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 296 毫秒
1.
Subsolidus phase relations in the low-Y2O3 portion of the system ZrO2-Y2O3 were studied using DTA with fired samples and X-ray phase identification and lattice parameter techniques with quenched samples. Approximately 1.5% Y2O3 is soluble in monoclinic ZrO2, a two-phase monoclinic solid solution plus cubic solid solution region exists to ∼7.5% Y2O3 below ∼500°C, and a two-phase tetragonal solid solution plus cubic solid solution exists from ∼1.5 to 7.5% Y2O3 from ∼500° to ∼1600°C. At higher Y2O3 compositions, cubic ZrO2 solid solution occurs.  相似文献   

2.
Phase equilibria of the ZrO2–SmO1.5 system have been studied by X-ray diffraction, scanning electron microscopy, and energy-dispersive X-ray spectroscopy. The compositions of phases in the tetragonal+fluorite, fluorite+pyrochlore, and fluorite+B-Sm2O3 two-phase fields have been determined for samples quenched from temperatures between 1400° and 1700°C. The heat content of the fluorite phase with 30 mol% SmO1.5 and of the pyrochlore phase with 50 mol% SmO1.5 has been measured in the temperature range 200°–1400°C using high-temperature drop calorimetry. The transition between pyrochlore and fluorite phases is clearly first order in the SmO1.5-rich region, while no fluorite+pyrochlore two-phase region has been detected for the samples with ZrO2 excess. Based on the obtained experimental results and literature data, the phase diagram and thermodynamic properties were optimized using the CALPHAD approach.  相似文献   

3.
Phase equilibria in the system ZrO2─InO1.5 have been investigated in the temperature range from 800° to 1700°C Up to 4 mol%, InO1.5 is soluble in t -ZrO2 at 1500°C. The martensitic transformation temperature m → t of ZrO2 containing InO1.5 is compared with that of ZrO2 solid solutions with various other trivalent ions with different ionic radii. The diffusionless c → t ' A phase transformation is discussed. Extended solid solubility from 12.4 ± 0.8 to 56.5 ± 3 mol% InO1.5 is found at 1700°C in the cubic ZrO2 phase. The eutectoid composition and temperature for the decomposition of c -ZrO2 solid solution into t -ZrO2+InO1.5 solid solutions were determined. A maximum of about 1 mol% ZrO2 is soluble in bcc InO1.5 phase. Metastable supersaturation of ZrO2 in bcc InO 1.5 and conditions for phase separation are discussed.  相似文献   

4.
The thermodynamics and kinetics of cubic → tetragonal phase transformations in ZrO2-Y2O3 alloys were investigated by using thermodynamic stability analysis and kinetic computer simulations to explore the possibility of a spinodal mechanism during decomposition. Based on a simple free energy model, it is shown that, depending on the alloy composition, a cubic phase aged within the t + c two-phase field may result in three different sequences of phase transformations: (1) direct nucleation and growth of the equilibrium t-phase from the c-phase matrix; (2) formation of a metastable t'-phase followed by nucleation and growth of the equilibrium c- and t-phases; and (3) formation of a transient t'-phase followed by its spinodal decomposition into two tetragonal phases with one of the tetragonal phases eventually transforming to the equilibrium c-phase. The temporal microstructure evolutions for different compositions were studied by using computer simulations based on the time-dependent Ginzburg-Landau (TDGL) model which incorporates the long-range elastic interactions.  相似文献   

5.
Polycrystalline, Al-doped samples in the TiO2-SnO2 ceramic system were made and annealed in the two-phase field above the [001] coherent spinodal. The samples were examined, after various annealing times, using X-ray diffraction and Raman spectroscopy. It was observed that time constants obtained from X-ray studies were in good agreement with those obtained from Raman shifts, and both showed similar effects of grain size on the decomposition kinetics. These results suggest that Raman spectroscopy may be a useful technique for the study of phase transformations in ceramics.  相似文献   

6.
The subsolidus miscibility gap for the TiO2-SnO2 system was redetermined. The critical temperature, 1430°C, is intermediate between that determined by Padurow, 1330°C, and that determined by Garcia and Speidel, 1475°C. Although the phase boundary is slightly asymmetric (the critical composition occurs at 47 mol% TiO2), it fits the regular-solution model down to 1200°C. Calculations of the coherent spinodal using the regular-solution model indicated depression of the spinodal below Tc , by 105°, 310°, and 387° for composition fluctuations along the [001], <101>, and <100> directions, respectively. These depressions of the spinodal are much greater than those calculated by Stubican and Schultz; this discrepancy is believed to result from an error in the latter workers' calculations. During the present work, positive deviations from Vegard's law were found in this system. Both the magnitude and the sign of the deviation can be predicted using a theory based on nonlinear second-order elasticity.  相似文献   

7.
The phase diagrams in the Al2O3–Cr2O3 and V2O3–Cr2O3 systems have been assessed by thermodynamic modeling with existing data from the literature. While the regular and subregular solution models were used in the Al2O3–Cr2O3 system to represent the Gibbs free energies of the liquid and solid phases, respectively, the regular solution model was applied to both phases in the V2O3–Cr2O3 system. By using the liquidus, solidus, and/or miscibility gap data, the interaction parameters of the liquid and solid phases were optimized through a multiple linear regression method. The phase diagrams calculated from these models are in good agreement with experimental data. Also, the solid miscibility gap and chemical spinodal in the V2O3–Cr2O3 system were estimated.  相似文献   

8.
An optimal set of thermodynamic functions for the ZrO2─YO1.5 system are obtained using phase diagram and thermodynamic data. The liquid is described by a subregular solution model. Both cubic ZrO2 and YO1.5 solid solutions are regarded as one cubic solution, which is also treated as a subregular solution. The ordered Zr3Y4O12 phase is treated as a stoichiometric compound. A regular solution model is applied to the other solid solutions. Tentative equilibrium boundaries between monoclinic and tetragonal ZrO2 solid solutions are evaluated from information about the T 0 line. The calculated phase diagram and thermodynamic functions agree well with experimental data.  相似文献   

9.
Time-index of refraction isotherms were measured for B2O3 glass starting from both a high and a low temperature in the transformation region. The equilibrium index values at each temperature, obtained from both types of approach curve, were identical. As in the case of the density values, the equilibrium refractive index curve as a function of temperature for this glass is not a straight line. The two-relaxation-times (crossover) model was applied to B2O3 glass and fitted the data as well as it did in previous experiments with borosilicate crown and GeO2 glasses. The reverse crossover which was predicted by the model was experimentally confirmed with the B2O3 glass. The spectrum of relaxation times narrowed with decreasing temperature, indicating approach to another region of single relaxation associated with the low-temperature Arrhenius region. The relaxation times for the low-temperature crossover agreed well with those from the high-temperature curves, indicating complete linearity in the experiments. The spectrum of relaxation times was slightly asymmetrical at constant pressure and very asymmetrical at constant volume.  相似文献   

10.
An extensive X-ray study of CeO2–Nd2O3 solid solutions was performed, and the densities of solid solutions containing various concentrations of NdO1.5 were measured using several techniques. Solid solutions containing 0–80 mol% NdO1.5 were synthesized by coprecipitation from Ce(NO3)3 and Nd(NO3)3 aqueous solutions, and the coprecipitated samples were sintered at 1400°C. A fluorite structure was observed for CeO2–NdO1.5 solid solutions with 0–40 mol% NdO1.5, which changed to a rare earth C-type structure at 45–75 mol% NdO1.5. The change in the lattice parameters of CeO2–NdO1.5 solid solutions, when plotted with respect to the NdO1.5 concentration, showed that the lattice parameters followed Vegard's law in both the fluorite and rare earth C-type regions. The maximum solubility limit for NdO1.5 in CeO2 solid solution was approximately 75 mol%. The relationship between the density and the Nd concentration indicated that the defect structure followed the anion vacancy model over the entire range (0–70 mol% NdO1.5) of solid solution.  相似文献   

11.
12.
Stable and metastable phase relationships in the system ZrO2–ErO1.5 were investigated using homogeneous samples prepared by rapid quenching of melts and by arc melting. The rapidly quenched samples were annealed in air for 48 h at 1690°C or for 8 months at 1315°C. Two tetragonal phases ( t - and t '-phases) were observed after quenching samples heated at 1690°C to a room temperature, whereas one t -phase and cubic ( c -) phase were found in those treated at 1315°C. Since the t '-phase is obtained through a diffusionless transformation during cooling from a high-temperature c -phase, t - and c -phases can coexist at high temperature. The t - and c -phases field spans from 4 to 10 mol% ErO1.5 at 1690°C and from 3 to 15 mol% ErO1.5 at 1315°C. The equilibrium temperature T t-m 0 between the t - and monoclinic ( m -) phases estimated from As and Ms temperatures decreased with increasing ErO1.5 contents.  相似文献   

13.
The two-phase field involves a ZrO2-TiO2 solid solution containing no more than 4 mol% TiO2, and a BaTiO3-BaZrO3 solid solution containing a maximum of 75 mol% BaTiO3. The field narrows above 1300°C, probably because of the intrusion of a liquid-phase field into the ternary, beginning near the BaO-TiO2 edge.  相似文献   

14.
Alumina and gallia were substituted separately for Na2O in amounts of 0.2, 0.5, 1.0, 1.5, 2.0, and 3.0 wt% in three Na2O-SiO2 glass compositions (82, 84, and 86 wt% SiO2) within the immiscibility region. The immiscibility regions for each system extend to ∼1.5 mol% of the added oxide. In general, the addition reduced the immiscibility temperature ( T m), but at the edge of the immiscibility region (82% SiO2) the Na2O loss effect initially increased T m. A structural model of the miscibility of Al2O3 added to silicate glasses is presented.  相似文献   

15.
The pseudoternary system ZrO2-YO1.5CrO1.5 was studied between 1300° and 1600C in air by °a quenching method. No ordered phase of the type ZrY6O11 was detected, but an ordered Zr3Y4O12 phase at 1300°C and YCrO3 were observed as intermediate compounds. Solid solutions ofZrO2 and YO1.5 coexisting with CrO1.5 and/or YCrO3 formed; the apex occurred between 26.5 and 27.5 wt% YO1.5 for the cubic ZrO2+CrO1.5+YCrO3, three-phase region; CrO1.5 is slightly soluble in ZrO2(ss).  相似文献   

16.
A partial phase diagram for the system Na3AlF6-Li3AlF6 was constructed from DTA and X-ray diffraction measurements. A region of solid solution extends from Na3AlF6 to the limiting composition Na2LiAlF6. At compositions between Na1.5Li1.5AlF6 and NaLi2AlF6 a cubic phase resembling the mineral cryolithionite is stable over a narrow range of temperatures, which shifts progressively to higher temperatures with increasing lithium content. A single phase can be quenched from samples annealed within this range except near the ideal cryolithionite composition (Na1.5Li1.5AlF6) where the upper temperature limit is too low to permit recombination of two solid phases to cryolithionite even with prolonged annealing. Cryolithionite precipitated from aqueous solution, a sample of the mineral, and the solid solutions containing from 53 to 65 mole % Li3AlF6 have identical powder patterns except for differences of lattice constants. Solid solubility of Na3AlF6 in Li3AlF6 reaches 30 mole % at the eutectic temperature.  相似文献   

17.
NASICON-type structured Li1.5Al0.5Ge1.5(PO4)3– x Li2O Li-ion-conducting glass–ceramics were successfully prepared from as-prepared glasses. The differential scanning calorimetry, X-ray diffraction, nuclear magnetic resonance, and field emission scanning electron microscope results reveal that the excess Li2O is not only incorporated into the crystal lattice of the NASICON-type structure but also exists as a secondary phase and acts as a nucleating agent to considerably promote the crystallization of the as-prepared glasses during heat treatment, leading to an improvement in the connection between the glass–ceramic grains and hence a dense microstructure with a uniform grain size. These beneficial effects enhance both the bulk and total ionic conductivities at room temperature, which reach 1.18 × 10−3 and 7.25 × 10−4 S/cm, respectively. In addition, the Li1.5Al0.5Ge1.5(PO4)3–0.05Li2O glass–ceramics display favorable electrochemical stability against lithium metal with an electrochemical window of about 6 V. The high ionic conductivity, good electrochemical stability, and wide electrochemical window of LAGP–0.05LO glass–ceramics suggest that they are promising solid-state electrolytes for all solid-state lithium batteries with high power density.  相似文献   

18.
The average grain size of ZrO2(+Y, o,) materials sintered at 1400°C was observed to depend significantly on the Y2O3 content. The average grain size decreased by a factor of 4 to 5 for Y2O3 contents between 0.8 and 1.4 mol% and increased at Y2O3 contents of 6.6 mol%. Grain growth control by a second phase is the concept used to interpret these data; compositions with a small grain size lie within the two-phase tetragonal + cubic phase field, and the size of the tetragonal grains is believed to be controlled by the cubic grains. This interpretation suggests that the Y2O3-rich boundary of the two-phase field lies between 0.8 and 1.4 mol% Y2O3. Transformation toughened materials fabricated in this binary system must have a composition that lies within the two-phase field to obtain the small grain size required, in part, to retain the tetragonal toughening agent.  相似文献   

19.
The pseudobinary system CoNb2O5–CoTa2O6 was investigated. CoNb2O6 crystallizes in either the columbite or rutile structure, whereas CoTa2O6 assumes only the trirutile structure. In an argon atmosphere at about 1400°C, CoNb2O6 undergoes a phase transition from columbite or rutile. Between 1000° and 1400°C the solubility of CoTa2O6 in CoNb2O6 is about 10 mole %; in the same temperature region the solubility of CoNb2O6 in CoTa2O6 varies from about 40 to 70 mole %. The extensive solubility of CoNb2O6 in CoTa2O6 is explained by the ability of niobium to induce some disorder in the trirutile phase. The columbite to rutile transformation is also discussed on this basis.  相似文献   

20.
Transmission electron microscopic analyses defined the structures and compositions in single-phase and two-phase La2O3-doped Y2O3 materials fabricated by the transient solid second-phase sintering. The composition in single-phase, 10-mol%-La2O3-doped, sintered and annealed samples was found to be uniform, indicating that diffusivity was sufficiently high for homogenization in the single-phase field. Two-phase, 16-mol%-La2O3-doped, sintered and annealed samples showed two morphologies: (1) intragranular, lath-like, monoclinic second-phase particles (twinned and untwinned) and (2) equiaxed cubic matrix. The second-phase particles were identified as the monoclinic phase derived from the high-temperature hexagonal phase through a rapid phase transition. A short, high-temperature anneal (2200°C for 1 min) of 9 mol% La2O3-Y2O3 composition was found to retain the hexagonal phase. Microchemical analyses of the phases suggested adjustments to the Y2O3-La2O3 phase diagram. Observation of the interactions of the intragranular second-phase particles with crack propagation indicated crack deflection as one of the mechanisms responsible for toughening (1.5 vs 0.9 MPa · m1/2).  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