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 tensile creep behavior of two ceramic composite systems exhibiting duplex microstructures was studied relative to their single-phase constituents in the temperature and stress ranges of 1100–1350°C and 35–75 MPa. The equivolumetric compositions in the Al2O3: c –ZrO2 (8 mol% Y2O3) and Al2O3:Y3Al5O12 systems both exhibit lower creep rates than either of their single-phase constituents. This effect is attributed to Y3+ (and possibly Zr4+) present in the A12O3 as a segregant which lowers the creep rate by ∼2 orders of magnitude. It is believed that the segregation of Y3+ to the A12O3 grain boundaries hinders the interface reaction believed to control the creep. If one of the single-phase constituents is taken to be the Y3+-doped Al2O3, the creep of the duplex microstructures can be modeled using standard composite theory applied to flow.  相似文献   

2.
The diffusive transport of chromium in both pure and Y-doped fine-grained alumina has been investigated over the temperature range 1250°–1650°C. From a quantitative assessment of the chromium diffusion profile in alumina, as obtained from electron microprobe analysis, it was found that yttrium doping retards cation diffusion in the grain-boundary regime by over an order of magnitude. The Arrhenius equations for the undoped and Y-doped samples were determined to be: δ D b=(4.77±0.24) × 10−7 exp (−264.78±47.68 (kJ/mol)/RT)(cm3/s) and δDb=(6.87±0.18) × 10−8 exp (−284.91±42.57 (kJ/mol)/RT)(cm3/s), respectively. Finally, to elucidate the mechanism for this retardation, the impact of yttrium doping on diffusion activation energies and prefactors was examined.  相似文献   

3.
The microstructure in Y2O3-stabilized tetragonal zirconia polycrystal (Y-TZP) sintered at 1300°–1500°C was examined to clarify the role of Y3+ ions on grain growth and the formation of cubic phase. The grain size and the fraction of the cubic phase in Y-TZP increased as the sintering temperature increased. Both the fraction of the tetragonal phase and the Y2O3 concentration within the tetragonal phase decreased with increasing fraction of the cubic phase. Scanning transmission electron microscopy (SEM) and X-ray energy dispersive spectroscopy (EDS) measurements revealed that cubic phase regions in grain interiors in Y-TZP generated as the sintering temperature increased. High-resolution electron microscopy and nanoprobe EDS measurements revealed that no amorphous layer or second phase existed along the grain-boundary faces in Y-TZP and Y3+ ions segregated at their grain boundaries over a width of ∼10 nm. Taking into account these results, it was clarified that cubic phase regions in grain interiors started to form from grain boundaries and the triple junctions in which Y3+ ions segregated. The cubic-formation and grain-growth mechanisms in Y-TZP can be explained using the grain boundary segregation-induced phase transformation model and the solute drag effect of Y3+ ions segregating along the grain boundary, respectively.  相似文献   

4.
The volatility of UO2±x and the phase relations in the system uranium-oxygen were studied using thermogravimetric techniques. Chemical reactions describing the loss of uranium from UO2±x at temperatures between 1100° and 2200°C in oxygen pressures between approximately 102 and 10−6 torr are proposed. Results were obtained requiring the consideration of UO4(g) as the uranium-bearing vapor species above UO2±x. Evidence supporting the existence of UO4(g) included the volatilization of material with an oxygen-to-uranium ratio of 4 during the decomposition of UO2+x(0.Z > × > 0) to near-stoichiometric UO2 in vacuum above 1500°C and the dependence of the evaporation rate of the uranium dioxide on the oxygen pressure between 1200° and 1500°C. The equilibrium oxygen pressures over compositions between UO2.02 and UO2.63 in the UO2+x and U3O8-y regions and over the boundary between these phases were measured between 1000° and 1600°C. The equilibrium oxygen-to-uranium ratio of UO2±x was less than 2 above 1700°C in vacuum.  相似文献   

5.
The influence of Nd2O3 doping on the reaction process and sintering behavior of BaCeO3 is investigated. Formation of BaCeO3 is initiated at 800°C and completed at 1000°C. When Nd2O3 is added to the starting materials, the formation of BaCe1–xNdxO3–δ is delayed and the temperature for complete reaction is increased to 1100°C. Only a BaCe1-xNdxO3–δ solid solution with an orthorhombic crystal structure is present in the specimens for x ≤ 0.1. A secondary phase rich in Ce and Nd is formed within grains and at grain boundaries, when the Nd2O3 content is greater than the solubility limit (x ≥ 0.2). Pure BaCeO3 is difficult to sinter, even at 1500°C, and only a porous microstructure could be obtained. However, doping BaCeO3 with Nd2O3 markedly enhances its sinterability. The enhancement of the sinterability of Nd2O3-doped specimens at x ≤ 0.1 is attributed to the increase in the concentration of oxygen ion vacancies, which increases the diffusion rate. At x ≥ 0.2, the grain size is abnormally coarsened, which is caused by the formation of a liquid phase. While this liquid phase accelerates sintering, its beneficial effect on densification is counteracted by the segregation of the secondary grain-boundary phase which inhibits sintering.  相似文献   

6.
CeO2 samples doped with 10, 1.0, and 0.1 mol% Y2O3 and undoped CeO2 samples of high purity were studied by impedance spectroscopy at temperatures <800°C and under various oxygen partial pressures. According to microstructural investigations by SEM and analytical STEM (equipped with EDXS), the grain boundaries were free of any second phase, providing direct grain-to-grain contacts. An amorphous siliceous phase was detected at only a few triple junctions, if at all; as a result, its contribution to the grain-boundary resistance was negligible. Nevertheless, the specific grain-boundary conductivities were still 2–7 orders of magnitude lower than the bulk conductivities, depending on dopant concentration, temperature, and oxygen partial pressure. The charge carrier transport across the grain boundaries occurred only through the grain-to-grain contacts, whose properties were then determined by the space-charge layer. The space-charge potential in acceptor-doped CeO2 was positive, causing the simultaneous depletion of oxygen vacancies and accumulation of electrons in the space-charge layer. The very low grain-boundary conductivities can be accounted for by the oxygen-vacancy depletion; the accumulation of electrons became evident in weakly doped and undoped CeO2 at high temperatures and under low oxygen partial pressures.  相似文献   

7.
High-temperature piezoelectric ceramics based on W6+-doped Bi4Ti3O12 (W-BIT) were prepared by both the conventional mixing oxides and the chemical coprecipitation methods. Sintering was carried out between 800° and 1150°C in air. A rapid densification, >99% of the theoretical density (rhoth) at 900°C/2 h, took place in the chemically prepared W6+-doped Bi4Ti3O12 ceramics, whereas conventionally prepared BIT-based materials achieved a lower maximum density, ∼94% of rhoth, at higher temperature (1050°C). The microstructure study revealed a platelike morphology in both materials. Platelike grains were larger in the conventionally prepared W-BIT-based materials. The sintering behavior could be related both to the agglomeration state of the calcined powders and to the enlargement of the platelets at high temperature. The W6+-doped BIT materials showed an electrical conductivity value 2-3 orders of magnitude lower than undoped samples. The electrical conductivity increased exponentially with the aspect ratio of the platelike grains. The addition of excess TiO2 produced a further decrease of the electrical conductivity.  相似文献   

8.
Fluorescence spectra resulting from uv excitation of pure, Ca2+-doped, and Y3+-doped ThO2 were resolved into discrete bands at 395.4, 434.6, and 464.0 nm. These bands are related to absorption bands in the pure material at 238.9, 227.3, and 252.8 nm, respectively. The Ca2+ - and Y3+-doped ThO2 specimens showed an apparent absorption edge at 240 nm (213 nm is the true edge in pure ThO2). However, the fluorescence studies showed that this edge is, in fact, spurious, resulting from enhancement of the high-energy absorbance bands by doping.  相似文献   

9.
The A-site cation diffusion in LaFeO3 has been examined by inter-diffusion experiments between LaFeO3 and NdFeO3. Dense, polycrystalline bodies were annealed in contact at temperatures between 1100° and 1300°C in ambient air. The bulk- and grain-boundary inter-diffusion coefficients were calculated from concentration profiles determined by electron probe micro analysis of cross sections. The bulk- and grain-boundary inter-diffusion coefficients showed Arrhenius-type behavior with activation energies 610±30 and 600±100 kJ/mol, respectively. Based on the assumption of 1 nm thick grain boundaries the grain-boundary inter-diffusion coefficient was ∼4 orders of magnitude higher than the bulk inter-diffusion coefficient.  相似文献   

10.
Al2O3-5 vol% SiC particle composites doped with 800 ppm rare-earth impurities (Y3+, Nd3+, and La3+) were fabricated by hot-pressing at a temperature of 1550°C. Doping of rare-earth impurities in Al2O3-SiC composites led to a fracture- mode change from transgranular in dopant-free composites to intergranular in rare-earth-doped composites. The fracture mode change obviously increased the crack deflection so that the fracture toughness of rare-earth-doped composites was higher than that of the composites without dopants, especially for the Nd3+- and La3+-doped composites. It was found that the fracture-mode change originated from a weak grain-boundary bonding caused by co-segregation of the rare-earth dopants and Si4+ ions dissolved from the SiC particle surfaces.  相似文献   

11.
The breakup of dislocation dipoles in plastically deformed samples of undoped and 30-ppm-MgO-doped sapphire (α-Al2O3) was monitored using conventional TEM techniques. Dislocation dipoles break up into prismatic dislocation loops in a sequential process during annealing; i.e., dislocation loops are pinched off at the end of a dislocation dipole. This pinch-off process is primarily controlled by pipe diffusion, and pipe diffusion coefficients at temperatures between 1300° and 1500°C were estimated by monitoring the kinetics of the dipole breakup process. We determined D PU= 8.1(–4.3+9.1) × 10–3 exp [–(4.5 ± 1.3 eV )/ kT )] m2/s for the undoped material. The pipe diffusion kinetics for the MgO-doped crystal was determined at 1250° and 1300°C and was about 6 times higher than for undoped sapphire. Finally, climb dissociation of the dislocations constituting the perfect dipoles in sapphire is common; annihilation of one set of partials can result in the formation of faulted dipoles, which can pinch off to form faulted dislocation loops. D PU for faulted dipoles in the undoped material was determined at 1300° and 1350°C, and was about 4–10 times higher than for perfect dipoles.  相似文献   

12.
The grain-boundary diffusion product, D'δ , of 51Cr in MgO and Cr-doped MgO as a function of grain-boundary orientation and point-defect concentration was determined at T =1200° to 1450°C. A large degree of anisotropy was found in the grain-boundary diffusion behavior in MgO. The ratio of D'δ|| parallel to D'δ perpendicular to the growth direction, D'||/D' , is 102 for a 5° (100) tilt boundary, decreased to ∼2 in boundaries with tilt angles > 10°. The decrease in D'||/D' is due to a large increase in D' with increasing tilt angle. The results indicate that grain-boundary diffusion in MgO is connected to the orientation of dislocations and the mechanism is one of dislocation pipe diffusion. The grain-boundary diffusion product D'δ increases with increasing Cr concentration in MgO and is ∼4 times larger for MgO containing 0.56 at. % Cr than for the undoped MgO. For all bicrystals studied, the activation energies are within 180 ± 20 kJ/mol which is 60% of the activation energy for 51Cr diffusion in undoped MgO.  相似文献   

13.
The solid solubility of R ions (R = Ho3+, Dy3+, and Y3+) in the BaTiO3 perovskite structure was studied by quantitative electron-probe microanalysis (EPMA) using wavelength-dispersive spectroscopy (WDS), scanning electron microscopy (SEM), and X-ray diffractometry (XRD). Highly doped BaTiO3 samples were prepared using mixed-oxide technology including equilibration at 1400° and 1500°C in ambient air. The solubility was found to depend mainly on the starting composition. In the TiO2-rich samples a relatively low concentration of R incorporated preferentially at the Ba2+ lattice sites (solubility limit ∼Ba0.986R0.014Ti0.9965(V"Ti")0.0035O3at 1400°C). In BaO-rich samples a high concentration of R entered the BaTiO3 structure at the Ti4+ lattice sites (solubility limit ∼BaTi0.85R0.15O2.925(VO••)0.075at 1500°C). Ho3+, Dy3+, and Y3+incorporated preferentially at the Ti4+ lattice sites stabilize the hexagonal polymorph of BaTiO3. The phase equilibria of the Ho3+–BaTiO3 solid solutions were presented in a BaO–Ho2O3–TiO2phase diagram.  相似文献   

14.
In this work, 800 ppm of Zr4+ dopants were added to Al2O3-5 vol% SiC particle composite. Zr4+ doping led to a weak Al2O3 grain-boundary bonding so that the fracture mode changed from transgranular in undoped composite to intergranular in Zr4+-doped composite. The fracture mode change increased the fracture toughness of the composite. Transmission electron microscopy and energy-dispersive spectroscopy examinations revealed that the weak grain-boundary bonding in the doped composite was caused by the segregation of Zr4+ and Si4+ ions at the Al2O3 grain boundary.  相似文献   

15.
The Mode I fracture toughness ( K I C ) of a small-grained Si3N4 was determined as a function of hot-pressing orientation, temperature, testing atmosphere, and crack length using the single-edge precracked beam method. The diameter of the Si3N4 grains was <0.4 µm, with aspect ratios of 2–8. K I C at 25°C was 6.6 ± 0.2 and 5.9 ± 0.1 MPa·m1/2 for the T–S and T–L orientations, respectively. This difference was attributed to the amount of elongated grains in the plane of crack growth. For both orientations, a continual decrease in K IC was observed through 1200°C, to ∼4.1 MPa·m1/2, before increasing rapidly to 7.5–8 MPa·m1/2 at 1300°C. The decrease in K IC through 1200°C was a result of grain-boundary glassy phase softening. At 1300°C, reorientation of elongated grains in the direction of the applied load was suggested to explain the large increase in K IC. Crack healing was observed in specimens annealed in air. No R -curve behavior was observed for crack lengths as short as 300 µm at either 25° or 1000°C.  相似文献   

16.
Ce3+-doped yttrium aluminum garnet (Ce:YAG) phosphor powders were synthesized by heterogeneous precipitation process using three different aluminum sources: α-phase, θ-phase, and boehmite (AlOOH). Mixtures of yttrium and cerium nitrate solutions containing various aluminum sources were precipitated by ammonia solution in normal and reverse strike methods. The influence of pH was studied in the normal strike method by maintaining the solutions at pH 7, 9, and 11 during precipitation. Dried precipitates were double calcined at 1300°C/16 h and 1300–1500°C/24 h, at a ramping rate of 10°C/min, with an intermittent wet ball milling in water. Structural evolution of the resultant phosphors was studied by powder XRD. In the normal strike method, a highly pure YAG phase was formed by α- alumina (pH 7, 11) and θ-alumina (pH 11) while boehmite source ended up with mixed phases of YAlO3 (YAP) and Y4Al2O9 (YAM) along with YAG phase at all pH values of precipitation. However, in the reverse strike process, the θ-phase of alumina gave an extremely pure Ce:YAG phase at a relatively lower calcination temperature (1400°C/24 h) compared with the α-phase and also showed more intense emission of yellowish-green light under blue (λ=469 nm) excitation. Scanning electron microscopy revealed 1–2 μm sized particles with least agglomeration in the reverse strike method.  相似文献   

17.
Beryllium nitride (Be3N2) vaporizes congruently in the range 1640° to 1960°K by the reaction Be, N2( c ) = 3Be( g ) + N2( g ). The equilibrium nitrogen partial pressure, in atmospheres, at the composition for congruent sublimation is given by the expression log P N2= [(–1.952 ± 0.038) × 104] T −1+ (6.509 ± 0.207). The measured enthalpy of decomposition (370 ± 5 kcal at 298° K) yields an enthalpy of formation for Be3N2( c ) of –136 ± 6 kcal/mole. The upper limit to the evaporation coefficient at 1600° to 2000°K can be set as 10–4 by comparison of equilibrium data to Langmuir data obtained with a sample of 18% porosity. The apparent enthalpy of activation for the reaction is 409 ± 7 kcal/mole at 1800°K for the porous Langmuir specimen. An expression is developed to predict the temperature dependence of the reduced apparent pressures in Knudsen studies of substances with low evaporation coefficients in terms of the enthalpy of activation. The variation in temperature dependence of the Langmuir measurements and Knudsen measurements with three different-sized orifices is consistent with predictions from this expression.  相似文献   

18.
Lanthanum zirconate (La2Zr2O7) was prepared by coprecipitating lanthanum nitrate and zirconyl oxychloride at pH 10, followed by ethanol washing. The initial high surface area of ∼304 m2·g−1 decreased very rapidly with increased sintering temperature and decreased to an immeasurably small value after heating at 1200°C for 15 h. The major parameters studied were phase evolution, crystallite size, porosity, surface area reduction, and shrinkage during sintering. Three temperature regions were identified based on these studies: below the crystallization temperature, between the crystallization temperature and ∼1100°C, and above 1100°C. The main contribution of surface area reduction in the region 800°–1100°C was due to surface diffusion; the main contribution above 1100°C was due to grain-boundary diffusion coupled with surface diffusion.  相似文献   

19.
Porous silicon carbide (SiC) membrane supports sintered at 1500°–1800°C were prepared by cold isostatic pressing (CIP) under different pressures and using different amounts of alumina additive (0%–4%). The relationship between processing factors and pore size and microstructure was examined. Varying the sintering temperature, the CIP pressure and the amount of additive used were found to be effective for controlling pore size and microstructure. The pore size and particle size of the membrane support prepared without alumina were found to increase with increasing sintering temperature. This was attributed to surface diffusion. Densification of the undoped support did not occur, however, because of concurrent pore development. In the SiC membrane support containing 4% alumina, small particles and a pore size of around 100 nm were retained. This was because of the formation of a limited amount of SiO2–Al2O3 liquid phase during sintering.  相似文献   

20.
Thermodynamic Properties of the System Indium-Oxygen   总被引:1,自引:0,他引:1  
The free-energy change for the reaction 2In( l )+3NiO( s ) = In2O3( s )+3Ni( s ) was determined from 550° to 800°C from emf measurements on solid-oxide galvanic cells. The results were used to develop an equation for the standard molar free energy of formation of In2O3, i.e. ΔG°ln2O3= -215,550+72.63 T ±450 cal/ mol. The standard molar enthalpy and entropy of formation of In2O3 at 298°K were calculated to be -214,000±1500 cal/mol and -69.03±o0.22 eu, respectively, using the available thermo-chemical data. The absolute entropy of In2O3 at 298°K was calculated to be 32.23±0.22 eu. The free-energy results of this study were used in conjunction with literature data to calculate partial pressures of the gaseous species over In2O3 for different experimental conditions.  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