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 515 毫秒
1.
Microhardness anisotropy profiles for the (100) and (111) planes of single-crystal stoichiometric MgAl2O., spinel were determined at room temperaturé. The (100) microhardness profile has ahardness maximum in tiie [001] and a minimum in the [O11], which supports the previous suggestion that the primary slip system is the {111}〈11¯0〉. The microhardness of the (111) plane is independent of indenter orientation, also consistent, with a {111}〈11¯0〉 primary slip system. It is concluded that these microhardness profiles are in accord with other experimental observations that the {111}〈11¯0〉 is the primary slip system in stoichiometric MgAl2O4 spinel.  相似文献   

2.
Sintered AIN specimens were deformed by Vickers hardness (HV) indentations. Compared with Al2O3 the HV hardness values indicate a much higher plasticity of AIN at room temperature, but above 600°C a higher ductility for Al2O3. Deformed AIN specimens were examined by transmission electron microscopy. Basal and prismatic glide with the slip systems (0001) 〈1120〉 and {1 1 00}〈11 2 0〉 were frequently observed. This results in four linearly independent slip systems. The critical resolved shear stress for single prismatic slip seems to be even smaller than for basal slip. However, thermally activated dislocation reactions are frozen up to at least 1000°C. Thus, prismatic slip is suppressed as soon as more than one slip direction is activated.  相似文献   

3.
Bulk BaTiO3 ceramics with 〈111〉-texture have been prepared by the modified templated grain growth method, using platelike Ba6Ti17O40 particles as templates, and the mechanism of texture development is examined. The Ba6Ti17O40 particles induce the abnormal growth of BaTiO3 grains, and a structure similarity between {001} of Ba6Ti17O40 and {111} of BaTiO3 gives 〈111〉-texture to abnormally grown BaTiO3 grains. Thus, the 〈111〉-texture develops in the BaTiO3 matrix. The use of platelike Ba6Ti17O40 particles has been extended to a 0.65Pb(Mg1/3Nb2/3)O3–0.35PbTiO3 matrix, but the matrix phase is decomposed by extensive chemical reactions between the matrix and template phases.  相似文献   

4.
Alpha alumina whiskers grown by the vapor reaction 2A1 + 3H2O° Al2O3+ 3H2 were examined by X-ray microtopography. Topographs of whiskers PO to 60 μm in diameter were taken by (1120) and (00012) diffraction; most a and c whiskers (growth direction 〈1120〉 and 〈0001〉, respectively) were free from dislocations. Chemical etching of a whisker cross section with orthophosphoric acid or potassium bisulfate did not produce etch pits. Growth of alumina whiskers through axial dislocation appears improbable.  相似文献   

5.
During wetting experiments of pure aluminum on the (0001) plane of sapphire, an aluminum sample was unintentionally contaminated with 31 atoms/parts per million of iron. Transmission electron microscopy investigations of the sample after cooling showed the formation of FeAlO3 at the aluminum-Al2O3 interface. Formation of FeAlO3 confirms the existence of a high oxygen activity at the liquid aluminum-sapphire interface, which is considered to be the reason for the strong adhesion between aluminum and Al2O3. A low indexed orientation relationship between FeAlO3 and Al2O3 was determined, where [22¯1]FeAlO3‖ [11¯00]Al2O3 and (110)FeAlO3‖ (0001)Al2O3.  相似文献   

6.
The entire data base for the dependence of the nonstoichiometry, x , on temperature and chemical potential of oxygen (oxygen potential) in 〈U1– z Ce z O2+ x 〉 and 〈U1– z Ln z O2+ x 〉 was retrieved from the literature and represented by a thermodynamic method. The method reproduces the behavior of the experimental data and also results in partial molal Gibbs free energy quantities that are useful for any thermodynamic calculation involving these nonstoichiometric phases. The behavior of these systems is also compared with that for 〈U1– z Pu z O2± x 〉.  相似文献   

7.
Stoichiometric polycrystalline magnesium aluminate spinel has been irradiated at 25° and 650°C with 2.4-MeV Mg+ ions to a fluence of 1.4 × 1021 ions/m2 (∼35 dpa (displacement per atom) peak damage level). Microindentation hardness measurements and transmission electron microscopy combined with energy dispersive X-ray spectroscopy measurements were used to characterize the irradiation effects. The room-temperature hardness of spinel increased by about 5% after irradiation at both temperatures. There was no evidence for amorphization at either irradiation temperatures. Interstitial-type dislocation loops lying on {110} and {111} planes with Burgers vectors along 〈110〉 were observed at intermediate depths (∼1 μm) along the ion range. The 〈110〉{111} loops are presumably formed from 〈111〉{111} loops as a result of a shear on the anion sublattice. Only about 0.05% of the calculated displacements were visible in the form of loops, which indicates that spinel has a high resistance to aggregate damage accumulation. The peak damage region contained a high density of dislocation tangles. There was no evidence for the formation of voids or vacancy loops. The specimen irradiated at 650°C was denuded of dislocation loops within ∼1 μm of the surface.  相似文献   

8.
We report a nanobelt structure along the 〈110〉 direction in perovskite–spinel (BiFeO3–CoFe2O4 or BFO–CFO) composite epitaxial thin films deposited on SrTiO3 (STO) substrates by pulsed laser deposition. The results reveal the nucleation and growth mechanism of a nanostructure: first nucleation of perovskite phase regions on STO, followed by segregation of spinel out of the perovskite phase, and finally by evolution of spinel nanobelts that are elongated along the 〈110〉. The reason for control of the nanostructure evolution is a ledge growth mechanism of spinel along the 〈110〉 direction.  相似文献   

9.
Single-crystal and polycrystalline films of Mg-Al2O4 and MgFe2O4 were formed by two methods on cleavage surfaces of MgO single crystals. In one procedure, aluminum was deposited on MgO by vacuum evaporation. Subsequent heating in air at about 510°C formed a polycrystalline γ-Al2O8 film. Above 540°C, the γ-Al2O, and MgO reacted to form a single-crystal MgAl2O4 film with {001} MgAl2O4‖{001} MgO. Above 590°C, an additional layer of MgAl2O4, which is polycrystalline, formed between the γ-Al2O3 and the single-crystal spinel. Polycrystalline Mg-Al2O4 formed only when diffusion of Mg2+ ions proceeded into the polycrystalline γ-Al2O3 region. Corresponding results were obtained for Mg-Fe2O4. MgAl2O4 films were also formed on cleaved MgO single-crystal substrates by direct evaporation, using an Al2O3 crucible as a source. Very slow deposition rates were used with source temperatures of ∼1350°C and substrate temperatures of ∼800°C. Departures from single-crystal character in the films may arise through temperature gradients in the substrate.  相似文献   

10.
Crystallographic notation for Al2O3 is reviewed, with particular reference to the correct basis to be used in describing slip systems. A Groves-and-Kelly calculation showed that the combination of pyramidal slip on {11¯02}<11¯01> and basal slip on (0001){112¯0} will allow homogeneous deformation of Al2O3 polycrystals. Furthermore, operation of either the {101¯1}<1¯011> or the {011¯2}<2¯021> slip system will also satisfy the Von Mises criterion, since each system is capable of 5 independent deformation modes. Electron microscopy of an Al2O3 polycrystal deformed ≅5% at 1150°C under a hydrostatic confining pressure confirmed that pyramidal slip had occurred.  相似文献   

11.
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   

12.
Planar defects in the metastably retained h-BaTiO3 exhibiting α-fringe pattern have been characterized via transmission electron microscopy (TEM). The eligible fault vectors were determined by adopting the invisibility criteria of 2πg·R = 0 or 2 n π augmented by high-resolution imaging. Three stacking faults, F1, F2, and F3, of the extrinsic nature have been fully analyzed. The eligible fault vectors for faults F1 and F3 contained a basal component respectively of ⅓[0001] and ⅙[0001] and a common prismatic component of ⅓〈10[1-macr]0〉. However, only three of the 〈10[1-macr]0〉 vectors are the eligible prismatic component for the fault vectors RF1=⅓[0[1-macr]11], ⅓[10[1-macr]1], and ⅓[[1-macr]101], and RF3=⅙[02[2-macr]1], ⅙[2[2-macr]01], and ⅙[[2-macr]021] that have fulfilled the invisibility criteria. On the other hand, all fault vectors RF21=⅙〈[4-macr]223〉 for fault F2, containing six vectors of the 〈[2-macr]110〉 family, is eligible. Unlike the faults of πRF=⅙〈[2-macr]203〉 found in the D019 intermetallics of Ni3Sn and Co3W, neither fault F1 nor F3 is the π-rotation type. Fault F2, however, is a π-rotation fault since a 60°-rotation clockwise about [0001] has produced another eligible fault vector.  相似文献   

13.
Dihedral angles, Ψs, from surface thermal grooves were measured using a metal reference line technique for polycrystalline MgO, undoped Al2O3, and MgO-doped Al2O3. The values of Ψs span the following ranges: 89° to 116° for MgO at 1520 K, 76° to 166° for undoped Al2O3 at 1870 K, and 90° to 139° for MgO-doped Al2O3 at 1870 K. For all three systems, the median Ψs values are 105° to 113°, implying that the median γgbs is 1.1 to 1.2, in contrast to metals where γgbs ranges typically from 1/4 to 1/2. The widths in Ψs distributions were different for the three materials with the width increasing in the following order: MgO, MgO-doped Al2O3, undoped Al2O3. For all three materials, the grainboundary grooves and their corresponding Ψs were not symmetrical with respect to the surface normal. The asymmetry for MgO was due to the pinning of the grain boundaries by the surface thermal grooves. The range of inclination angles of the grain boundary to the surface was a function of Ψs, with the maximum inclination angles of ∼13°, in quantitative agreement with theory.  相似文献   

14.
Potassium-containing plate-like precipitates having the βm-alumina structure were identified within intergranular spinel inclusions in MgO-doped sintered aluminum oxide ceramics. Transmission electron microscopy revealed the corresponding epitaxial relation: (0001) βm{111} spinel and 1120 βm110 spinel. Chemical analyses and quantitative metallography suggested that in α-Al2O3 the 1800°C solubility limits of magnesium and potassium do not exceed 300 to 400 ppm and 5 to 10 ppm by weight, respectively.  相似文献   

15.
The crack systems associated with room-temperature Vickers indentations on {001} surfaces in 9.4-mol%-Y2O3-stabilized cubic ZrO2 single crystals are examined. The indentation-induced radial cracks are not always perpendicular to the free surface, as is usually assumed, and the surface traces are therefore not a reliable guide to the nature of the cracking. In fact, indentations with 〈100〉 diagonals do not form cracks on {010}, but form secondary radial cracks on inclined planes and noncoplanar median cracks on (110) or (1 1 0).  相似文献   

16.
A non-Arrhenius segregation isotherm is derived which includes the change in the heat of segregation with surface coverage due to impurity—impurity interactions. It is shown that a linear dependence of log ( Xs ) on the reciprocal temperature, where Xs is the surface atomic ratio, can derive either from a constant heat of segregation, i.e., Arrhenius behavior, or from a heat of segregation that varies as X−1s . This isotherm is then used to calculate the equilibrium surface coverages of Ca at the {001} surface of MgO1 Mg at the {0001} surface of α-Al2O3, γ at the {1012} and {1120} surfaces of α-Al2O3, and Na at the {111} and {110} surfaces of Li2O from the calculated heats of segregation. Where possible, comparisons are made with experiment. The more useful operational definition of the heat of segregation, namely, that derived from the measured coverage or that defined atomistically and obtained by calculation, is discussed.  相似文献   

17.
Thin films of Pb(Zr0.4TiO.6)O3 produced by chemical solution deposition were used to study the effects of stress from different platinized single-crystal substrates on film orientation and resulting electrical properties. Films deposited on MgO preferred a (001) orientation due to compressive stress on the film during cooling through the Curie temperature ( T C). Films on Al2O3 were under minimal stress at T C, resulting in a mixture of orientations. Those on Si preferred a (111) orientation due to templating from the bottom electrode. Films oriented in the 〈001〉 direction demonstrated lower dielectric constants and higher P r and − d 31 values than (111) films.  相似文献   

18.
The interface structures of calcium–alumino–silicate (CAS) glassy intergranular films (IGFs) formed between the combined basal and prism orientations of α-Al2O3 crystals were studied using molecular dynamics simulations. Preferential adsorption of specific ions from the IGFs to the contacting surfaces of the alumina crystals was observed. This segregation of specific ions to the interface enables formation of localized, ordered structures between the IGF and the crystal. However, the segregation behavior of the ions is anisotropic, depending on the orientation of the α-Al2O3 crystals. The results show that the enrichment of Ca atoms at the basal interface inhibits growth in the 〈0001〉 direction. However, at the (11 2 0) prism plane, Ca ions have little effect on the epitaxial adsorption of Al and O ions from the IGF onto the (11 2 0) surface. Increasing alumina concentration in the glassy IGF enhances adsorption of Al ions onto the prism surface, with little effect on the basal surface, indicating the tendency of growth in the 〈11 2 0〉 direction on the prism plane, but limited growth on the basal plane. These results are consistent with the experimental data regarding anisotropic grain growth in alumina sintered in the presence of CAS IGFs.  相似文献   

19.
Stacking faults in hot-pressed hexagonal BaTiO3 (the 6H-polytype) ceramic have been analyzed adopting the 2π g·R F=0 (or 2n π ) invisibility criteria, and dislocations were examined by the g·b =0 effective invisibility criteria, both using transmission electron microscopy. Perfect basal dislocations with bB= 1/3〈 1 2 1 0〉 have dissociated into pairs of prism-plane half-partials with bPr= 1/3〈01 1 0〉 by glide in the (0002) fault plane of an α-type extended planar stacking fault. Consequently, planar fault FA divided by half-partials was segmented into the initially faulted regions (FA1, FA3, FA5) of extended planar stacking faults and the complex-faulted regions (FA2, FA4) of complex extrinsic stacking faults superposed by the stacking fault ribbon created by half-partials. High-resolution imaging enables unambiguous differentiation of stacking sequence for all fault segments parted by half-partials. Change of the stacking sequence across half-partials reveals the atomic configurations of the faulted regions and the shear directions along 〈01 1 0〉 in (0002). The stacking fault energy of h-BaTiO3 derived from the separation between half-partials is compared with data from the literature. The discrepancy is attributed to non-equilibrium configuration and to the stacking sequence altered in the faulted regions embedded with half-partials.  相似文献   

20.
A thermal etching technique is developed to reveal dislocations in MgO · nAl2O3 (n=2.9) single crystals decorated by Al2O3 precipitation. Large plastic deformation leads to a larger density of thermal etch pits. Also, dense arrays of pits, parallel to the traces of slip planes with a surface, are observed around a Vickers identation. The cracks developed by an indentation run selectively along the 〈100〉 directions and accompany dislocations arrayed in the possible slip planes.  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