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 140 毫秒
1.
The pyroelectric properties of (1− x )Pb(Mg1/3Nb2/3)O3− x PbTiO3 (PMN− x PT) single crystals with various compositions and orientations have been investigated using a dynamic method. Excellent pyroelectric performances can be achieved in 〈111〉-oriented rhombohedral PMN− x PT (0.24≤ x ≤0.30) crystals, where the measurement direction corresponds to the polar axis of the crystal. At room temperature, the pyroelectric coefficient and the detectivity figure of merit ( F d ) for the 〈111〉-oriented PMN–0.28PT single crystal are 8.55 × 10−4 C·(m2·K)−1 and 9.89 × 10−5 Pa−1/2 (100 Hz), respectively, superior to those of the widely used pyroelectric materials. They are also weak temperature dependent and nearly independent of frequency. These outstanding pyroelectric performances make the single crystals a promising candidate for uncooled infrared detectors and thermal imagers.  相似文献   

2.
Single crystals of Al2O3 containing 0.5 wt% Fe were exposed to low p o2 atmospheres at 1500°C to produce precipitate phases. Analytical TEM identified the precipitate phases as spinel (hercynite) and iron, with the following orientation relationships: 〈001〉Fe‖〈2 2 01〉s with {1 1 0}Fe‖{11 2 0}s, 〈111〉Fe‖〈10 1 0〉s with {0 1 1}Fe‖ (0001)s, and 〈1 1 0〉H‖〈01 1 0〉s with {111}H‖ (0001)s. The three phase fields observed — (Fe, Al)2O3, spinel +α-Al2O3, and iron +αAl2O3— are in accordance with phase stability diagrams. Precipitation kinetics indicate that oxygen is mobile in the reduced region of the crystal.  相似文献   

3.
A dielectric loss study of porous MgO indicates that H2O absorption on MgO probably leads to the formation of surface defects as well as hydroxide ions. The ac conductivity, σac followed the equation σac1 + KP H2O0.27 from 550° to 800°C when the water partial pressure was varied between 7 × 10−5 and 3 × 10−2 atm.  相似文献   

4.
Planar faults, designated F 1– F 6 and F a , with intermixed dislocations in metastably retained hexagonal-BaTiO3 ceramics, were found and analyzed by transmission electron microscopy. Only faults with one, two, and four extra c -layers parallel to the basal plane have been identified. Fault vectors RF1=RFa=1/6[02 2 1], RF2=RF3=1/3[0 1 11], RF4=1/3[01 1 2], and RF'1=RF5= RF6=RF'a=1/6[0 2 21] were determined adopting the 2πg·R=0 (or 2 n π) criteria in combination with high-resolution imaging. Further, the embedded dislocations were half-partials with Burgers vectors b=1/3〈10 1 0〉 determined by the g·b=0 effective invisibility criteria in conjunction with the eligible fault vectors. A rotation by 60° about the c -axis was found between fault segments with RF1=1/6[02 2 1] and RF'1=1/6[20 2 1] located on either side of a partial. Basal dislocations with bB=1/3〈1 2 10〉 have dissociated into two prism-plane Shockley half-partials with bPr=1/3〈1 1 00〉 by glide in the fault plane (0002) according to 1/3〈1 2 10〉→1/3〈1 1 00〉+1/3〈0 1 10〉. The fault segment F'1 encompassed by two half-partials is an extrinsic complex stacking fault.  相似文献   

5.
Cation ordering and domain boundaries in perovskite Ca[(Mg1/3Ta2/3)1− x Ti x ]O3 ( x =0.1, 0.2, 0.3) microwave dielectric ceramics were investigated by high-resolution transmission electron microscopy (HRTEM) and Rietveld analysis. The variation of ordering structure with Ti substitution was revealed together with the formation mechanism of ordering domains. When x =0.1, the ceramics were composed of 1:2 and 1:1 ordered domains and a disordered matrix. The 1:2 cation ordering could still exist until x =0.2 but the 1:1 ordering disappeared. Neither 1:2 nor 1:1 cation ordering could exist at x =0.3. The space charge model was used to explain the cation ordering change from 1:2 to 1:1 and then to disorder. A comparison between the space charge model and random layer model was also conducted. HRTEM observations showed an antiphase boundary inclined to the (111) c plane with a projected displacement vector in the 〈001〉 c direction and ferroelastic domain boundaries parallel to the 〈100〉 c direction.  相似文献   

6.
Crystals of β-Ca2SiO4 (space group P 121/ n 1) were examined by high-temperature powder X-ray diffractometry to determine the change in unit-cell dimensions with temperature up to 645°C. The temperature dependence of the principal expansion coefficients (αi) found from the matrix algebra analysis was as follows: α1= 20.492 × 10−6+ 16.490 × 10−9 ( T - 25)°C−1, α2= 7.494 × 10−6+ 5.168 × 10−9( T - 25)°C−1, α3=−0.842 × 10−6− 1.497 × 10−9( T - 25)°C−1. The expansion coefficient α1, nearly along [302] was approximately 3 times α2 along the b -axis. Very small contraction (α3) occurred nearly along [     01]. The volume changes upon martensitic transformations of β↔αL' were very small, and the strain accommodation would be almost complete. This is consistent with the thermoelasticity.  相似文献   

7.
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   

8.
Vibrational Spectra of Hydrothermally Prepared Hydroxyapatites   总被引:1,自引:0,他引:1  
Apatites can be used as bioceramic materials for tooth and bone implants. Infrared (4000 to 400 cm−1) and laser-Raman spectra were obtained for hydrothermally prepared hydroxyapatites with the general formula X10(PO4)6(OH)2 where X= Ca2+ Sr2+, Ba2+, or Pb2+ Fundamental vibrational modes were identified and empirical band assignments made. Analysis of the vibrational spectra led to the reassignment of the v2 mode to a higher wave-number, lower-intensity band (e.g. ∼470 cm−1 for Ca hydroxyapatite). Effects of solid-state interaction were noted in the vibrational spectra and interpreted by factor-group analysis. An "effective" factor group C6h, which neglects the locations of the OH ions, best describes the vibrational spectra of the hydroxyapatites.  相似文献   

9.
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   

10.
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   

11.
Magnesium oxide crystals and whiskers were grown by vapor phase reactions between MgO and tungsten, hydrogen, and carbon. The large number of morphologies obtained are described in terms of two, three, and fourfold growth symmetry and are related to the growth conditions by current heterogeneous nucleation theory. Threefold (or 〈111〉) growth occurred principally with MgO + H2 reactions, and in all experiments 〈100〉 and 〈110〉 growth occurred under closely similar conditions of temperature and vapor supersaturation. Observations of growth steps suggest that {100} surfaces are favored only at the highest vapor supersaturation. Tensile strength measurements of whiskers with l/d ratios of about 300 show increasing strength with decreasing cross section. A value of 8.13 × 104 psi was observed for a whisker 9.76μ in diameter. The results are of particular interest to the formation of refractory oxides crystallizing in the simple cubic NaCl structure.  相似文献   

12.
Thermal expansion of the low-temperature form of BaB2O4 (β-BaB2O4) crystal has been measured along the principal crystallographic directions over a temperature range of 9° to 874°C by means of high-temperature X-ray powder diffraction. This crystal belongs to the trigonal system and exhibits strongly anisotropic thermal expansions. The expansion along the c axis is from 12.720 to 13.214 Å (1.2720 to 1.3214 nm), whereas it is from 12.531 to 12.578 Å (1.2531 to 1.2578 nm) along the a axis. The expansions are nonlinear. The coefficients A, B , and C in the expansion formula L t = L 0(1 + At + Bt 2+ Ct 3) are given as follows: a axis, A = 1.535 × 10−7, B = 6.047 × 10−9, C = -1.261 × 10−12; c axis, A = 3.256 × 10−5, B = 1.341 × 10−8, C = -1.954 × 10−12; and cell volume V, A = 3.107 × 10−5, B = 3.406 × 10−8, C = -1.197 × 10−11. Based on α t = (d L t /d t )/ L 0, the thermal expansion coefficients are also given as a function of temperature for the crystallographic axes a , c , and cell volume V.  相似文献   

13.
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   

14.
The defect structure of acceptor-doped (Fe, Al, Cr) polycrystalline strontium titanate was investigated by measuring the equilibrium electrical conductivity as a function of oxygen activity (10−21≤ao2≤1) and temperature (85°C≤ T ≤ 1050°C). The electrical conductivity was n type with ∼−1/4 dependence on a 02 for a02<10−10. For a02>10−8, the observed data for Fe- or Al-doped samples were proportional to ∼1/4.5 power of a02. In this region for Cr-doped samples, the value of m in.σp αa02+1/m varies from ∼4.80 to ∼9.00 as the concentration of Cr is increased from 460 to 10 000 ppm. The onset of p -type conductivity depends on the amount. of acceptor impurity added to the sample. The absolute values of the conductivity in the acceptor-doped samples were lower in the n-type region than those for undoped SrTiO3. The conductivity minima shift toward lower oxygen activity with increasing acceptor concentration. For the entire oxygen activity rangae used in this study, the defect structure of SrTiO3 is dominated by the added acceptor impurities.  相似文献   

15.
The effect on the γ-Al2O3-to-α-Al2O3 phase transition of adding divalent cations was investigated by differential thermal analysis, X-ray diffractometry, and surface-area measurements. The cations, Cu2+, Mn2+, Co2+, Ni2+, Mg2+, Ca2+, Sr2+, and Ba2+, were added by impregnation, using the appropriate nitrate solution. These additives were classified into three groups, according to their effect: (1) those with an accelerating effect (Cu2+ and Mn2+), (2) those with little or no effect (Co2+, Ni2+, and Mg2+), and (3) those with a retarding effect (Ca2+, Sr2+, and Ba2+). The crystalline phase formed by reaction of the additive with γ-Al2O3 at high temperature was a spinel-type structure in groups (1) and (2) and a magnetoplumbite-type structure in group (3). In groups (2) and (3), a clear relationship was found between the transition temperature and the difference in ionic radius of Al3+ and the additive (Δ r ): The transition temperature increased as Δ r increased. This result indicates that additives with larger ionic radii are more effective in suppressing the diffusion of Al3+ and O2− in γ-Al2O3, suppressing the grain growth of γ-Al2O3, and retarding the transformation into α-Al2O3.  相似文献   

16.
Excitation of Tm3+ to 3 H 4 using the 791 nm pump source showed the frequency up-converted blue emission (∼480 nm) due to the Tm3+:1 G 43 H 6 transition in Tm3+/Nd3+ codoped CaO·Al2O3 glasses. Intensity and lifetime changes with rare-earth concentrations suggested the efficient energy transfer of Tm3+:3 H 4→ Nd3+:4 F 5/2 and Nd3+:4 F 3/2→ Tm3+:1 G 4. The latter transfer enabled Tm3+ to reach its 1 G 4 level, and the blue emission became possible through the 1 G 43 H 6 transition. Quantitative analysis with rate equations proved that these two transitions were the most efficient among all the possible energy transfer routes between Tm3+ and Nd3+. Calculated up-conversion efficiency of the Tm3+/Nd3+ combination in CaO·Al2O3 glass was 6.6 × 10−3, and it was ∼4 orders of magnitude larger than those reported for other oxide glasses.  相似文献   

17.
Anti-phase boundary (APB) domains resulting from the orthorhombic ( o 1, Ibmm (No. 74)) to orthorhombic ( o 2, Pbnm (No. 62)) phase transformation at ∼290°C in pressureless-sintered barium cerate (BaCeO3) ceramics have been investigated by transmission electron microscopy and large-angle convergent beam electron diffraction (LACBED). APB exhibiting symmetric π-fringe patterns in both bright-field and dark-field images, lying in {011) with the fault vector R =1/2〈111〉, were determined adopting the invisibility criteria of 2π g · R =0 or 2 n π. The curved boundaries due to symmetry change can be related to the low symmetry phase by a pure translation upon phase transition. The displacement vector, R , relating the domains is the translational symmetry element from the origin to the lattice point that was lost upon transition. The formation of such domains induced by the o 1→ o 2 phase transition is discussed in terms of change in the Bravais lattice and loss of lattice point upon the transition. The identification has also confirmed the existence of a higher symmetry orthorhombic o 1 phase along the transition sequence of R 3 c → Ibmm → Pbnm . The increased spatial frequency of APB domains is attributed to A-site ordering due to vacancies created by BaO loss and/or disordering of the B-site substitutional defects generated extrinsically when doped with either a Y3+-acceptor or a Nb5+-donor.  相似文献   

18.
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   

19.
Abstract. For the SETAR (2; 1,1) model

where {at(i)} are i.i.d. random variables with mean 0 and variance σ2(i), i = 1,2, and {at(l)} is independent of {at(2)}, we consider estimators of φ1, φ 2 and r which minimize weighted sums of the sum of squares functions for σ2(1) and σ2(2). These include as a special case the usual least squares estimators. It is shown that the usual least squares estimators of φ1, φ2 and r are consistent. If σ2(1) ≠σ2(2) conditions on the weights are found under which the estimators of r and φ1 or φ2 are not consistent.  相似文献   

20.
Lanthanum-doped ceria powder with a composition Ce0.8La0.2O1.9 was prepared by heating the oxalate solid solution (Ce0.8La0.2)2(C2O4)3 at 873 K in air. As-prepared powder was densified to 96%–97% relative density by sintering in air at 1773 K for 4 h. The electronic current of the disk sample was measured in a temperature range from 773 to 1113 K by the direct current polarization method using a Hebb–Wagner ion-blocking cell. A linear relationship, which was theoretically predicted, was measured between log σe (electronic conductivity) and E (applied voltage) in the applied voltage range of 0.2–1.0 V. The σe was proportional to P O2−1/4.3≈ P O2−1/4.6 in the oxygen partial pressure range of 10−2–10−8 Pa, and to P O2−1/6.7≈ P O2−1/7.1 in the oxygen partial pressure range of 10−7.5–10−22 Pa. The apparent activation energy of the electronic conduction was 1.87–1.94 eV. The hole conduction was also measured in the P o2 range of 102–105 Pa. The transport number of oxide ion was 0.96–1.00 at 773–1113 K under an oxygen partial pressure of 10−5 Pa.  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