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 620 毫秒
1.
An extensive series of 30 Cu exchanged zeolites and Cu impregnated silicas and aluminas have been tested in their capacities to stabilize the bis(μ-oxo)dicopper core. This core shows a remarkably activity towards methane, as it selectively hydroxylates methane into methanol at the low temperature of 125 °C. UV–vis spectroscopy is an easy approach to detect the presence of this bis(μ-oxo)dicopper core since it is characterized by an intense charge transfer band at 22 700 cm−1. In this way it was found that after calcination, only the Cu exchanged zeolites ZSM-5 and MOR are capable of stabilizing this core. In addition, an optimum in the Si/Al ratio and in the calcination temperature were observed, indicating that this core requires a rather specific coordination environment. For ZSM-5, the optimal Si/Al ratio for bis(μ-oxo) dicopper core formation is between 12 and 30 and the amount of this core increases with increasing copper loading above Cu/Al = 0.2. Calcination in O2 should be done at temperatures higher than 280 °C and lower than 700 °C. After reaction with methane at low temperature (150 °C), it was found that only Cu-ZSM-5 and Cu-MOR yielded methanol, whereas all the other Cu based materials yielded almost no methanol. At higher temperatures (200 °C) however, Cu-FER and Cu-BEA showed comparable methanol yields as Cu-ZSM-5 and also the methanol yield of Cu-MOR increased at this higher reaction temperature, indicating that a second not yet identified Cu-oxygen species is activated in the FER, BEA and MOR zeolites at higher temperatures.  相似文献   

2.
Rhodium catalysts, supported on six γ-Al2O3 supports with different crystallinities, were exposed to sequential treatments in hydrogen at 500°C, in oxygen at 760°C, in hydrogen at 500°C and at 760°C, respectively. Samples were characterized by X-ray diffraction and hydrogen chemisorption at various stages in the sequential treatment. Based on the characterization results, it is concluded that the formation of crystalline Rh2O3 is a function of γ-Al2O3 crystallinity; formation of crystalline Rh2O3 increased with increasing crystallinity of γ-Al2O3 during treatment in oxygen at 760°C. The crystalline Rh2O3 formed during treatment in oxygen at 760°C was reduced to Rh metal by hydrogen at 500°C, but most of the Rh did not adsorb hydrogen at room temperature. Subsequent treatment in hydrogen at 760°C increased the hydrogen adsorption capacity by as much as a factor of three. X-ray line broadening measurements showed that oxygen treatment of reduced Rh/γ-Al2O3 at 760°C followed by hydrogen reduction at 500°C resulted in significant increases in Rh crystal size; further treatment in hydrogen at 760°C resulted in additional sintering of Rh.  相似文献   

3.
The possibility of producing ceramic powders suitable for the fabrication of microporous filters was investigated through the thermal treatment of the powder mixtures of a high-purity (99.09% SiO2) quartz and clinoptilolite type of natural zeolite. The quartz and zeolite, mixed in the ratio of 3:1 by weight, was wet ground in a ball mill, the powder was sieved on a 90-μm screen, and the undersize was dried and sintered in the powder form at the temperatures of 1000, 1100 and 1200 °C for 7 h in an air furnace. The powder sinter products were deagglomerated by gentle crushing in an agate mortar and then characterized by phase composition, density, and specific surface area measurements. The added zeolite facilitated the transformation of quartz to cristobalite. The phase transformation of quartz to cristobalite first appeared at around 1000 °C, and, at 1200 °C, led to a ceramic powder sufficiently high in cristobalite content for the fabrication of the microporous ceramic bodies. Re-sintering at 1200 °C of the pressed forms of the ceramic powder resulted in microporous (0.5–3 μm) ceramics with a high porosity of 48.5%, and a three-point bend strength of 140 kg/cm2. The ceramics obtained may have potential for filter applications.  相似文献   

4.
Palladium cation exchanged zeolites (ZSM-5, mordenite and ferrierite) were studied as catalysts for methane combustion. Pd-zeolites showed much higher activities than PdO/Al2O3. For comparable palladium loadings, PdO/Al2O3 requires a reaction temperature of ca. 70–80°C higher than Pd-ZSM-5 for conversions between 50–100%. The catalytic activity of Pd-ZSM-5 seems to be related to its reducibility. Temperature-programmed reduction experiments with carbon monoxide showed a lower reduction temperature (ca. 157°C) for Pd-ZSM-5 than for PdO/Al2O3 (225°C). Further, the positioning of the palladium by ion exchange offers a highly dispersed form of PdII supported on the high surface area zeolite.  相似文献   

5.
Fully dense composites of 0–30 wt% discrete TiN particles distributed in a ß-sialon matrix of overall composition Si5·5Al0·5O0·5N7·5 have been prepared by hot isostatic pressing at 1650 and 1750°C. Pressureless sintering at 1775°C gave materials with an open porosity. Typical sizes of the TiN particles were 1–3 μm, and no intergranular glassy phase was observed in the prepared materials. The grain size of ß-sialon was below 1 μm in the materials HIPed at 1650°C, and 1–2 μm at 1750°C. The Vickers hardness was fairly constant for the TiN-ß-sialon composites with up to 15 wt% TiN added: Hv10 around 17·5 GPa for materials HIPed at 1650° and around 17 GPa at 1750°C, whereas at higher TiN contents the hardness decreased to around 16 GPa. The indentation fracture toughness of the ß-sialon ceramic increased approximatively from 3 to 4 MPam1/2 at an addition of 15 wt% TiN particulates. The fracture toughness could be further increased to 5 MPam1/2 by addition of small amounts of Y2O3 and A1N to a ß-sialon composite with 30 wt% TiN.  相似文献   

6.
Initial tests have been carried out with the fuel cell arrangement La0.6Sr0.4Co0.2Fe0.8O3Ce0.9Gd0.1O1.95Ni/YSZ, incorporating dense film (5–10 μm) Ce0.9Gd0.1O1.95 electrolyte tape cast onto the supporting anode, to investigate the feasibility of intermediate temperature operation (500–700°C). A good open circuit voltage of approx. 0.8 V was obtained at 550°C using moist hydrogen as the fuel. Slightly lower open circuit voltages were found at higher temperatures, which may have been caused by minor gas leakage and the electronic conductivity of the electrolyte. Power outputs in excess of 100 mW/cm2 were obtained at 650°C, and the cell resistance was 0.8Ω cm2 at this temperature. This resistance, and the greater resistance at lower temperature, was predominantly due to the cathode according to AC impedance measurements. Experiments were also carried out at 600°C using direct methanol fuels at the anode; the maximum power output was approximately half of that obtained with hydrogen.  相似文献   

7.
A titanosilicate, named EMS-3, has been synthesized using the organic base TMAOH (tetramethylammonium hydroxide) and Na ion, as structure directing agents, under hydrothermal conditions. The XRD pattern of EMS-3 contains sharp and broad reflections typical of a partially disordered structure. EMS-3 has low thermal stability, indeed XRD pattern shows a decreased resolution after thermal treatment at 300 °C. An amorphous phase is observed at 500 °C. EMS-3 has a high cationic exchange capacity. A small amount of TMA cation has been detected in all the exchanged forms of EMS-3, as showed by IR spectroscopy and thermogravimetric analyses. Cation exchanged EMS-3’s show higher thermal stability than as-synthesized EMS-3. In particular, the typical features of EMS-3 are maintained, although Sr-EMS-3 exhibits poorly resolved XRD pattern after thermal treatment at 500 °C. Divalent cation exchanged EMS-3’s have higher specific surface area than as-synthesized EMS-3, due to a lower amount of extra-framework atoms inside the porous structure. CO2 adsorption capacity is higher than both CH4 and N2. Both the charge and the ionic radius of the exchanged cations affect the adsorption properties. Ca-EMS-3 has the highest adsorption capacity for all the adsorbates.  相似文献   

8.
A critical replication and re-evaluation of Charnell’s procedure to the synthesis of zeolites A and X has been carried out, aiming at reliable protocols for preparation of large and uniform crystals of the respective zeolites in a scale of 50 g per batch. Triethanolamine, as an organic additive to the reacting sodium aluminosilicate hydrogel, increases the viscosity of the system, and reduces the reactivity of aluminium species towards nucleation and crystal growth through forming chelated complex compounds. The reactivity of silicate species has to be accordingly adjusted, by choosing proper source materials. Zeolite A crystals, which possess shapes of edge-truncated cubes, and sizes of 35–40 μm in edge-length, have been synthesized with a starting gel having the composition 1.7Na2O:Al2O3:0.7SiO2:165H2O:6.1TEA in 2-l batch-size, using dissolved metallic aluminium and colloidal silica. The crystallization has been accomplished at 85 °C within 21 days. When the gel has a starting composition 2.3Na2O:Al2O3:1.3SiO2:300H2O:10TEA (1-l batch), 70–80 μm large zeolite X crystals can be obtained at 85 °C in 35 days. Both starting gels are prepared at 0 °C, then quickly heated to the crystallization temperature.  相似文献   

9.
A method to quantify DRIFT spectral features associated with the in situ adsorption of gases on a NOx adsorber catalyst, Pt/K/Al2O3, is described. To implement this method, the multicomponent catalyst is analysed with DRIFT and chemisorption to determine that under operating conditions the surface comprised a Pt phase, a pure γ-Al2O3 phase with associated hydroxyl groups at the surface, and an alkalized-Al2O3 phase where the surface –OH groups are replaced by –OK groups. Both DRIFTS and chemisorption experiments show that 93–97% of the potassium exists in this form. The phases have a fractional surface area of 1.1% for the 1.7 nm-sized Pt, 34% for pure Al2O3 and 65% for the alkalized-Al2O3. NO2 and CO2 chemisorption at 250 °C is implemented to determine the saturation uptake value, which is observed with DRIFTS at 250 °C. Pt/Al2O3 adsorbs 0.087 μmol CO2/m2and 2.0 μmol NO2/m2, and Pt/K/Al2O3 adsorbs 2.0 μmol CO2/m2and 6.4 μmol NO2/m2. This method can be implemented to quantitatively monitor the formation of carboxylates and nitrates on Pt/K/Al2O3 during both lean and rich periods of the NOx adsorber catalyst cycle.  相似文献   

10.
Sodium-doped hydroxylapatite powder was synthesized by the wet chemical method. Powder behavior, thermal stability, mechanical strength and biocompatibility were investigated.

The synthesized hydroxylapatite powder consisted of needles 1800 Å in width and 260 Å in length. The particle size, specific area (BET) and Ca/P atomic ratio were 0·1–0·3 μm, 29·9 m2/g and 1·62 respectively. A large amount of absorbed water existed in the powder, and evaporated on heating to 1000°C. Differential thermal analysis showed that no phase transformation occurred during heating to 1250°C. After heating at 1250°C for 1 h, the O---H bond was still found in the synthesized powder, by IR spectrophotometry. The optimum sintering condition was heating at 1200°C for 4 and this resulted in 680 MPa compressive strength, 1-1·3 μm mean grain size and 99% T.D.

The synthesized hydroxylapatite showed no cytotoxicity and had excellent tissue compatibility. This powder possesses a high potential for bone implantation.  相似文献   


11.
Powders of pure and 5% ytterbium substituted strontium cerate (SrCeO3/SrCe0.95Yb0.05O3−δ) were prepared by spray pyrolysis of nitrate salt solutions. The powders were single phase after calcination in nitrogen atmosphere at 1100 °C (SrCeO3) and 1200 °C (SrCe0.95Yb0.05O3−δ). Dense SrCeO3 and SrCe0.95Yb0.05O3−δ materials were obtained by sintering at 1350–1400 °C in air. Heat treatment at 850 and 1000 °C, respectively, was necessary prior to sintering to obtain high density. The dense materials had homogenous microstructures with grain size in the range 6–10 μm for SrCeO3 and 1–2 μm for SrCe0.95Yb0.05O3−δ. The electrical conductivity of SrCe0.95Yb0.05O3−δ was in good agreement with reported data, showing mixed ionic–electronic conduction. The ionic contribution was dominated by protons below 1000 °C and the proton conductivity reached a maximum of 0.005 S/cm above 900 °C. In oxidizing atmosphere the p-type electronic conduction was dominating above 700 °C, while the contribution from n-type electronic conduction only was significant above 1000 °C in reducing atmosphere.  相似文献   

12.
In this study, cylindrical porous glass tablets were plated by palladium using electroless plating technique. Hypophosphite and Co(II) complexes were used as reducing agents in the prepared plating baths. Experiments were carried out in an especially designed glass vessel in which helium gas was continuously bubbled through the solution to create uniform concentration and to remove hydrogen gas from the surface for the case of hypophosphite-based procedure. XRF analysis of the upper layer of the composite membrane prepared by the hypophosphite-based bath showed a Pd/Si ratio of 4.6. SEM photographs indicated impregnation of Pd into the substrate upto 200 μm. However, the thickness of the dense Pd layer was only about 15 μm. SEM photographs and XRF results showed that hypophosphite-based bath was much more successful than the Co(II) complex-bath in Pd plating. Permeation experiments carried out at different temperatures showed that the contribution of surface diffusion to the permeation was significant at low temperatures and solution–diffusion mechanism was not important in the 40–200 °C temperature range for these membranes. The selectivity ratio for H2/N2 was found to be about 7 at 200 °C.  相似文献   

13.
The present work has provided new fundamental information on the effects of aging (mileage, km) of a commercial Pd-Rh (9:1, w/w) three-way catalyst (TWC) on (a) the H2 chemisorption, (b) the redox properties of washcoat material, and (c) the dynamic oxygen storage and release properties of TWC. Hydrogen chemisorption in the 25–400 °C range was found to decrease with increasing aging of TWC. On the other hand, H2 chemisorption performed at 200 and 400 °C resulted in a significantly larger amount of chemisorbed hydrogen (exceeding the monolayer value based on the noble metals) compared to that obtained after chemisorption at 25 °C. This is due to the onset of a hydrogen spill over process in the 200–400 °C range. The features of H2-TPD spectra were found to strongly depend on the aging of TWC. Hydrogen chemisorption at 25 °C followed by TPD after a given pre-treatment of the catalyst surface might be considered as a procedure for a good estimate of the metal dispersion of a commercial Pd-Rh TWC. The redox properties of the oxygen storage components of the washcoat of a commercial Pd-Rh TWC were found to drastically change with increasing aging of TWC. It was found that the oxygen storage capacity (OSC) of the TWC investigated decreases significantly with catalyst aging in the 0–56,000 km range and in the temperature range of 500–750 °C. In the case of fresh TWC, the presence of 20 ppm SO2 or 10% CO2 in a 1.5% O2/He gas mixture used for oxygen storage (oxidizing gas) was found to result in a large decrease in the amount of dynamic OSC measured by alternating switches between oxidizing and reducing feed gas compositions. The shift in time of the peak maximum of the transient response of CO2 obtained during the switch O2/He → He → CO/He (t) can be ascribed to the alteration of the kinetics of the oxygen back-spillover process and not to the kinetics of CO oxidation on the metal surface.  相似文献   

14.
The catalytic effect of a heteropolyacid, H4SiW12O40, on nitrobenzene (20 and 30 μM) oxidation in supercritical water was investigated. A capillary flow-through reactor was operated at varying temperatures (T=400–500 °C; P=30.7 MPa) and H4SiW12O40 concentrations (3.5–34.8 μM) in an attempt to establish global power-law rate expressions for homogenous H4SiW12O40-catalyzed and uncatalyzed supercritical water oxidation. Oxidation pathways and reaction mechanisms were further examined via primary oxidation product identification and the addition of various hydroxyl radical scavengers (2-propanol, acetone, acetone-d6, bromide and iodide) to the reaction medium. Under our experimental conditions, nitrobenzene degradation rates were significantly enhanced in the presence of H4SiW12O40. The major differences in temperature dependence observed between catalyzed and uncatalyzed nitrobenzene oxidation kinetics strongly suggest that the reaction path of H4SiW12O40-catalyzed supercritical water oxidation (average activation Ea=218 kJ/mol; k=0.015–0.806 s−1 energy for T=440–500 °C; Ea=134 kJ/mol for the temperature range T=470–490 °C) apparently differs from that of uncatalyzed supercritical water oxidation (Ea=212 kJ/mol; k=0.37–6.6 μM s−1). Similar primary oxidation products (i.e. phenol and 2-, 3-, and 4-nitrophenol) were identified for both treatment systems. H4SiW12O40-catalyzed homogenous nitrobenzene oxidation kinetics was not sensitive to the presence of OH√ scavengers.  相似文献   

15.
The nitridation of elemental silicon powder at 900–1475 °C was studied by X-ray photoelectron spectroscopy (XPS), X-ray excited Auger electron spectroscopy (XAES), XRD, thermal analysis and 29Si MAS NMR. An initial mass gain of about 12% at 1250–1300 °C corresponds to the formation of a product layer about 0·2 μm thick (assuming spherical particles). XPS and XAES show that in this temperature range, the surface atomic ratio of N/Si increases and the ratio O/Si decreases as the surface layer is converted to Si2N2O. XRD shows that above 1300 °C the Si is rapidly converted to a mixture of - and β-Si3N4, the latter predominating >1400 °C. In this temperature range there are only slight changes in the composition of the surface material, which at the higher temperatures regains a small amount of an oxidised surface layer. By contrast, in the interval 1400–1475 °C, the 29Si MAS NMR chemical shift of the elemental Si changes progressively from about −80 ppm to −70 ppm, in tandem with the growth of the Si3N4 resonance at about −48 ppm. Possible reasons for this previously unreported change in the Si chemical shift are discussed. ©  相似文献   

16.
The effect of microwave irradiation on zeolite Na-A has been studied in comparison with classical heat treatment. The thermal behavior and the structural transformations of the zeolite were investigated by X-ray powder diffraction. The structure of zeolite Na-A changed concurrently with microwave treatment. The longer the irradiation period, the larger was the amount of zeolite transformed into another crystalline phase (low-carnegieite), while a decreasing zeolite content remained detectable. On the other hand, during classical heat treatment at 400–800°C, zeolite Na-A went through subsequent structural changes; at about 800°C nepheline (NaAlSiO4) crystallized as a stable phase along with a considerable amount of amorphous material.  相似文献   

17.
The single gas H2 and N2 permeability of a 4 μm thick dense fcc-Pd66Cu34 layer has been studied between room temperature and 510 °C and at pressure differences up to 400 kPa. Above 50 °C the H2 flux exhibits an Arrhenius-type temperature dependence with JH2=(5.2±0.3) mol m−2 s−1 exp[(−21.3 ± 0.2) kJ mol−1/(R·T)]. The hydrogen transport rate is controlled by the bulk diffusion although the pressure dependence of the H2 flux deviates slightly from Sieverts’ law. A sudden increase of the H2 flux below 50 °C is attributed to embrittlement.  相似文献   

18.
Complete Ni2+ exchange of a single crystal of zeolite X of composition Na92Si100Al92O384 per unit cell was attempted at 73°C with flowing aqueous 0.05 M NiCl2 (pH=4.3 at 23°C). After partial dehydration at 23°C and ≈10−3 Torr for two days, its structure, now of composition Ni2(NiOH)35(Ni4AlO4)2(H3O)46Si101Al91O384 per unit cell, was determined by X-ray diffraction techniques at 23°C (space group Fd , a0=24.788(5) Å). It was refined using all intensities; R1=0.080 for the 236 reflections for which Fo>4σ(Fo), and wR2=0.187 using all 1138 unique reflections measured. At four crystallographic sites, 45 Ni2+ ions were found per unit cell. Thirty of these are at two different site III′ positions. Twenty of those are close to the sides of 12-rings near O–Si–O sequences, where each coordinates octahedrally to two framework oxygens, to three water molecules which hydrogen bond to the zeolite framework, and to an OH ion. The remaining 10 are near O–Al–O sequences; only three members of a likely octahedral coordination sphere could be found. In addition, two Ni2+ ions are at site I, eight are at site I′, and five are at site II. Forty six H3O+ ions per unit cell, 24 at site II′ and 22 at site II, each hydrogen bond triply to six rings of the zeolite framework. Each of the 22 H3O+ ions also hydrogen bonds to a H2O molecule that coordinates to a site III′ Ni2+ ion. Six of the eight sodalite cages each contain four H3O+ ions at site II′; the remaining two each contains a tetrahedral orthoaluminate anion at its center. Each tetrahedral face of each orthoaluminate ion is centered by a site I′ Ni2+ ion to give two Ni4AlO4 clusters. The five site II Ni2+ ions each coordinate to a OH ion. With 46 H3O+ ions per unit cell, the great tendency of hydrated Ni2+ to hydrolyze within zeolite X is demonstrated. With a relatively weak single-crystal diffraction pattern, with dealumination of the zeolite framework, and with an apparent decrease in long-range Si/Al ordering likely due to the formation of antidomains, this crystal like others treated with hydrolyzing cations appears to have been damaged by Ni2+ exchange and partial dehydration.  相似文献   

19.
The structure and density of individual ammonium chloride particles formed at 0 and −20°C by homogeneous nucleation were studied using electron microscopy and X-ray diffraction. The crystal size apparently increased at the lower temperature and many of the particles formed at −20°C were single crystals or had an oriented polycrystalline structure. These results differ from those reported previously for particles formed at room temperature (23–26°C), which showed an amorphous or randomly-oriented fine crystal structure. Coagulation was more frequently observed as the temperature decreased and the porosity present in the particles appeared to be much finer and more uniform. The density of these particles decreased from about 0.26 g cm−3 for particles of size 0.1–0.2 μm to approximately 0.1 g cm−3 for particles slightly smaller than 1 μ.  相似文献   

20.
The process for alumina gel casting was developed using an inorganic binder The monohydroxy aluminium oxide (boehmite, AlOOH) was incorporated with ultrafine alumina of particles (0·5 μm) and the slurry rheology was studied and presented. The effect of boehmite in slurry viscosity was observed with respect to different amounts of boehmite and time. The alumina 54 vol% slurry with 10 wt% boehmite showed the viscosity of 880 mPa s at 93 s−1. An external coagulating agent, HMTA, was incorporated with alumina–boehmite slurry and the effective change in slurry viscosity with respect to concentration and time was studied. The addition of HMTA results in faster gelation and the optimum concentration was determined as 0·21 mol L−1. The alumina gelcast body was dried under humidity conditions at 40°C, RH 70%. The defect free dried green body was obtained and the total linear drying shrinkage was calculated as 3·2% and the green density observed was 59·3% of theoretical value. The sintered density of 98% (TD) was achieved at 1450°C in 2 h. The mechanical hardness of sintered alumina measured as 2286 kg mm−2. The sintered ceramic showed an extremely fine grained microstructure with an average grain size <2 μm. The boehmite acts as an excellent binder and sintering aid for alumina ceramics.  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