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 709 毫秒
1.
The hydration characteristics of calcium monoaluminate were studied using an effective water/aluminate ratio of 0.15 at 20° or 80°C, from a few minutes of two months. The material hydrated at 80°C shows a large shrinkage while at the lower temperature a continuous expansion occurs. The product at 80°C shows a much higher strength than that hydrated at 20°C. The main initial hydration products are 2Ca0, Aℓ203, 8H20 and alumina gel. Microcracks are developed in the products hydrated at 20°C while at the higher temperature a very compact mass results. The data indicate that it is possible to obtain a durable high alumina cement by using a low water/cement ratio and hydrating at higher temperatures, and under these conditions C3AH3---C3AH6 bond is favoured.  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.
In the present study, we have examined sulfation of cerium oxide via impregnation of (NH4)2SO4, followed by heating in the temperature range of 220–720°C, using Raman Spectroscopy. Based on the SO and SO stretching frequencies in the range of 900–1400 cm−1, a wide range of surface oxysulfur species and bulk cerium-oxy-sulfur species are identified. At 220°C, a mixture of (NH4)2SO4 crystals, SO2−4(aq) and HSO1−r(aq) is found to have formed on ceria's surface, whereas complete conversion of (NH4)2SO4 to SO2−4(aq) and HSO1−4(aq) occurs at 280°C. At 350°C, formation of a mixture of surface pyrosulfate S2O2−7(surf.0, consisting of two SO oscillators and a bulk type compound identified as Ce(IV)(SO4)x(SO3)2−x (0 < x < 2) have been observed. Upon introduction of moisture, the former transforms to HSO1−4(surf.), whereas the latter remains unchanged. At 400°C, only the bulk type compound can be observed. At 450°C, only Ce2(SO4)3 is generated and remains stable until 650°C. Further increase in the temperature to 720°C results in the formation of CeOSO4. The present study not only provides a more thorough understanding of the sulfation of cerium oxide at a molecular level, but also demonstrates that Raman spectroscopy is a highly effective technique to characterize sulfation of metal oxides.  相似文献   

4.
The storage and release of NO2 on alumina-supported barium oxide has been studied with particular attention to the stoichiometry of the two processes. At 400 °C the storage process is characterised by a short period of complete uptake, possibly as nitrito or nitro species, followed by a slower partial uptake in which approximately one NO is released for every three NO2 lost. The latter reaction appears to supply the oxygen necessary to store NO2 as nitrate ions. Molecular O2 has little direct involvement even if in large excess. The second storage reaction also occurs, but to a much lesser extent, with Al2O3 alone. During temperature programmed desorption, release of NOx from Al2O3 peaks at 430 °C with evolution of NO2 and some O2. Release from BaO/Al2O3 exhibits an additional peak near 520 °C corresponding to formation of NO and a higher O2 concentration. The NO may arise from NO2 since BaO/Al2O3 has activity for NO2 decomposition by 500 °C. Although CO2 at low concentration is rapidly taken up by BaO/Al2O3 at 400 °C it is displaced by NO2 and does not interfere with storage. Thermodynamic calculations show that the formation of Ba(NO3)2 by the reaction of NO2 with bulk BaCO3 under the conditions used here is more favourable above 380 °C if NO is evolved than if O2 is consumed.  相似文献   

5.
Well crystallised aluminium borate Al18B4O33 has been synthesised from alumina and boric acid with a BET area of 18 m2/g after calcination at 1100 °C. Afterwards, 2 wt.% Pd/Al18B4O33 was prepared by conventional impregnation of Pd(NO3)2 aqueous solution and calcination in air at 500 °C. The catalytic activity of Pd/Al18B4O33 in the complete oxidation of methane was measured between 300 and 900 °C and compared with that of Pd/Al2O3. Pd/Al18B4O33 exhibited a much lower activity than Pd/Al2O3 when treated in hydrogen at 500 °C or aged in O2/H2O (90:10) at 800 °C prior to catalytic testing. Surprisingly, a catalytic reaction run up to 900 °C in the reaction mixture induced a steep increase of the catalytic activity of Pd/Al18B4O33 which became as active as Pd/Al2O3. Moreover, the decrease of the catalytic activity observed around 750 °C for Pd/Al2O3 and attributed to PdO decomposition into metallic Pd was significantly shifted to higher temperatures (820 °C) in the case of Pd/Al18B4O33. The existence of two distinct types of PdO species formed on Al18B4O33 and being, respectively, responsible for the improvement of the activity at low and high temperature was proposed on the basis of diffuse reflectance spectroscopy and temperature-programmed desorption of O2.  相似文献   

6.
The system Al2O3–ZrO2 was studied by differential thermal analysis in inert atmosphere and in vacuum. The eutectic was located at 1866°C and 40% mass of ZrO2. Zirconia solid solution at the eutectic temperature is up to 1.1±0.3% mass of Al2O3. Enthalpy of melting of this eutectic is 1080±90 J/g. Pure ZrO2 transforms from monoclinic to tetragonal at 1162±7°C, but the saturated solid solution of ZrO2, with 0.7±0.2% mass Al2O3 at this temperature, transforms at 1085±5°C. Inverse transitions occur with hysteresis correspondingly at 1055±5 and 995±5°C. Enthalpy of transformation of pure ZrO2 from monoclinic to tetragonal phase is 42±5 J/g (5.2±0.6 J/mol) but only 30±5 J/g for a ZrO2 saturated solid solution.  相似文献   

7.
The influence of sintering and poling conditions on dielectric properties and microstructures of the system 0·125Pb(Mg1/3Nb2/3)O3−0·875Pb (Zr0·5Ti0·5)O3 was investigated. Specimens were prepared by the conventional mixed-oxide technique. On account of eliminating the pyrochlore phase and lowering the sintering temperature, the calcined 0·125PZT−0·875PMN ceramic was doped with 4PbO.B2O3 glass powder. The 4PbO.B2O3 glass frit not only has a low flow temperature, but also a high polarizability. Additions of 4PbO.B2O3 to the perovskite 0·125PMN–0·875PZT solid solution will form a liquid phase, which served as a densification aid for the ceramics. With additions of 0·2 wt% glass frit, densities in excess of 98% of theoretical were obtained after sintering at 115°C. By variation of the fabrication processes, the influence of sintering and poling conditions on the properties of the ceramics was studied.  相似文献   

8.
Low-temperature sintering of PZT ceramics   总被引:3,自引:0,他引:3  
The required sintering temperature of Pb(Zr0·52Ti0·48)O3 ceramics (abbreviated as PZT 52/48) can be lowered to about 1000°C by incorporating Li2CO3, Na2CO3 or Bi2O3. A dielectric constant of about 1000 and a planar coupling factor of between 45% and 65% are obtained in PZT 52/48 ceramics sintered at 1025°C, with added Li2CO3 and Bi2O3. The optimal amount of the additives, which can be deduced from the densification, the dielectric and piezoelectric properties of the sintered PZT 52/48 ceramics, is 0·375 wt% of Li2CO3 together with an equal mole fraction of Bi2O3. A planar coupling factor of 65% is obtained. This is explained, with the aid of X-ray diffraction (XRD) analysis, by a maximum c/a ratio and consequently by a large spontaneous polarization. The PZT 52/48 ceramics sintered with Li2CO3 and Bi2O3 under the optimal conditions can have ε33T of about 1000, kp higher than 60%, Qm around 100 and tan δ less than 2·0%.  相似文献   

9.
The durability of plasma-sprayed metals bonded with a polyimide adhesive has been studied. Metal adherend surfaces were prepared for adhesive bonding by plasma-spraying inorganic powders on aluminum and titanium. The plasma-sprayed materials included Al2O3, AlPO4. MgO, and SiO2 on aluminum, and TiO2, TiSi2, MgO, and SiO2 on titanium. The coatings were sprayed at two different thicknesses. Durability studies of samples prepared in a wedge-type geometry were carried out. Bonded specimens were maintained in an environmental cycle that included exposure to the conditions; low temperature, - 20°C; relative humidity at elevated temperature, 70% RH at 66°C; elevated temperature (160°C) in air, high temperature (160°C) in vacuum (130 torr, 0.2 atm.), and room temperature. Crack growth rate and mode of failure were determined. The results of the durability tests indicate that thin coatings (25 μm) of plasma-sprayed materials perform better than thicker (150 μm) coatings. The crack growth rate for thin coatings (25 μm) of Al2O3, AlPO4, SiO2, and MgO plasma-sprayed on aluminum was equivalent to that for phosphoric acid anodized aluminum. Similarly, the durability performance for titanium samples prepared with a 25 μm-thick TiO2, TiSi2, and SiO2 plasma-sprayed coatings was equivalent to that for a Turco®-prepared titanium surface. Although the evaluation of durability as a function of surface chemistry was an objective of the study, it was not possible to evaluate the effect, since most failures occurred within the adhesive (cohesive failure) during the environmental tests. That failure occurred in the adhesive indicates that the coating-adherend and the coating-adhesive interactions are sufficiently robust to prevent interfacial failure under the experimental conditions investigated.  相似文献   

10.
The thermal decomposition up to 400 °C of ammonium ferric citrate hydrate, of unknown structure and formula weight, was studied by thermogravimetry, differential thermal analysis, infrared (IR) spectroscopy and X-ray diffractometry. The possible identities of the formula weight and the intermediate products of calcination are discussed. The results revealed that the parent material is amorphous and contains two moles of water and two moles of ammonia. Decomposition takes place via six weight-loss processes, three endothermic (90–230 °C) and three exothermic (240–298 °C), leading eventually to the formation of Fe2O3. The intermediate solid products are mainly unstable amorphous oxycarbonates, as indicated by X-ray and IR spectroscopies. The gas-phase decomposition products identified by IR spectroscopy are NH3, CO2, CO, CH3COCH3, CH4 and NH4OH. Surface area measurement and scanning electron microscopy showed that Fe2O3, the final product at 400 °C, hada surface area of 40 m2/g and good crystalline and porous character.  相似文献   

11.
Dispersing La2O3 on δ- or γ-Al2O3 significantly enhances the rate of NO reduction by CH4 in 1% O2, compared to unsupported La2O3. Typically, no bend-over in activity occurs between 500° and 700°C, and the rate at 700°C is 60% higher than that with a Co/ZSM-5 catalyst. The final activity was dependent upon the La2O3 precursor used, the pretreatment, and the La2O3 loading. The most active family of catalysts consisted of La2O3 on γ-Al2O3 prepared with lanthanum acetate and calcined at 750°C for 10 h. A maximum in rate (mol/s/g) and specific activity (mol/s/m2) occurred between the addition of one and two theoretical monolayers of La2O3 on the γ-Al2O3 surface. The best catalyst, 40% La2O3/γ-Al2O3, had a turnover frequency at 700°C of 0.05 s−1, based on NO chemisorption at 25°C, which was 15 times higher than that for Co/ZSM-5. These La2O3/Al2O3 catalysts exhibited stable activity under high conversion conditions as well as high CH4 selectivity (CH4 + NO vs. CH4 + O2). The addition of Sr to a 20% La2O3/γ-Al2O3 sample increased activity, and a maximum rate enhancement of 45% was obtained at a SrO loading of 5%. In contrast, addition of SO=4 to the latter Sr-promoted La2O3/Al2O3 catalyst decreased activity although sulfate increased the activity of Sr-promoted La2O3. Dispersing La2O3 on SiO2 produced catalysts with extremely low specific activities, and rates were even lower than with pure La2O3. This is presumably due to water sensitivity and silicate formation. The La2O3/Al2O3 catalysts are anticipated to show sufficient hydrothermal stability to allow their use in certain high-temperature applications.  相似文献   

12.
A 1% Pd catalyst (38% dispersion) was prepared by impregnating a γ-alumina with palladium acetylacetonate dissolved in acetone. The behaviour of this catalyst in oxidation and steam reforming (SR) of propane was investigated. Temperature-programmed reactions of C3H8 with O2 or with O2 + H2O were carried out with different stoichiometric ratios S(S =[O2]/5[C3H8]). The conversion profiles of C3H8 for the reaction carried out in substoichiometry of O2 (S < 1) showed two discrete domains of conversion: oxidation at temperatures below 350°C and SR at temperatures above 350°C. The presence of steam in the inlet gases is not necessary for SR to occur: there is sufficient water produced in the oxidation to form H2 and carbon oxides by this reaction. Contrary to what was observed with Pt, an apparent deactivation between 310 and 385°C could be observed with Pd in oxidation. This is due to a reduction of PdOx into Pd0, which is much less active than the oxide in propane oxidation. Steam added to the reactants inhibits oxidation while it prevents the reduction of PdOx into Pd0. Compared to Pt and to Rh, Pd has a higher thermal resistance: no deactivation occurred after treatment up to 700°C and limited deactivation after treatment up to 900°C, provided that the catalyst is maintained in an oxygen-rich atmosphere during the cooling.  相似文献   

13.
The influence of catalyst pre-treatment temperature (650 and 750 °C) and oxygen concentration (λ = 8 and 1) on the light-off temperature of methane combustion has been investigated over two composite oxides, Co3O4/CeO2 and Co3O4/CeO2–ZrO2 containing 30 wt.% of Co3O4. The catalytic materials prepared by the co-precipitation method were calcined at 650 °C for 5 h (fresh samples); a portion of them was further treated at 750 °C for 7 h, in a furnace in static air (aged samples).

Tests of methane combustion were carried out on fresh and aged catalysts at two different WHSV values (12 000 and 60 000 mL g−1 h−1). The catalytic performance of Co3O4/CeO2 and Co3O4/CeO2–ZrO2 were compared with those of two pure Co3O4 oxides, a sample obtained by the precipitation method and a commercial reference. Characterization studies by X-ray diffraction (XRD), BET and temperature-programmed reduction (TPR) show that the catalytic activity is related to the dispersion of crystalline phases, Co3O4/CeO2 and Co3O4/CeO2–ZrO2 as well as to their reducibility. Particular attention was paid to the thermal stability of the Co3O4 phase in the temperature range of 750–800 °C, in both static (in a furnace) and dynamic conditions (continuous flow). The results indicate that the thermal stability of the phase Co3O4 heated up to 800 °C depends on the size of the cobalt oxide crystallites (fresh or aged samples) and on the oxygen content (excess λ = 8, stoichiometric λ = 1) in the reaction mixture. A stabilizing effect due to the presence of ceria or ceria–zirconia against Co3O4 decomposition into CoO was observed.

Moreover, the role of ceria and ceria–zirconia is to maintain a good combustion activity of the cobalt composite oxides by dispersing the active phase Co3O4 and by promoting the reduction at low temperature.  相似文献   


14.
The coupled substitution of Bi3+ for Pb2+ and B3+ for Ge4+ was successfully achieved in ferroelectric Pb5Ge3O11. Large single crystals of optical quality were grown from the melt. Deterioration of the crystal quality could not be observed within a period of two years under ambient conditions. The dielectric permittivity εr, electric conductivity p and pyroelectric coefficient γ are investigated as a function of temperature. The room temperature data (εr = 43, γ = 2.4·10-4Cm-2 K-1, p = 7·107 Ω, volume specific heat s = 2·106 Jm-3K-1) show the usefulness of Pb5-xBixGe3-xBxO11 around x = 0.1 as a material for pyroelectric thermal detectors operating at room temperature. Also the difficulty of producing lead germanate silicate mixed crystals in the size and quality needed for vidicon applications is overcome by additional replacement of some Pb and Ge by Bi and B.  相似文献   

15.
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   

16.
In a paste CA2 hydrates very slowly at 20°C but the reaction is more rapid at 40°C. At both temperatures the products of hydration comprise C2AH8, gibbsite and alumina gel after 7 days. For mixtures of CA2 + CA the initial hydration at 20°C is dominated by CA. In the presence of ground granulated blastfurnace slag CA2 or a mixture of CA2 + CA will form strätlingite on hydration at 40°C within 3 days. Slag addition reduces the amount of heat evolved during the first 24 hours of hydration at both 20° and 40°C.  相似文献   

17.
Spontaneous precipitation in the aqueous system Mg+2-Na+-SO3-2-SO4-2 affected by mixing solutions of MgSO4 and Na2SO3, together was studied for temperature varying from 40 to 80°C and for pH from 5,5 to 9. The initial composition of a precipitating system was 0·67 and 1·17 mol of MgSO3 and Na2SO4 per liter and 0·82 mol of MgSO3 and Na2SO4 together with 0·83 mol of MgSO4 per liter, respectively. Depending on the prevailing reaction conditions, solid phase consisting of MgSO3 6H2O, MgSO3 3H2O, Mg2NaOH(SO3)2 H2O or their mixture is formed. Each solid phase forms crystals of typical size and shape. The precipitation diagrams drawn in the temperature and pH coordinates for three different initial composition of the studied system are presented.  相似文献   

18.
The nucleation and growth of Pd clusters in mordenite were investigated using in situ extended X-ray absorption fine structure (EXAFS) spectroscopy and Fourier transform infrared (FTIR) spectroscopy of absorbed CO. Calcination of [Pd(NH3)4]2+-exchanged mordenite at 350°C in O2 results in decomposition of the amine complex and formation of square-planar Pd2+ oxo species within the mordenite pores. Reduction of these species at 150°C in H2 yields Pd clusters with an average nuclearity of 3. On an average two 2.22 Å Pd-O bonds are associated with each Pd3 cluster; we infer that this interaction serves to anchor the clusters within the pores. After reduction at 150°C, the FTIR spectrum of irreversibly adsorbed CO is indicative of a mixture of Pd+, Pdδ+, and Pd0 carbonyl species. Reduction at 350°C produces larger intrazeolitic Pd clusters (average nuclearity of 6) that exhibit only a weak interaction with the mordenite, as evidenced by their facile aggregation in the presence of CO at 30°C. Reduction at 450°C yields large 20 Å Pd clusters that we infer are located on external mordenite surfaces or locally disrupt the intracrystalline structure.  相似文献   

19.
Iron (III) adsorption from aqueous solutions onto raw and pretreated clinoptilolite was investigated here. Various parameters for iron removal; initial solution pH, contact time and metal ion concentration were optimized. The equilibrium data were modeled by both the Langmuir and Freundlich adsorption isotherms at optimal conditions. Adsorption capacities of raw samples and those pretreated with Na2S2O8 at 20 °C , 70 °C and with HNO3 at 20 °C were all similar but samples pretreated with HNO3 at 70 °C were significantly different; iron (III) removal from samples pretreated with HNO3 decreased with increasing pretreatment temperature. Tests with Fe+3 solutions containing phenol, CsCl or KCl, indicated the continued presence of these ions in zeolite which either promoted or retarded the adsorption of iron. The Fe+3 adsorption capacity of clinoptilolite pretreated with HNO3 at 70 °C was about two times greater with, than without, CsCl and KCl. The kinetics of iron adsorption from aqueous solution were also investigated using the first-order Lagergren equation and a pseudo-second-order model.  相似文献   

20.
(CH3)NHgBrI2 is an electro-optic material which shows a structural phase transition with a symmetry change from orthorhombic to monoclinic at T = 108°C upon heating. In this contribution we report first experimental data on the optical properties (refractive indices, absorption coefficients), on the elastic and electro-optical properties. It is shown, that the dispersion of the refractive indices is adequately described by a single term Sellmeier oscillator model. Measurements of the temperature dependence of the birefringence and of the elastic constants show pronounced changes at the structural phase transition. The biaxial crystal with the three main indices na = 1.761±0.001, nb = 1.802±0.001 and nc = 1.755±0.001 at room temperature and a wavelength λ=633nm becomes optically uniaxial at T = (50±0.5)°C with na=nc=1.750, due to different temperature dependences na(T) and nb(T). Room temperature electro-optic coefficients and the corresponding half-wave voltages have been determined, and indicate that the electro-optic activity of (CH3)4NHgBrI2 is about two orders of magnitude smaller than in KH2PO4.  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