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 15 毫秒
1.
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   

2.
The phase relations of the systems ZrO2–TiO2 and ZrO2–TiO2–SiO2 were investigated. X-ray diffraction techniques served as the principal means of analysis. The binary system ZrO2–TiO2 was found to be one of partial solid solutions with no intermediate compounds. A eutectic point was found to exist at 50 to 55 weight % ZrO2 and 1600°C. A preliminary investigation of the ternary system ZrO2–TiO2–SiO2, although not extensive, resulted in a better understanding of this system, with a fairly accurate location of some of its boundary lines. A eutectic point was located at 2% ZrO2, 10% TiO2, and 88% SiO2 at approximately 1500°C.  相似文献   

3.
Studies made on low-hafnium-content ZrO2, show that the monoclinic-tetragonal inversion temperature is 1170°C., and it is raised to approximately 1190°C. in the "natural" ZrO2, which contains approximately 2% HfO2. No explanation could be found for the knownmarked hysteresis during cooling, when the reverse polymorphic transformation takes dace at 1040°C. In the system ZrO2-ThO2 the monoclinic-tetragonal ZrO2, inversion temperature is lowered to 1000°C., although the maximum solid solution extent of ZrO2, in Thon and vice versa is approximately only 2% at this temperature. Below about 400°C. under hydrothermal conditions it was possible to prepare a continuous, although metastable series of solid solutions with the fluorite structurewith compositions varying from ThO2, to nearly pure ZrO2. Contrary to earlier work only 8 mole ZrO2, dissolves in UO2 and less than 4 mole of UO, in ZrO2 at temperatures up to 13OO0C. A continuous series of solid solutions could be made between Th2 and UO2 at 13OO°C., and extensive defect fluorite solid solutions could be prepared between Tho2 and U3O8; there is some evidence for exsolution into uranium-rich and thorium-rich members at low temperatures.  相似文献   

4.
The effect of ZrO2 on crystallographic order, microstructure, and microwave dielectric properties of Ba(Zn1/3Ta2/3)O3 (BZT) ceramics was investigated. A small amount of ZrO2 disturbed the 1:2 cation ordering. The average grain size of the BZT significantly increased with the addition of ZrO2, which was attributed to liquid-phase formation. The relative density increased with the addition of a small amount of ZrO2, but it decreased when the ZrO2 content was increased. Variation of the dielectric constant with ZrO2 addition ranged between 27 and 30, and the temperature coefficient of resonant frequency increased abruptly as the ZrO2 amount exceeded 2.0 mol%. The Q value of the BZT significantly improved with the addition of ZrO2, which could be explained by the increased relative density and grain size. The maximum Q × f value achieved in this investigation was ∼164 000 GHz for the BZT with 2.0 mol% ZrO2 sintered at 1550°C for 10 h.  相似文献   

5.
The equilibrium temperature ( T 0) at which the tetragonal and monoclinic phases of either ZrO2 or HfO2 coexist is generally defined by the middle temperature of A s (the onset transformation temperature on heating) and M s (the onset transformation temperature on cooling). It cannot be directly determined due to the athermal nature of the martensitic transformation. Practically, the determination of T 0 is important for the prediction of A s and M s in ZrO2 or HfO2-based materials. In this work, the ZrO2–HfO2 system was studied experimentally by differential thermal analysis (DTA) to obtain the martensitic tetragonal ⇔ monoclinic transformation temperatures in the temperature range of 1273–1973 K. The T 0 temperatures obtained for ZrO2 and HfO2 are 1367±5 and 2052±5 K, respectively. They are adopted for the assessments of the Gibbs energy parameters of these two oxides. A reasonably calculated ZrO2−HfO2 phase diagram is presented.  相似文献   

6.
During constant-rate heating to 350°C in concentrated NaOH solutions, cubic ZrO2 crystallized at ∼120°C from hydrated amorphous ZrO2; these One cubic ZrO2 particles abruptly changed into needlelike monoclinic ZrO2 single crystals at 300°C. Crystallization and phase transformation were studied by XRD, TEM, and EPMA. Cubic ZrO2 appears to crystallize via collapse of the ZrO2−nH2O structure and subsequent slight rearrangement of the lattice. The abrupt formation of mono-clinic ZrO2 was considered to result when the very fine cubic ZrO2 particles coagulated in a highly oriented fashion.  相似文献   

7.
A ZrO2 coating was prepared on Hi-Nicalon fiber and single-crystal Si by chemical vapor deposition (CVD) using ZrCl4, CO2, and H2 as precursors at 1050°C. The effects of oxygen partial pressure on the nucleation behavior of the CVD-ZrO2 coating were systematically studied by intentionally varying the controlled amount of O2 into the CVD chamber. Characterization results suggested that the number density of tetragonal ZrO2 nuclei apparently decreased with increasing the oxygen partial pressure from 4 × 10−3 to 1.6 Pa. Also, the coating layer became more columnar and contained larger monoclinic ZrO2 grains. The observed relationships between the oxygen partial pressure and the nucleation and morphologic characteristics of the ZrO2 coating were attributed to the grain size and oxygen deficiency effects, which have been previously reported to cause the stabilization of the tetragonal ZrO2 phase in bulk ZrO2 specimens.  相似文献   

8.
The vertical section Ti-ZrO2 within the Ti-Zr-O system was investigated by metallographic, X-ray diffraction, electron probe, and melting point studies. Analyses were conducted using arcmelted specimens which had been equilibrated and quenched from temperatures of 600° to 1600°C. The Ti-ZrO2 section is similar to the Zr-ZrO2 system. At high temperatures, considerable amounts of Zr and O go into solid solution in Ti, stabilizing α-Ti to 30 wt% ZrO2. From 30 to 98 wt% ZrO2 an α-Ti+ZrO2 region is defined, and at compositions above 98 wt% ZrO2, single-phase ZrO2( ss ) exists. At low temperatures an α-Ti+(Ti,Zr)3O field exists from 22 to 32 wt% ZrO2; this region decreases in size with increasing temperature until it disappears at 1200°C. Above 32 wt% ZrO2, a three phase α-Ti+ (Ti,Zr)3O+ZrO2 field exists; its stability extends from 1200°C at 30 wt%   相似文献   

9.
The oxidation of hot-pressed SiC-particle (SiCp)/zirconia (ZrO2)/mullite composites with various ZrO2 contents, exposed in air isothermally at 1000° and 1200°C for up to 500 h, was investigated; an emphasis was placed on the effects of the ZrO2 content on the oxidation behavior. A clear critical volume fraction of ZrO2 existed for exposures at either 1000° or 1200°C: the oxidation rate increased dramatically at ZrO2 contents of >20 vol%. The sharp transition in the oxidation rate due to the variation of ZrO2 content could be explained by the percolation theory, when applied to the oxygen diffusivity in a randomly distributed two-phase medium. Morphologically, the composites with ZrO2 contents greater than the critical value showed a large oxidation zone, whereas the composites with ZrO2 contents less than the critical value revealed a much-thinner oxidation zone. The results also indicated that the formation of zircon (ZrSiO4) at 1200°C, through the reaction between ZrO2 and the oxide product, could reduce the oxidation rate of the composite.  相似文献   

10.
A reexamination of the system CaO·SiO2–ZrO2 has been conducted in order to use this system to obtain ZrO2-toughened wollastonite materials. The results have shown that CaO·SiO2–ZrO2 is a pseudobinary system with an invariant peritectic point at 1467°± 2°C, which is in disagreement with previously reported results.  相似文献   

11.
Fracture toughness of ZrO2-toughened alumina could he increased by macroscopic interfaces, such as those existing in laminated composites. In this work, tape casting was used to produce A/A or A/B laminates, where A and B can be Al2O3, Al2O3/5 vol% ZrO2, and Al2O3/l0 vol% ZrO2. An increase of toughness is observed, even in the Al2O3/Al2O3 laminates.  相似文献   

12.
Metastable tetragonal ZrO2 phase has been observed in ZrO2–SiO2 binary oxides prepared by the sol–gel method. There are many studies concerning the causes of ZrO2 tetragonal stabilization in binary oxides such as Y3O2–ZrO2, MgO–ZrO2, or CaO–ZrO2. In these binary oxides, oxygen vacancies cause changes or defects in the ZrO2 lattice parameters, which are responsible for tetragonal stabilization. Since oxygen vacancies are not expected in ZrO2–SiO2 binary oxides, tetragonal stabilization should just be due to the difficulty of zirconia particles growing in the silica matrix. Furthermore, changes in the tetragonal ZrO2 crystalline lattice parameters of these binary oxides have recently been reported in a previous paper. The changes of the zirconia crystalline lattice parameters must result from the chemical interactions at the silica–zirconia interface (e.g., formation of Si–O–Zr bonds or Si–O groups). In this paper, FT-IR and 29Si NMR spectroscopy have been used to elucidate whether the presence of Si–O–Zr or Si–O is responsible for tetragonal phase stabilization. Moreover, X-ray diffraction, Raman spectroscopy, and transmission electron microscopy have also been used to study the crystalline characteristics of the samples.  相似文献   

13.
A furnace for use in conjunction with the X-ray spectrometer was developed which was capable of heating small powdered specimens in air to temperatures as high as 1850°C. This furnace was also used for the heating and quenching of specimens in air from temperatures as high as 1850°C. An area of two liquids coexisting between 20 and 93 weight % TiO2 above 1765°± 10°C. was found to exist in the system TiO2–SiO2, which is in substantial agreement with the previous work of other investigators. The area of immiscibility in the system TiO2–SiO2 was found to extend well into the system TiO2–ZrO2–SiO2. The two liquids were found to coexist over a major portion of the TiO2 (rutile) primary-phase area with TiO2 (rutile) being the primary crystal beneath both liquids. The temperature of two-liquid formation in the ternary was found to fall about 80°C. with the first additions of ZrO2 up to 3%. With larger amounts of ZrO2 the change in the temperature of the boundary of the two-liquid area was so slight as to be within the limits of error of the temperature measurement. Primary-phase fields for TiO2 (rutile), tetragonal ZrO2, and ZrTiO4 were found to exist in the system TiO2–ZrO2–SiO2. SiO2 as high cristobalite is known to exist in the system TiO2–ZrO2–SiO2.  相似文献   

14.
In order to determine the effect of the coarse tail of the ZrO2 size distribution of the monoclinic ZrO2 content of zirconia-toughened alumina, quantitative hot-stage X-ray measurements were made on (1) A12O3-10% Zr02 (submicrometer with a coarse tail >1 μm), (2) A12O3-10% ZrO2 with the coarse tail of the ZrO2 carefully removed by centrifugation, and (3) Al2O3-10% ZrO2 in which 10% of Zr02(1) and 90% of Zr02(2) were used. The monoclinic content of (3) was compared with the weighted average of(1) and (2). A disproportionate amount of monoclinic ZrO2 was found in (3) and it was concluded that transformation of the coarse particles promotes transformation in finer particles. Microstructural examination supports this conclusion. Results were interpreted as being due both to autocatalytic transformation and to constraint relief from microcracking.  相似文献   

15.
The transformation of ultrafine powders (particle size, 0.01 to 0.04 μm) of the system ZrO2–Al2O3, prepared by spraying their corresponding nitrate solutions into an inductively coupled plasma (ICP) of ultrahigh temperature, was investigated. The powders were composed of metastable tetragonal ZrO2 ( mt- ZrO2) and γ-Al2O3. On heating, the mt- ZrO2 (or tetragonal ZrO2, t -ZrO2) was retained up to 1200°C. At 1380°C the transformation to monoclinic ZrO2 ( m -ZrO2) occurred and the amount of the m -ZrO2 decreased with the increase in Al2O3 content, thus indicating the stabilization of the t -ZrO2 by the Al2O3, which seems to be explained in terms of the retardation of grain growth.  相似文献   

16.
This paper clarifies the formation reaction of ZrO2 crystals which appear as extrinsic scatterers in fluoride fibers. EPMA analysis indicates that BaO exists at grain boundaries of BaF2 purified by sublimation. BaO reacts with ZrF4 to form ZrO2 at 600°C during a glass-melting process. The ZrO2 formation reaction is influenced by H2O. Ba(OH)2, which is formed by the reaction between BaO and water vapor, melts at 370° to 420°C and reacts with ZrF4 to form ZrO2 at 450° to 520°C. When low-oxide-content BaF2 is used for fiber preparation, scatterers significantly decrease.  相似文献   

17.
Mullite composites toughened with ZrO2 (with or without a MgO or Y2O3 stabilizer) and/or SiC whiskers (SiC( w )) were fabricated by hot-pressing powders prepared from Al, Si, Zr, and Mg(Y) alkoxide precursors by a sol–gel process. Micro-structures were studied by using XRD. SEM, and analytical STEM. Pure mullite samples contained prismatic, preferentially oriented mullite grains. However, the addition of ZrO2, as well as the hot-pressing temperature, affected the morphology and grain size in the composites; a fine, uniform, equiaxed microstructure was obtained. The effect of SiC( W ) was less pronounced than that of ZrO2. Glassy phases were present in mullite and mullite/SiC( W ) composites, but were rarely observed in Al2O3-rich or ZrO2-containing samples. The formation of zircon due to the reaction between ZrO2 and SiO2 and the considerable solid solution of SiO2 in ZrO2 prevented the formation of the glassy phase, whereas the reaction between Al2O3 and MgO in MgO-containing samples formed a spinel phase and also deprived the ZrO2 phase of the stabilizer. Intergranular ZrO2 particles were either monoclinic or tetragonal, depending on size and stabilizer content; small intragranular ZrO2 inclusions were usually tetragonal in structure.  相似文献   

18.
Infrared absorption spectra are shown for monoclinic ZrO2 and for cubic stabilized ZrO2. Nine bands are reported in monoclinic ZrO2 in the region 800 to 200 cm−l, whereas only one broad band is observed in cubic ZrO2 over the same frequency range.  相似文献   

19.
The control of the microstructure of Ce-doped Al2O3/ZrO2 componsites by the valence change of cerium ion has been demonstrated. Two distinctively different types of microstructure, large Al2O3 grains with intragranular ZrO2 particles and small Al2O3 grains with intergranular ZrO2 particles, can be obtained under identical presintering processing conditions. At doping levels greater than ∼ 3 mol% with respect to ZrO2, Ce3+ raises the alumina grain-boundary to zirconia particle mobility ratio. This causes the breakaway of grain boundary from particles and the first type of microstructure. On the other hand, Ce4+ causes no breakaway and produces a normal intergranular ZrO2 distribution. The dramatic effect of Ce3+ on the relative mobility ratio is found to be associated with fluxing of the glassy boundary phase and is likewise observed for other large trivalent cation dopants. The ZrO2 second phase acts as a scavenger for these trivalent cations, provided their solubility limit in ZrO2 is not exceeded.  相似文献   

20.
Alumina and Al2O3/ZrO2 (1 to 10 vol%) composite powders were mixed and consolidated by a colloidal method, sintered to >98% theoretical density at 1550°C, and subsequently heat-treated at temperatures up to 1700°C for grain-size measurements. Within the temperature range studied, the ZrO2 inclusions exhibited sufficient self-diffusion to move with the Al2O3 4-grain junctions during grain growth. Growth of the ZrO2, inclusions occurred by coalescence. The inclusions exerted a dragging force at the 4-grain junctions to limit grain growth. Abnormal grain growth occurred when the inclusion distribution was not sufficiently uniform to hinder the growth of all Al2O3 grains. This condition was observed for compositions containing ≤2.5 vol% ZrO2, where the inclusions did not fill all 4-grain junctions. Exaggerated grains consumed both neighboring grains and ZrO2, inclusions. Grain-growth control (no abnormal grain growth) was achieved when a majority (or all) 4-grain junctions contained a ZrO2 inclusion, viz., for compositions containing ≥5 vol% ZrO2. For this condition, the grain size was inversely proportional to the volume fraction of the inclusions. Since the ZrO2 inclusions mimic voids in all ways except that they do not disappear, it is hypothesized that abnormal grain growth in single-phase materials is a result of a nonuniform distribution of voids during the last stage of sintering.  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