Al2O3, and MgAl2O4 has been investigated by high-resolution trans-mission electron microscopy, temperature-programmed reduction, and IR spectroscopy of adsorbed CO. The reduction of PdAg2(OAc)4(HOAc)4 supported on γ-Al2O3 and MgAl2O4 takes place in two steps (at 15–245 and 290–550°C) and yields Pd–Ag particles whose average size is 6–7 nm. The reduction of the Pd–Ag catalyst supported on α-Al2O3 occurs in a much narrower temperature range (15–200°C) and yields larger nanoparticles (~10–20 nm). The formation of Pd–Ag alloy nanoparticles in all of the samples is demonstrated by IR spectroscopy of adsorbed CO, which indicates a marked weakening of the absorption band of the bridged form of adsorbed carbon monoxide and a >30-cm–1 bathochromic shift of the linear adsorbed CO band. IR spectroscopic data for PdAg2/α-Al2O3 suggest that Pd in this sample occurs as isolated atoms on the surface of bimetallic nanoparticles, as is indicated by the almost complete absence of bridged adsorbed CO bands and by a significant weakening of the Pd–CO bond relative to the same bond in the bimetallic samples based on γ-Al2O3 and MgAl2O4 and in the monometallic reference sample Pd/γ-Al2O3.  相似文献   

13.
The catalytic activity of alumina-manganese catalysts in the oxidation of CO was studied. The MnO x -Al2O3 catalysts were prepared by an extrusion method with the introduction of mechanically activated components (manganese oxide and its mixtures with aluminum oxide, aluminum hydroxide, and a mixture of a manganese salt with aluminum hydroxide) into a paste of aluminum hydroxide followed by thermal treatment in air or argon at 1000°C. In the majority of cases, the catalysts contained a mixture of the phases of β-Mn3O4 (Mn2O3), α-Al2O3, and δ-Al2O3. The presence of low-temperature δ-Al2O3 suggested the incomplete interaction of manganese and aluminum oxides. It was found that the catalytic activity of MnO x -Al2O3 depends on the degree of interaction of the initial reactants, and its value is correlated with the amount of β-Mn3O4 in the active constituent. The intermediate thermal treatment of components at 700°C negatively affects the catalytic activity as a result of the formation of Mn2O3 and the coarsening of particles, which levels the results of mechanochemical activation. The greatest degree of interaction between Al- and Mn-containing components was reached in the selection of mechanochemical activation conditions by decreasing the size of grinding bodies, optimizing the time of mechanochemical activation, and using the mechanochemical activation of precursor mixtures. As a result of mechanochemical activation, the initial reactants were dispersed, the amounts of MnO2 and Mn2O3 changed, and defects were formed; this strengthened the interaction of components and increased catalytic activity.  相似文献   

14.
Investigations on Metal Catalysts. XXII. Chemisorption of Hydrogen on Pt? Ru and η-Al2O3 Supported Pt? Ru Catalysts Measurements of the H2 chemisorption by volumetric method were carried out on several series of Pt? Ru and η-Al2O3 supported Pt? Ru catalysts. Addition of Ru to Pt and vice versa effects a remarkable influence on the sorption behaviour of the starting samples. For mixtures of carrier-free catalysts and η-Al2O3 as well as Pt? Ru/η-Al2O3 catalysts a hydrogen-spillover effect was found.  相似文献   

15.
Investigations on Metal Catalysts. XI. Investigations on Pt? η-Al2O3 Catalysts Modifieded by Iron, Cobalt, and Nickel Pt? Me? η-Al2O3 catalysts (M: Fe, Co, Ni) were characterized by magnetic investigations, reflectance spectra and determination of dispersity (chemisorption of CO, oxygen-hydrogen titration), respectively. The phase structure of platinum-rich catalysts is composed of a high degree by Pt3Fe super-structure. All the Pt? Fe? η-Al2O3 catalysts contained FeIII ions in octahedral symmetry. The dispersity of the metallic components is determined essentially by their phase structure.  相似文献   

16.
Catalysts of Nb2O5/γ-Al2O3 were prepared by aqueous solution impregnation. The state of niobia species on surface of γ-Al2O3 is characterized by using the technology of X-ray power diffraction (XRD) and analyzed using the “incorporation model”. The acidity and the nature of acid sites of the catalysts were evaluated by means of Fourier transform infrared (FTIR) spectroscopy of adsorbed pyridine. The catalytic activity of Nb2O5/γ-Al2O3 catalysts was evaluated by a condensation reaction from isobutene and isobutyraldehyde to 2,5-dimethyl-2,4-hexadiene. The results of XRD indicate that the dispersion capacity of niobia on γ-Al2O3 is about 0.76 mmol Nb per 100m2 γ-Al2O3, which is almost identical to the theoretical value (0.75 mmol Nb per 100m2 γ-Al2O3) calculated by the “incorporation model”. The results of Py-IR and catalytic activity evaluation indicate that the acidity feature is related to the state of dispersed niobia species as well as the loading of niobia onto the surface of γ-Al2O3 support.  相似文献   

17.
Investigations on Metal Catalysts. XXVII. On the Influence of Several Carriers on the Dispersion of Nickel Nickel supported catalysts (carrier: η-, ?-, α-Al2O3, TiO2) were characterized by chemically determined degree of reduction, CO chemisorption, magnetic susceptibility measurements and FMR. The influence of interaction between carrier and active component in the unreduced state on the size and number of Ni crystallites is discussed.  相似文献   

18.
The behavior of the manganese-alumina system with Mn:Al = 1:1 on heating in air and vacuum was studied. The starting samples were mixtures of β-Mn3O4, α-Mn2O3, and γ-Al2O3. On heating to 950°C in air, the samples were partially oxidized into α-Mn2O3, and corundum α-Al2O3 formed along with mixed manganese-alumina cubic spinel, whose composition was close to Mn2AlO4. In vacuum at 1200°C, the starting sample with a ratio of Mn:Al = 1:1 transformed into the manganese-alumina spinel Mn1.5Al1.5O4, which retained its cubic structure after slow cooling in vacuum. When cooled in air, this solid solution delaminated, and a nanocrystalline Mn2.8Al0.2O4 phase formed, whose structure was β-Mn3O4 type tetragonal spinel.  相似文献   

19.
Manganese oxides supported on γ-Al2O3, amorphous SiO2, MCM-41, and TiO2 prepared by an impregnation method were used as heterogeneous catalysts for epoxidation of alkenes with 30 % H2O2 in the presence of NaHCO3 aqueous solution. The effect of support and manganese loading on their activity was studied. The 1.3-MnO x /γ-Al2O3 exhibited superior epoxidazing activity of styrene, compared with other supported MnO x . Hydrogen temperature-programmed reduction, UV–vis and ESR analyses suggested that Mn2+ (catalytic activity species) dominated in 1.3 % MnO x /γ-Al2O3 due to a strong interaction between MnO x and γ-Al2O3. Recycling studies showed the catalyst was a heterogeneous one and retained its activity after recycling four times.  相似文献   

20.
Temperature-programmed thermal decomposition of γ- and α-manganese oxyhydroxide has been studied between 20 and 670°C under vacuum and under a low pressure (10 Torr) of oxygen. Solid products at various temperatures have been analyzed by X-ray diffractometry. Under vacuum γ-MnOOH decomposed below 400°C to a mixture of Mn5O8, α-Mn3O4, and water according to the reaction scheme: 8MnOOH → Mn5O8 + Mn3O4 + 4H2O. Above this temperature Mn5O8 was converted to α-Mn3O4 as a result of oxygen removal. The vacuum dehydration at 250°C of oxyhydroxide rich in α-MnOOH led to the formation of a new modification of Mn2O3 isostructural with corundum (α-Al2O3). In oxygen both oxyhydroxides decomposed to β-MnO2. γ-MnOOH transformed directly to β-MnO2 while α-MnOOH appeared to transform via corundum-phase Mn2O3 as an intermediate.  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