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.
Isothermal oxidation of dense TiC ceramics, fabricated by hot-isostatic pressing at 1630°C and 195 MPa, was performed in Ar/O2 (dry oxidation), Ar/O2/H2O (wet oxidation), and Ar/H2O (H2O oxidation) at 900°–1200°C. The weight change measurements of the TiC specimen showed that the dry, wet, and H2O oxidation at 850°–1000°C is represented by a one-dimensional parabolic rate equation, while the oxidation in the three atmospheres at 1100° and 1200°C proceeds linearly. Cross-sectional observation showed that the dry oxidation produces a lamellar TiO2 scale consisting of many thin layers, about 5 μm thick, containing many pores and large cracks, while H2O-containing oxidation decreases pores in number and diminishes cracks in scales. Gas evolution of CO2 and H2 with weight change measurement was simultaneously followed by heating the TiC to 1400°C in the three atmospheres. Cracking in the TiO2 scale accompanied CO2 evolution, and the H2O-containing oxidation produced a small amount of H2. A piece of single crystal TiC was oxidized in 16O2/H218O to reveal the contribution of O from H2O to the oxidation of TiC by secondary ion mass spectrometry.  相似文献   

2.
The final color of a ceramic body is generally determined by the contents of Fe3+ ions. The Central European market accepts kaolinitic clays for the production of tableware, electro-insulators, and wall tiles only if the Fe2O3 in chemical analyses of the kaolin does not exceed 1.0 wt%. The chemical analyses calculate the content of all iron components as if they were in the form of Fe2O3 and then their quantity excludes even good-quality clay from the ceramic industry of white bodies. We found that oxidizing firing of the samples fired at over 1180°C changes the color in cases where the initial Fe3+ component is in the form of hydro ferric oxide [FeO(OH)] or if the ferric ions were added to the white kaolin samples in the form of ferric nitrate [Fe(NO3)3·9H2O]. The Mössbauer spectroscopy confirms the presence of only diluted Fe3+ ions. The UV-ViS-NIR spectroscopy confirms that even if the concentration of Fe2O3 is above 3.0 wt%, these ions of Fe3+ in the case of their initial hydro ferric oxide formed in clay are incorporated into the mullitic structure in tetrahedral coordination. The iron-coloring effect depends on the coordination of Fe3+ ions with the studied discoloring effect of fired bodies—the very small and well-distributed particles enter into the formatted mullitic structure.  相似文献   

3.
Single-crystal X-ray and electron-diffraction studies show the existence in one polymorph of 4CaO.Al2O3. 13H2O of a hexagonal structural element with α= 5.74 a.u., c = 7.92 a. u. and atomic contents Ca2(OH)7- 3H2O. These structural elements are stacked in a complex way and there are probably two or more poly-types as in SiC or ZnS. Hydrocalumite is closely related to 4CaO.A12O3.13H2O, from which it is derived by substitution of CO32-for 20H-+ 3H2O once in every eight structural elements; similar substitutions explain the existence of compounds of the types 3CaO Al2O3.Ca Y 2- xH2O and 3CaO Al2O3 Ca Y xH2O. On dehydration, 4CaO.Al2O3.13H2O first loses molecular water and undergoes stacking changes and shrinkage along c. At 150° to 250°C., Ca(OH)2 and 4CaO.3Al2O3.3H2O are formed and, by 1000°C., CaO and 12CaO.7Al2O8. The dehydration of hydrocalumite follows a similar course, but no 4CaO.3Al2O3.3H2O is formed.  相似文献   

4.
The potassium ions in potassium β-ferrite ((1 + x)K2O ·11Fe2O3) crystals were exchanged with Na+, Rb+, Cs+, Ag+, NH4+, and H3O+ in molten nitrates or in concentrated H2SO4. On the other hand, spinel and hexagonal ferrites were formed by soaking the crystals in the melt of divalent salts. The crystals of K+, Rb+, and Cs+β-ferrites decomposed to form α-Fe2O3 at high temperatures of 800° to 1100°C. In addition, H3O+, NH4+, and Ag+β-ferrites decomposed to form α-Fe2O3 at relatively low temperatures of 350° to 650°C, in accordance with the stabilities of the inserted ions. The electrical properties of some β-ferrites were measured.  相似文献   

5.
Samarium ions (Sm2+) incorporated into aluminosilicate glasses by a sol-gel process showed persistent spectral hole burning at room temperature. Gels of the system Na2O-Al2O3SiO2 synthesized by the hydrolysis of Si(OC2H5)4, Al(OC4H9)3, CH3 COONa, and SmCl3·6H2O were heated in air at 500°C, then reacted with H2 gas to form Sm2+ ions. Whereas Al3+ ions effectively dispersed the Sm3+ ions in the glass structure, Na+ ions were not effective. The Al2O3-SiO2 glasses proved appropriate for reacting the Sm3+ ions with H2 gas and exhibited the intense photoluminescence of Sm2+ ions. The reaction of Sm3+ ions with H2 in the Al2O2-SiO2 glasses was determined by first-order kinetics, and the activation energy equaled 95 kJ/mol. At 800°C, the maximum photoluminescence of the Sm2+ ions was achieved within 20 min.  相似文献   

6.
Reactive Ceria Nanopowders via Carbonate Precipitation   总被引:3,自引:0,他引:3  
Nanocrystalline CeO2 powders have been successfully synthesized via a carbonate precipitation method, using ammonium carbonate (AC) as the precipitant and cerium nitrate hexahydrate as the cerium source. The AC/Ce3+ molar ratio ( R ) affects significantly precursor properties, and spherical nanoparticles can be produced only in a narrow range of 2 < R ≤ 3. The precursor, having an approximate composition of Ce(OH)CO3·2.5H2O, decomposes to CeO2 at temperatures ≥300°C. The CeO2 powder calcined at 700°C exhibits high reactivity and can be densified to >99% of theoretical at 1000°C.  相似文献   

7.
Equilibrium NiO solubility measurements were made in Li2CO3/K2CO3 mixtures as a function of temperature, ambient gas environment, and salt composition for gases containing 3.1% H2O. The equilibrium solubility was found to increase with increasing temperature, CO2 partial pressure, and cation fraction Li+ and to decrease slightly with increasing O2 partial pressure. These results were coupled with theoretical predictions of (1) O2, H2, and CO2 activities and (2) compositional gradients which develop across the electrolyte structure of an operating carbonate fuel to explain the two distinct regions of nickel precipitation across the fuel cell. Precipitation occurs near the cathode because NiO solubility decreases commensurate with electrolyte compositional gradients, whereas the second region deeper within the electrolyte matrix results from a decreased O2 activity.  相似文献   

8.
The synthesis and characterization of yttrium hydroxyl carbonate (Y(OH)CO32−) and yttrium nitrate hydroxide hydrate (Y(OH)NO3H2O) precursor materials as well as Y2O3 nanoparticles are reported. The resultant precursor particle size is about 10–12 nm with a narrow size distribution by the enzymatic decomposition method, whereas the particle size was smaller than those acquired by the homogeneous and alkali precipitation methods. The formation of Y(OH)CO32− and Y(OH)NO3H2O species was also evident from the fourier-transform infrared spectrometry (FT-IR) analysis. Precipitated Y(OH)CO32− precursors have an amorphous nature whereas Y(OH)NO3H2O precursors have a crystalline nature, which was manifested from the XRD analysis. Moreover, precipitated Y(OH)NO3H2O precursors were found in the agglomerated form and Y(OH)CO32− was established in the monodispersed form, as determined from the FE-SEM, TEM and DLS measurements. It was demonstrated that calcination of precursor materials at 900°C eventually removed the inorganic anions from the precursors and consequently produced crystalline Y2O3 nanoparticles, which was evident from the XRD and FT-IR analysis. The EDS analysis confirms Er3+ doping in the Y2O3 nanoparticles. The morphology and the size of the Y2O3 nanoparticles are almost unchanged before and after the calcination.  相似文献   

9.
Low-Temperature Synthesis of Praseodymium-Doped Ceria Nanopowders   总被引:1,自引:0,他引:1  
Praseodymium-doped ceria (CeO2) nanopowders have been synthesized via a simple but effective carbonate-coprecipitation method, using nitrates as the starting salts and ammonium carbonate as the precipitant. The precursors produced in this work are ammonium rare-earth double carbonates, with a general formula of (NH4)0.16Ce1− x Pr x (CO3)1.58·H2O (0 < x ≤ 0.20), which directly yield oxide solid solutions on thermal decomposition at a very low temperature of ∼400°C. Praseodymium doping causes a gradual contraction of the CeO2 lattice, because of the oxidation of Pr3+ to smaller Pr4+, and suppresses crystallite coarsening of the oxides during calcination. Dense ceramics have been fabricated from the thus-prepared nanopowders via pressureless sintering for 4 h at a low temperature of 1200°C.  相似文献   

10.
The stability of MoSi2 in combustion gas at 1370° and 1600°C was evaluated using SOLGASMIX-PV thermodynamic modeling, periodic weight measurements, and characterization via XRD, SEM, EDS, and image analysis. Passive oxidation occurred at both temperatures. During an initial stage of exposure, specimen surfaces oxidized to form MoO3(g) and amorphous SiO2 via reduction of CO2 and H2O. After a short time (<6.5 min at 1370°C, <1 min at 1600°C), the oxidation mechanism switched; Mo5Si3 and amorphous SiO2 formed as oxidation products. The first mechanism esulted in the formation of 46.1 vol% at 1370°C and 42.6 vol% at 1600°C of the amorphous silica surface coating. The attainment of a near-terminal weight gain implied silica formation was limited by H2O and CO2 diffusion through the silica coating.  相似文献   

11.
The wetting of polycrystalline alumina by a colored, calciamagnesia aluminosilicae glass was found to be dependent on temperatutre between 1300° and 1500°C, but independent of gas atmosphere effects. Neither the oxygen partial pressure, oveer the range of 10-6 to 10-10 Pa, the gas buffer system (Co/CO2 or H2/H2O), nor pre-equilibration of the substrate surface with the atmosphere at tge exoperimental temperature before solid-liuid interface formation affected the stable contact angle. An initial drop in contact angle the stable contact angle. An initial drop in contact angle occurring within the first hour is attributed to repaid dissolution of alumina and the formation of a stable glass/alumina interface. The contact angle after an 8-h isothermal hold decreased from 1300° to 1500°C. The solid-liquid interfacial energy, μMS1, controls the wetting behavior. Changes in μMs1 are attributed to he breakup of the silica network as temperature increases.  相似文献   

12.
Two wet-chemical routes have been used to synthesize Sc2O3 nanopowders from nitrate solutions employing ammonia water (AW) and ammonium hydrogen carbonate (AHC) as the precipitants. The precursors and the resultant oxides are characterized by elemental analysis, X-ray diffractometry, differential thermal analysis/thermogravimetry, high-resolution scanning electron microscopy, and Brunauer-Emmett-Teller analysis. Crystalline γ-ScOOH· n H2O ( n ≈ 0.5) is the only phase obtained by the AW method. This phase dehydrates to Sc2O3 at ∼400°C, yielding hard aggregated nanocrystalline Sc2O3 powders. Three types of precursors have been synthesized by the AHC method, depending on the AHC/Sc3+ molar ratio ( R ): amorphous basic carbonate [Sc(OH)CO3·H2O] at R ≤ 3, crystalline double carbonate [(NH4)Sc(CO3)2·H2O] at R ≥ 4, and a mixture of the two phases at 3 < R < 4. Among these precursors, only the basic carbonate shows spherical particle morphology, ultrafine particle size (∼50 nm), and weak agglomeration. Sc2O3 nanopowders (∼28 nm) with high surface area (∼49 m2/g) have been prepared by calcining the basic carbonate at 700°C for 2 h.  相似文献   

13.
Amorphous WO3 or WO3 or H2O is formed by hydrolysis of tungsten ethoxide. The temperature of hydrolysis influences the crystallization of WO3·H2O. Tungsten hydrate (WO3·H2O) has an orthorhombic unit cell with a=0.5235 nm, b = 1.0688 nm, and c=0.5123 nm. Orthorhombic WO3 crystallizes at 350° to 500°Cfrom amorphous WO3. Cubic WO3 is formed at 200° to 310°C with dehydration of WO3·H2O. WO3 transformations are examined by high-temperature X-ray diffraction. The kinetics of formation of the cubic modification have been studied by measuring the weight decrease with a thermobalcnce. Formation isotherms can be interpreted in terms of the first-order equation –In (1–f)=kt; activation energies are 110 and 80 kJ mol−1 for initial and final stages, respectively.  相似文献   

14.
Synthesis of Titanate Derivatives Using Ion-Exchange Reaction   总被引:3,自引:0,他引:3  
Two types of titanate derivatives, layered hydrous titanium dioxide (H2Ti4O9· n H2O) and potassium octatitanate (K2Ti8O17) with a tunnellike structure, were synthesized using an ion-exchange reaction. Fibrous potassium tetratitanate (K2Ti4O9· n H2O) was prepared by calcination of a mixture of K2CO3 and TiO2 with a molar ratio of 2.8 at 1050°C for 3 h, followed by boiling-water treatment of the calcined products for 10 h. The material then was transformed to layered H2Ti4O9· n H2O through an exchange of K+ ions with H+ ions using HCl. K2Ti8O17 was formed by a thermal treatment of KHTi4O9· n H2O. Pure KHTi4O9· n H2O phase was effectively produced by a treatment of K2Ti4O9 with 0.005 M HCl solution for 30 min. Thermal treatment at 250°–500°C for 3 h resulted in formation of only K2Ti8O17.  相似文献   

15.
Values of the spectral absorption coefficient (α) of liquid aluminum oxide were determined by transmission of a pulsed dye laser beam incident on continuous-wave (CW) CO2-laser-melted pendant drops attached to sapphire filaments. Measurements were made on molten drops of Verneuil sapphire at wavelengths of 0.450 and 0.633 μm, at ambient oxygen partial pressures from 10-10 to 1 bar in eight pure gases (Ar, CO, CO2, H2, H2O, HCI, N2 and O2), in CO/CO2 mixtures, and in H2/H2O mixtures, and at a temperature of ca. 2400 K. Specimens contaminated with iron, magnesium, silicon, and tungsten were also investigated in an oxygen atmosphere. At a wavelength of 0.633 μm, the value of α was greater than 50 cm-1 under reducing or inert gas conditions. It decreased to a minimum at intermediate oxygen partial pressures of 5 × 10-5 bar in CO/CO2 mixtures and 5 × 10-3 bar in H2/H2O mixtures, and increased at larger oxygen partial pressures. The specimens were opaque (α > 55 cm-1) in hydrogen, in HCI at pressures above 0.04 bar. Specimens contaminated with 5000-10000 ppm of Fe, Mg, Si, or W were also opaque. At a wavelength of 0.45 μm the liquid aluminum oxide specimens were opaque in Ar and oxygen, and gave α= 46 cm-1 in CO2. The dynamic response when the ambient gas was changed from CO2 to argon showed that the transmission maximum for = 0.45 μm was at p (O2) < 0.1 bar.  相似文献   

16.
The reduction by hydrogen of Fe3+, Ce4+, and Sn4+ in soda-lime-silica glass was found to be diffusion-limited in the glass transition temperature range (500°C). The reduction was studied in fibers by chemical analysis and in bulk samples by optical absorption. Reduction to Fe2+, Ce3+, and Sn2+ proceeds by a reaction of the type H2+2(ɛSi-O)-+2Fe3+→2Fe2++2(ɛSi–OH). Since the rate of reduction by hydrogen is quite similar for these systems and since reduction cannot be accomplished with CO, it is concluded that hydrogen is the diffusing species. A mechanism is developed in which hydrogen diffuses through the glass until it is trapped by a reducible ion.  相似文献   

17.
The compound BaFeO3- x exists in many forms, the hexagonal phase having a wide range in oxygen content (BaFeO2.63–2.92). The other phases have the limited compositions and the distinct structures of perovskite: triclinic I, BaFeO2.50; triclinic II, BaFeO2.64–2.67; rhombohedral I and II, BaFeO2.62–2.64; and tetragonal, BaFeO2.75–2.81. The phases contain Fe4+ ions correlating with the oxygen content. The hexagonal phase shows a continuous change in oxygen content with temperature down to BaFeO2.63. The perovskitelike phases, however, show characteristic transformations. The triclinic I phase oxidizes to the triclinic II form at 320° to 500°C and to the hexagonal form at 720° to 915°C in oxygen. These transformations are related to oxidation-reduction of iron ions (Fe3+⇌ Fe4+).  相似文献   

18.
Waveguides in Lithium Niobate   总被引:1,自引:0,他引:1  
Lithium niobate (LiNbO3 is considered to be the leading electrooptical material for fabrication of active waveguides, modulators, and switches for application in integrated optical circuits. LiNbO3 is a ferroelectric material with a high Curie temperature of 1210°C. This is essential to permit rapid diffusion at high temperature, e.g., >1000°C, without domain reversal. Its high electrooptic coefficient allows modulation at relatively low voltages. Waveguides are fabricated in LiNbO3 by Ti indiffusion around 1000°C, by Li2O outdiffusion, and, more recently, by Li+/H+ ion exchange. We report results of a study on diluted Li+/H+ ion exchange. This process is attractive because it allows control of the index change between 0.003 and 0.1 Furthermore, the fabricated waveguides are polarization selective, have high optical damage resistance, and do not exhibit index instability as is the case with complete Li+/H+ exchange. The variation in both the index of refraction and the diffusion coefficient with composition is very nonlinear. Mechanisms contributing to these nonlinear properties are discussed.  相似文献   

19.
A (Ce0.67Tb0.33)Mn x Mg1− x Al11O19 phosphor powder was synthesized, using a simple sol–gel process, by mixing citric acid with CeO2, Tb4O7, Al(NO3)3·9H2O, Mg(OH)2·4MgCO3·6H2O, and Mn(CH3COO)2. The phosphor crystallized completely at 1200°C, and the phosphor particle size was between 1 and 5 μm. The excitation spectrum was characteristic of Ce3+, while the emission spectrum was composed of lines from Tb3+ and Mn2+. The Mn2+ gave a green fluorescence band, and concentration quenching occurred when x > 0.10. The luminescent properties of the phosphor were explained by a configurational coordinate model.  相似文献   

20.
The compositions of anion-deficient zirconia and thoria in equilibrium with O2 were measured from 1 to 10−6 atm and 1400° to 1900°C; for ZrO2- x (po2 in atm, and T in °K), log x∼−0.890-[(0.400×104)/ T ]-[(log p )/6]; for ThO2- x , log x∼−1.870-[(0.340×104)/ T ]-[(log p )/6]. The ZrO2- x -Zr boundary was located at x=0.014 at 1800°C; thoria was single-phase over the entire range. Consistent results were obtained when O2/inert gas mixtures were used, but use of H2/H2O and CO/CO2 at 1000° to 1200°C gave abnormal and, in the latter case, erratic data; side reactions in these atmospheres are inferred. The monoclinic-tetragonal phase change of ZrO2 and the lattice thermal expansion, room-temperature Young's modulus, and strength properties of ZrO2 and ThO2 bodies were not appreciably altered by oxygen deficiency. The lattice dimensions decreased slightly with departure from stoichiometry.  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