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 46 毫秒
1.
Effect of a Liquid Phase on the Morphology of Grain Growth in Alumina   总被引:2,自引:0,他引:2  
In this investigation we have studied how the presence of a liquid phase affects the grain morphology and grain growth kinetics in Al2O3 at 1800°C using the growth of both matrix grains and large spherical single-crystal seeds growing into the matrix. The growth rates of the matrix grains were found to decrease in the following order: undoped Al2O3, AI2O3 with anorthite, AI2O3 with anorthite and MgO, and Al2O3 with MgO. Except for the samples doped with MgO alone, the matrix grains were faceted and appeared tabular in polished sections. In samples containing anorthite both with and without MgO, the single-crystal seeds exhibit basal facets with continuous liquid films and slow growth in the 〈0001〉 relative to all other crystallographic directions. When only MgO is added, the growth of the single-crystal seeds was not isotropic; however, no faceting was observed. We discuss how anisotropic growth rates caused by the anorthite additions can stimulate discontinuous grain growth in Al2O3.  相似文献   

2.
A stereological method has been used to determine the degree of grain boundary-pore contact during sintering of Al2O3. Al2O3 doped with 200 ppm MgO exhibits a degree of contact of 5.7 times that expected from random intersections with pores, while pure Al2O3 shows a pore contact factor of 4.8. These data are larger than the values of 2.8 for sintered or hot-pressed UO2, computed from published data, and values of 1.7 and 1.8 for sintered W and Cu powders, respectively. The degree of grain boundary-pore contact for each material remains constant throughout densification from pressed powder to near full density.  相似文献   

3.
Solubility of Magnesia in Polycrystalline Alumina at High Temperatures   总被引:1,自引:0,他引:1  
High-purity Al2O3 compacts were doped with 0–350 ppm (by weight) of MgO using a liquid immersion technique and equilibrated at temperatures between 1700° and 2000°C under hydrogen. The solubility limits of MgO in Al2O3 at temperatures of 1720° and 1880°C were very low, ∼75 and 175 ppm, respectively. Variation of MgO solubility with temperature could be represented by the equation, ln Mg/Al = 3.80–2.63 × 104/ T . The small MgO solubilities were understood by the high enthalpy (326 kJ/mol) of solution. The results of this study suggested that previous investigations on sintering and grain-growth mechanisms in MgO-doped Al2O3 were probably not done in single-phase Al2O3 solid solutions. However, the conclusions on sintering and grain-growth mechanisms in prior research work in MgO-doped A2O3 may be correct. The effects of SiO2 impurity and grain size on MgO solubility are discussed. Previous grain-growth experiments in MgO-doped Al2O3 are described that demonstrate the clearest evidence for grain-boundary mobility controlled by a solid-solution mechanism.  相似文献   

4.
Solubility of NiO in Al2O3 was determined by electron probe microanalysisy A diffusion couple method was used by coupling an NiO-doped Al2O3 polycrystal to a pure single crystal of Al2O3. The solubility of NiO in Al2O3 in air was 230 wt ppm (157 at. ppm of cations) and 170 wt ppm (116 at. ppm) at 2073 and 1973 K, respectively. The solubility of NiO in Al2O3 obtained in this work was compared with our previous work of the solubility of MgO in Al2o3.  相似文献   

5.
In a given batch more than 30%–40% of polycrystalline, MgO-doped Al2O3 tubes were converted into single crystals of sapphire by abnormal grain growth (AGG) in the solid state at 1880°C. Most crystals grew 4–10-cm in length in tubes with wall thicknesses of 1/2 and 3/4 mm and outer diameters of 5 and 7 mm, respectively, and had their c -axes oriented ∼ 90° and 45° to the tube axis. Initiation of AGG was associated with low values of bulk MgO concentration near 50 ppm. The unconverted tubes did not develop centimeter-size crystals but instead exhibited millimeter-size grains. The different grain structures in converted and unconverted tubes may be related to nonuniform concentration of MgO in the extruded tubes. The growth front of the migrating crystal boundary was typically nonuniformly shaped, and the interface between the single crystal and the polycrystalline matrix was composed of many "curved" boundary segments indicative of classical AGG in a single-phase material. The average velocities of many migrating crystal boundaries were quite high and reached ∼1.5 cm/h. The average grain boundary mobility at 1880°C was calculated as 2 × 10−10 m3/(N·s), representing the highest value reported so far in Al2O3 and within a factor of 2.5 of the calculated intrinsic mobility. Under similar experimental conditions sapphire crystals did not grow when a codopant of CaO, La2O3, or ZrO2 was added in concentrations of several hundred ppm.  相似文献   

6.
The grain-growth behavior of Al2O3 compacts with small contents (≤10 wt%) of various liquid-forming dopants was studied. Equiaxed and/or elongated grains were observed for the following dopants: MgO, CaO, SiO2, or CaO + TiO2. The platelike grains, defined as the abnormal grains larger than 100 μm with an aspect ratio ≥5 and with flat boundaries along the long axis, were observed when the boundaries were wet with the liquid phase and the codoping satisfied two conditions of size and valence. These dopings were Na2O + SiO2, CaO + SiO2, SrO + SiO2, or BaO + SiO2. However, an addition of MgO to the Al2O3 doped with CaO + SiO2 resulted in the change of grain shape from platelike to equiaxial. Equiaxed grains were also observed for the MgO + SiO2 doping, indicating that two conditions were necessary but not sufficient to develop the platelike grains. The fast growth rate of the platelike grains was explained by an increased interfacial reaction rate due to the codopants. AT the same time the codopants made the basal plane, which appeared as the flat boundaries, the lowest energy plane. The appearance of the platelike grains was favored in compacts with a small grain size and with a narrow size distribution at the onset of abnormal grain growth. Accordingly, the use of starting powders with a small particle size and narrow size distribution, smaller amounts of dopings, and high sintering temperature resulted in an increased number of the platelike grains.  相似文献   

7.
Grain-Growth Kinetics for Alumina in the Absence of a Liquid Phase   总被引:3,自引:0,他引:3  
The kinetics of grain growth in fully dense Al2O3 with and without MgO solute additions were measured for high-purity samples containing no liquid phases. The MgO was found to suppress the grain-boundary migration rate by a factor of 50. Compensating lattice defects are suggested to play a role in grain-growth inhibition. Implications of these results to the sintering of Al2O3 are discussed.  相似文献   

8.
AlN–AlN polytypoid composite materials were prepared in situ using pressureless sintering of AlN–Al2O3 mixtures (3.7–16.6 mol% Al2O3) using Y2O3 (1.4–1.5 wt%) as a sintering additive. Materials fired at 1950°C consisted of elongated grains of AlN polytypoids embedded in equiaxed AlN grains. The Al2O3 content in the polytypoids varied systematically with the overall Al2O3 content, but equilibrium phase composition was not established because of slow nucleation rate and rapid grain growth of the polytypoid grains. The polytypoids, 24 H and 39 R , previously not reported, were identified using HRTEM. Solid solution of Y2O3 in the polytypoids was demonstrated, and Y2O3 was shown to influence the stability of the AlN polytypoids. The present phase observations were summarized in a phase diagram for a binary section in the ternary system AlN–Al2O3–Y2O3 parallel to the AlN–Al2O3 join. Fracture toughness estimated from indentation measurements gave no evidence for a strengthening mechanism due to the elongated polytypoids.  相似文献   

9.
The effect of MgO and ZrO2 dopants, added separeately or simultaneously, on the grain size, denisity, and toughness of Al2O3 was studied. Small ZrO2 addotoopms (<100 ppm) had little effect, whereas larger amounts decreased the sintered density. Additions of Mgo Up to the solubility limit (∼ 300 ppm) increased both density and grain size; further additions had little effect on the density but strongly reduced grain size.  相似文献   

10.
Hard lead zirconate titanate (PZT) and PZT/Al2O3 composites were prepared and the alternating-electric-field-induced crack growth behavior of a precrack above the coercive field was evaluated via optical and scanning electron microscopy. The crack extension in the 1.0 vol% Al2O3 composite was significantly smaller than that in monolithic PZT and the 0.5 vol% Al2O3 composite. Secondary-phase Al2O3 dispersoids were found both at grain boundaries and within grains in the composites. A large number of dispersoids were observed at the grain boundaries in the 1.0 vol% Al2O3 composite. It appears that the Al2O3 dispersoids reinforce the grain boundaries of the PZT matrix as well as act as effective pins against microcrack propagation.  相似文献   

11.
Final-stage sintering has been investigated in ultrahigh-purity Al2O3 and Al2O3that has been doped individually with 1000 ppm of yttrium and 1000 ppm of lanthanum. In the undoped and doped materials, the dominant densification mechanism is consistent with grain-boundary diffusion. Doping with yttrium and lanthanum decreases the densification rate by a factor of ˜11 and 21, respectively. It is postulated that these large rare-earth cations, which segregate strongly to the grain boundaries in Al2O3, block the diffusion of ions along grain boundaries, leading to reduced grain-boundary diffusivity and decreased densification rate. In addition, doping with yttrium and lanthanum decreases grain growth during sintering. In the undoped Al2O3, surface-diffusion-controlled pore drag governs grain growth; in the doped materials, no grain-growth mechanism could be unambiguously identified. Overall, yttrium and lanthanum decreases the coarsening rate, relative to the densification rate, and, hence, shifted the grain-size-density trajectory to higher density for a given grain size. It is believed that the effect of the additives is linked strongly to their segregation to the Al2O3grain boundaries.  相似文献   

12.
The kinetics of grain growth in fully dense Al2O3 with and without MgO solute additions were measured. The MgO was found to retard boundary migration and to rkvelop more uniform microstructures. A grain-growth mechanism involving solute partitioning of segregated ions (calcium and magnesium) between different boundary types is proposed. Implications regarding the sintering of Al2O3 are discussed.  相似文献   

13.
The complicated texture of a thin sheet of MgO and MgO-doped Al2O3 (∼15 pm) was statistically resolved into four typical grain geometries. Grain size distribution was related to the ratio of expressions describing the growth (or shrinkage) of individual grains and the growth of grains of average size R. Grain-growth data for MgO indicate that the mobility of a grain boundary is independent of R (parabolic-grain-growth behavior) but depends strongly on the ratio of individual grain size to average grain size. When mobility is proportional to 1/R, the time dependence of grain growth is <1/2.  相似文献   

14.
The infiltration of glass melts into fully dense Al2O3 and MgO ceramics has been studied with emphasis on elucidating the penetration mechanism and the change in shape and size of the solid grains that accompany the penetration process. For Al2O3, penetrated by a Ca—Al—Si—O glass melt, the grains developed a prismatic shape consistent with interface-reaction-controlled grain growth. For MgO, penetrated by a Ca—Mg—Si—O glass melt, the grains maintained a spherical shape consistent with diffusion-controlled grain growth. When glass penetrated into the dense polycrystalline alumina specimen, it resulted in a homogeneous distribution of liquid phase and a uniform grain size throughout the whole specimen. In contrast, when glass penetrated the magnesia specimen, the volume fraction of liquid phase at the surface region (which was in direct contact with the melt) was higher than that in the center region. Furthermore, the average grain size was larger in the center, where the volume fraction of glass was lower. This microstructural inhomogeneity stayed uncorrected even after prolonged annealing treatments. Reasons for this behavior are discussed.  相似文献   

15.
The microstructures of niobium-based alumina composites prepared by pressureless sintering of compacts of attrition milled Al2O3, Nb, and Al powder mixtures were studied. The addition of a small amount of Al is assumed to assist in rapid sintering. X-ray diffraction analyses show that Al2O3, Nb, NbO, and the intermetallics AlNb2 and AlNb3 are present in the composites. Electron microscopy studies confirm the existence of these phases and reveal dense, fine-grained (≤500 nm) composites. Al2O3 and Nb grains form the matrix. NbO occurs as grains and additionally as small particles within Al2O3 grains and at Al2O3/Al2O3 grain boundaries. The intermetallic AlNb2 and AlNb3 phases do not exceed 300 nm in size if they occur at grain boundaries, and possess even smaller dimensions when occluded within Al2O3 grains or located at Al2O3 triple junctions. While the niobium intermetallics are expected to form during the heating cycle before reaching the sintering temperature, the NbO is assumed to form during the cooling cycle due to precipitation of oxygen dissolved in the niobium.  相似文献   

16.
The thermal expansion of Al2O3–MgO castables containing 5.5 wt% MgO and 1.36 wt% CaO and Al2O3–spinel castables containing 20 wt% spinel having 95 wt% Al2O3 and 1.7 wt% CaO was measured in the temperature range of 800–1650°C by dilatometry. A sharp increase in expansion from around 1425° to 1525°C, followed by a sharp decrease with further increasing temperature, is characteristic of Al2O3–MgO castables. The sharp increase in expansion is believed to be caused by the bond linkage between the CA6 and spinel grains in the bonding matrix, while the sharp decrease is apparently related to liquid-phase sintering. The sharp increase and decrease in expansion were not observed in Al2O3–spinel castables because of the much lower MgO (around 1 wt% MgO) and impurity contents. The magnitude of thermal expansion of calcium aluminate bonded castables containing self-forming or preforming spinels or both is dictated by the MgO content of the castables.  相似文献   

17.
Porous Al2O3 and SiC–dispersed-Al2O3 (Al2O3/SiC) nanocomposites with improved mechanical properties were fabricated using pulse electric current sintering (PECS). Microstructures with fine grains and enhanced neck growth, as well as high fracture strength, could be achieved via PECS of Al2O3. The incorporation of fine SiC particles into an Al2O3 matrix significantly increased the fracture strength of porous Al2O3. Based on microstructural observations, it was revealed that the refinement of Al2O3 grains and neck growth occurred by PECS and nanocomposite processing.  相似文献   

18.
The scavenging of a resistive siliceous phase via the addition of Al2O3 was studied, using imaging secondary-ion mass spectroscopy (SIMS), given the improved grain-boundary conductivity in 8-mol%-yttria-stabilized zirconia (8YSZ). The grain-boundary resistivity in 8YSZ decreased noticeably with the addition of 1 mol% of Al2O3. Strong SiO2 segregation at the grain boundaries was observed in a SIMS map of pure 8YSZ that contained 120 ppm of SiO2 (by weight). The addition of 1 mol% of Al2O3 caused the SiO2 to gather around the Al2O3 particles. The present observations provided direct and visual evidence of SiO2 segregation at the grain boundaries (which had a deleterious effect on grain-boundary conductivity) and the scavenging of SiO2 via Al2O3 addition.  相似文献   

19.
Compositions of alumina with a molybdenum dispersed phase were investigated in the 0 to 5 vol% Mo range. These compositions were also prepared with a 0.5 wt% MgO addition. All specimens were fabricated by hot-pressing, and near theoretical densities were achieved. Specimens were characterized by metallographic and X-ray diffraction analyses, and microhardness, elastic moduli, tensile strength, and fracture energy were determined. Results revealed that Mo additions did not affect grain growth; in contrast, MgO additions significantly inhibited grain growth. However, Mo additions did reduce the elastic moduli and microhardness but did not measurably affect the tensile strength. Tensile strength was dependent on grain size and fitted the G−1/3 relation. The fracture energy of Al2O3+5% Mo was 50% greater than that of Al2O3. Specimens were successfully hot-pressed with a micro-structure graded from that of Al2O3 to that of the 5% Mo composition.  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