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 31 毫秒
1.
The hydrated oxygen deficient complex perovskite-related materials Sr4(Sr2Nb2)O11·nH2O and Sr4(Sr2Ta2)O11·nH2O were studied at high water vapour pressures over a large temperature range by electrical conductivity measurements, thermogravimetry (TG), and X-ray powder diffraction (XRPD). In humid atmospheres both materials are known to exhibit protonic conductivity below dehydration temperatures, with peak-shaped maxima at about 500 °C. In this work we show that the peaks expand to plateaus of high conductivity from 500 to 700 °C at a water vapour pressure of 1 atm. However, in situ synchrotron XRPD of Sr4(Sr2Nb2)O11·nH2O as a function of temperature shows that these observations are in fact coincident with melting and dehydration of a secondary phase Sr(OH)2. The stability of Sr4(Sr2Nb2)O11·nH2O and Sr4(Sr2Ta2)O11·nH2O in humid atmospheres is thus insufficient, causing decomposition into perovskites with lower Sr content and SrO/Sr(OH)2 secondary phases. This, in turn, rationalizes the observation of peaks and plateaus in the conductivity of these materials.  相似文献   

2.
Photoluminescence (PL) of Al2O3 films obtained by anodization of thermally evaporated and annealed thin Al films on p++Si in 0.3 M oxalic acid has been investigated. Thermal annealing at 200–950 °C under the dry nitrogen atmosphere was used for deactivation of luminescence centres. Luminescence from as grown films was broad and located at 425 nm. This luminescence reached to highest level after annealing at 600 °C. Maximum 10 min was required for full optical activation and prolonged annealing up to 4 h did not change the luminescence intensity. Because of deep levels, absorption band edge of as grown films was shifted to the lower energy which is 3.25 eV. Annealing above 800 °C reduced the PL intensity and this observation was correlated with the blue shift of band edge as the defects annealed out. Disappearing PL intensity and blue shift of band edge absorption after annealing at 950 °C was mainly attributed to the oxygen-related defects and partly to impurities that may be originated from oxalic acid. AFM results did not show any hexagonally ordered holes but uniformly distributed nanosized Al2O3 clusters that were clearly seen. XRD measurements on as grown Al2O3 showed only [1 1 0] direction of α phase. Debye–Scherer calculation for this line indicates that cluster size is 35.7 nm. XRD and AFM pictures suggest that nanocrystalline Al2O3 are embedded in amorphous Al2O3.  相似文献   

3.
Single crystals of NdBa2Cu3O7−δ (Nd123) have been successfully grown by the top-seeded solution-growth (TSSG) method in 1%, 21% and 100% oxygen partial pressure atmosphere ((P(O2) = 0.01 atm, P(O)2) = 0.21 atm and P(O2) = 1.00 atm). Ba---Cu---O solvent with a Ba to Cu ratio of 3:5 was used in a Nd2O3 crucible. Nd is supplied by the reaction between the molten solvent and the Nd2O3 crucible. Compositions of Nd123 single crystals grown in different oxygen partial pressure atmospheres were analyzed by inductivity coupled plasma atomic emission spectrometry (ICP-AES) and confirmed to be Nd:Ba:Cu = 1.01:1.97:3.00 for P(O2) = 0.01 atm, Nd:Ba:Cu = 1.07:1.95:3.00 for P(O2) = 0.21 atm and Nd:Ba:Cu = 1.10:1.90:3.00 for P(O2) = 1.00 atm, respectively. The Nd123 single crystals grown in different oxygen partial pressure atmospheres were annealed in a pure oxygen gas flow, and the temperature dependence of the DC magnetization for these crystals was measured using a superconducting quantum interference device (SQUID) magnetometer. The Nd123 single crystal grown in P(O2) = 0.01 atm, and annealed at 340°C for 200 h showed a steep superconductive transition at 96 K. On the other hand, the Nd123 crystal grown in P(O2) = 0.21, 1.00 atm and, annealed at 340°C for 200 h exhibited a broad transition at 92 K for P(O2) = 0.21 atm and at 88 K for P(O2) = 1.00 atm, respectively. Therefore for Nd123 single crystal production with high quality superconductive characteristics, a low oxygen partial pressure atmosphere during crystal growth is found to be effective for minimizing the substitution of Nd ions into Ba sites.  相似文献   

4.
By using very small amount – 11‐mg – of sample powder, major oxides (Na2O, MgO, Al2O3, SiO2, P2O5, K2O, CaO, TiO2, MnO, and total Fe2O3) in ancient pottery (and igneous rocks) were determined with X‐ray fluorescence spectrometry. This minimized amount of sample was used to prepare a fused glass bead with 300 times the weight of lithium tetraborate as an alkali flux. Calibration standards were obtained by compounding chemical reagents (Na2CO3, MgO, Al2O3, SiO2, Na4P2O7, K2CO3, CaCO3, TiO2, MnO2, and Fe2O3) and the flux. Fewer 11 mg of reagents as oxides were able to give reliable calibration curves with good linearity (correlation coefficient: r > 0.995). Fewer 11 mg of sample was able to give reliable analytical results with good precision (relative standard deviation: <3% for more than 10.0 mass% of analyte, <10% for 1.0–10.0 mass% of analyte, and <20% for 0.1–1.0 mass% of analyte). Lower limits of detection were roughly a sub‐percentage of analyte in an unprepared sample (e.g. 0.3 mass% for Na2O, 0.5 mass% for MgO, 1.0 mass% for Al2O3, and 0.01 mass% for MnO). Composition of major oxides in artificial and natural aluminosilicate materials (including rock, stone, sand, sediment, and clay; and their products) should be fundamental information to be considered in detail. The present X‐ray determination based on very small amount of sample might be made readily accessible for destructive analysis of precious samples for archaeology (and geochemistry). Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

5.
Single crystals of α-Al2O3 (10×10 mm2, 0.4 mm thick) were annealed in vacuum at about 1500 °C in the ambience of boron. The OA studies on these samples showed bands at 203, 232 and 258 nm signifying that such a treatment leads to the formation of F and F+ centers in significant concentrations, these bands, however, were not found in the Al2O3 crystals processed in the similar manner in the absence of boron. The Al2O3:B samples were irradiated to different absorbed doses of 90Sr/90Y β-source and the continuous wave OSL (CW-OSL) was recorded on the samples using 470 nm blue light stimulation. These samples have shown a linear TL and CW-OSL response in the dose range of 20 mGy to 15 Gy. The minimum detectable dose, corresponding to 3σ limit of the variation of the output of the unirradiated dosimeters, was found to be 100 μGy. Irradiated samples stored in dark at room temperature for a period of two months show negligible fading. The TL and OSL sensitivities of the samples were found to be strongly dependent on process temperature and time. The TL response is marked by the absence of low temperature peak (<100 °C), unlike the case of α-Al2O3:C, implying that the boron doping does not lead to formation of shallow traps. The Al2O3:B samples show faster photoionisation cross-section as compared to α-Al2O3:C. This approach of processing of single crystal Al2O3 in the boron ambience thus represents a potential way of introducing dosimetrically pertinent defects in Al2O3 single crystals.  相似文献   

6.
AlN thin films have been grown on a-plane sapphire (Al2O3(112̄0)) substrates. X-ray diffraction measurements indicate the films are fully c-plane (0001) oriented with a full width at half maximum of the AlN(0002) rocking curves of 0.92. The epitaxial growth relationships have been determined by the reflection high energy electron diffraction analysis as AlN[11̄00]//Al2O3[0001] and AlN[112̄0]//Al2O3[11̄00]. Angular dependence of important surface acoustic wave (SAW) characteristics, such as the phase velocity and electromechanical coupling coefficient, has been investigated on the AlN(0001)/Al2O3(112̄0) structure. While the SAW is excited at all propagation angles with an angular dispersion of the phase velocity in the range of 5503–6045 m/s, a higher velocity shear-horizontal (SH) mode is observed only at 0°, 105° and 180° off the reference Al2O3[11̄00] over a 180° angular period. The phase velocity of the SH mode shows dispersion (6089–6132 m/s) as a function of the SAW wavelength. Temperature coefficients of frequency are also demonstrated for both modes. PACS 81.15.Hi; 77.84.-s; 77.65.Dq  相似文献   

7.
Cross-sectional observation of the oxide scale formed by oxidation of TiC single crystals with (100), (110) and (111) faces at 1500 °C for 30 min in a mixed gas of Ar/O2 (PO2=0.08 kPa) was performed by transmission electron microscope (TEM). The oxide scale was composed of outer (zone 2) and inner (zone 1) subscales. TEM and selected area electron diffraction combined with X-ray energy dispersion analysis showed that the zone 2 consists of rutile and pores and the zone 1 of carbon and titanium oxide, identified as Ti3O5 in the oxide scale formed on the (110) face. Zone 1 formed on the (100) and (111) faces showed crisscross patterns, in contrast to the (110) with the wavy lamellar pattern.  相似文献   

8.
Amorphous, nanocrystalline, and bulk AlO(OH) · xH2O crystals have six fundamental modes (FM) of vibration in a nonlinear AlO(OH) molecular structure. Most of them appear in groups of four IR and Raman bands. Their positions and relative intensities differ significantly in three specimens. The nanocrystals (monoclinic structure with z=8 molecules per unit cell) have four OH stretching bands at values enhanced by up to 360 cm−1 at 3120, 3450, 3560 cm−1 in comparison to those in bulk crystals or amorphous specimens. The first two bands are broad, bandwidth Δν1/2200 to 350 cm−1, while the other two are sharp, Δν1/290 cm−1. The sharp bands shift to 3525 and 3595 cm−1 after heating the sample at 100°C. They no longer appear after heating at 300 or 500°C for 2 h (the specimen decomposes to Al2O3), leaving behind only two bands at 3100 and 3400 cm−1. A Δν1/2 value of 500 cm−1 appears in the 3400 cm−1 in a delocalized distribution of H atoms. Two bands also occur at 3098 and 3300 cm−1 in bulk crystals (orthorhombic structure with z=4) or at 2990 and 3515 cm−1 in an amorphous sample. More than one bands appear in a FM vibration in occurrence of sample in more than one conformers. The amorphous sample has approximately the same conformer structure as the bulk crystals. An amorphous surface structure exists in nanocrystals with a group of three bands at 1420, 1510 and 1635 cm−1 in an interconnected network structure. It encapsulates the nanocrystals in an amorphous shell. Its volume fraction, 33% estimated from the integrated intensity in three bands, determines 2.2 nm thickness in the shell in spherical shape of nanocrystals in 35 nm diameter.  相似文献   

9.
Alternately Er doped Si-rich Al2O3 (Er:SRA) multilayer film, consisting of alternate Er-Si-codoped Al2O3 (Er:Si:Al2O3) and Si-doped Al2O3 (Si:Al2O3) sublayers, has been synthesized by co-sputtering from separated Er, Si, and Al2O3 targets. The dependence of Er3+ related photoluminescence (PL) properties on annealing temperatures over 700-1100 °C was studied. The maximum intensity of Er3+ PL, about 10 times higher than that of the monolayer film, was obtained from the multilayer film annealed at 950 °C. The enhancement of Er3+ PL intensity is attributed to the energy transfer from the silicon nanocrystals in the Si:Al2O3 sublayers to the neighboring Er3+ ions in the Er:Si:Al2O3 sublayers. The PL intensity exhibits a nonmonotonic temperature dependence: with increasing temperature, the integrated intensity almost remains constant from 14 to 50 K, then reaches maximum at 225 K, and slightly increases again at higher temperatures. Meanwhile, the PL integrated intensity at room temperature is about 30% higher than that at 14 K.  相似文献   

10.
The thermodynamic properties, spin–lattice relaxation times, T1, and spin–spin relaxation times, T2, of the 27Al, 87Rb, and 133Cs nuclei in MAl(SO4)2·12H2O (M=Rb and Cs) crystals were investigated, and the two crystals were found to lose H2O with increases in temperature. From our results for T1 and T2, we conclude that the discontinuities near Td in the T1 curves of the two crystals correspond to structural changes. In both crystals, below Td the water molecules surrounding the Al3+ and M+ nuclei form distorted octahedra, whereas above Td the water molecules around the Al3+ and M+ nuclei form regular octahedra and the environment of the Al3+ and M+ nuclei has cubic symmetry. Further, the T1 for the 27Al and 87Rb nuclei in RbAl(SO4)2·12H2O below Td were found to increase with increasing temperature, whereas the T1 for the 27Al and 133Cs nuclei in CsAl(SO4)2·12H2O were found to decrease. It is possible that this difference is due to the different characteristics of α- and β-type crystals.  相似文献   

11.
Compared to experiment, the adsorption energies, bonding properties, and electronic structure of two different Al2O3/B4C bridge sites with seven different Al2O3 surfaces are investigated by ab initio periodic density functional theory. The Al2O3/B4C ceramic sintered in Ar is synthesized and measured by XRD and TEM. The calculated results reveal that the densification of O_bridge site of Al2O3/B4C surface is better than that of Al_bridge. The Al2O3 (1 1 3)/B4C with O_bridge is the most favorable and stable. The electronic structure shows that the electron hybridization exists between Al, O atoms and C, B atoms. The results indicate that the calculated results are in good agreement with the experiment.  相似文献   

12.
We present a theoretical study of the collisions of atomic oxygen with O-precovered β-cristobalite (1 0 0) surface. We have constructed a multidimensional potential energy surface for the O2/β-cristobalite (1 0 0) system based mainly on a dense grid of density functional theory points by using the interpolation corrugation-reducing procedure. Classical trajectories have been computed for quasithermal (100–1500 K) and state-specific (e.g., collision energies between 0.01 and 4 eV) conditions of reactants for different O incident angles (θv). Atomic sticking and O2(adsorbed) formation are the main processes, although atomic reflection and Eley–Rideal (ER) reaction (i.e., O2 gas) are also significant, depending their reaction probabilities on the O incident angle. ER reaction is enhanced by temperature increase, with an activation energy derived from the atomic recombination coefficient (γO(θv = 0°, T)) equal to 0.24 ± 0.02 eV within the 500–1500 K range, in close agreement with experimental data. Calculated γO(θv = 0°, T) values compare quite well with available experimental γO(T) although a more accurate calculation is proposed. Chemical energy accommodation coefficient βO(T) is also discussed as a function of ER and other competitive contributions.  相似文献   

13.
The Ag2O–TiO2–SiO2 glasses were prepared by Ag+/Na+ ion-exchange method from Na2O–TiO2–SiO2 glasses at 380–450 °C below their glass transition temperatures (Tg), and their electrical conductivities were investigated as functions of TiO2 content and the ion-exchange ratio (Ag/(Ag+Na)). In a series of glasses 20R2xTiO2·(80−x)SiO2 with x=10, 20, 30 and 40 in mol%, the electrical conductivities at 200 °C of the fully ion-exchanged glasses of R=Ag were in the order of 10−5 or 10−4 S cm−1 and were 1 or 2 orders of magnitude higher than those of the initial glasses of R=Na. The glass of x=30 exhibited the highest increase of conductivity from 3.8×10−7 to 1.3×10−4 S cm−1 at 200 °C by Ag+/Na+ ion exchange among them. When the ion-exchange ratio was changed in 20R2O·30TiO2·50SiO2 system, the electrical conductivity at 200 °C exhibited a minimum value of 7.6×10−8 S cm−1 around Ag/(Ag+Na)=0.3 and increased steeply in the region of Ag/(Ag+Na)=0.5–1.0. When the ion-exchange temperature was changed from 450 to 400 °C, the conductivity of the ion-exchanged glass of x=30 decreased. The infrared spectroscopy measurement revealed that the ion-exchange temperature of 450 °C induced a structural change in the glass of x=30. The Tg of the fully ion-exchanged glass of x=30 was 498 °C. It was suggested that the incorporated silver ions changed the average coordination number of titanium ions to form higher ion-conducting pathway and resulted in high conductivity in the titanosilicate glasses.  相似文献   

14.
《Solid State Ionics》1986,20(1):17-23
The Al2O3 rich region of the Na2O:MgOAl2O3 system has been investigated by means of X-ray phase analytical methods. The main features of the isothermal sections at 1870 K and 1950 K have been determined. The two sections exhibit the same phase equilibria except in the Na2Al2O4-β″ region where a liquids phase appears at 1950 K. The magnesium stabilized β″-alumina phase with the nominal formula Na1xxAl11xO17 exhibits an extended homogeneity range with the magnesium content ranging over x=0.59-0.72. The homogeneity ranges are accompanied by variations in the unit cell dimentions. The composition of β″-alumina appears to correspond to a far lower relative sodium cntent than that indicated by the ideal formula at large magnesium + aluminium contents.  相似文献   

15.
Sodium tracer diffusivity and ionic conductivity have been measured from room temperature down to ?79°C on the same single crystals of sodium-β-alumina of composition 1.23Na2O·11Al2O3. The Haven ratio decreases from 0.35 at room temperature to 0.18 at ?79°C. The low value of HR and its temperature dependence can be explained in terms of high correlation effects for ionic diffusion due to large associated defects present at low temperatures.  相似文献   

16.
The article reports on the effect of the addition of copper in the Al2O3 film on its mechanical and optical properties. The Al–Cu–O films were reactively co-sputtered using DC pulse dual magnetron in a mixture of Ar + O2. One magnetron was equipped with a pure Al target and the second magnetron with a composed Al/Cu target. The amount of Al and Cu in the Al–Cu–O film was controlled by the length of pulse at the Al/Cu target. The Al–Cu–O films with ≤16 at.% Cu were investigated in detail. The addition of Cu in Al2O3 film strongly influences its structure and mechanical properties. It is shown that (1) the structure of Al–Cu–O film gradually varies with increasing Cu content from γ-Al2O3 at 0 at.% Cu through (Al8−2x,Cu3x)O12 nanocrystalline solid solution to CuAl2O4 spinel structure, (2) the Al–Cu–O films with ≥3 at.% Cu exhibit (i) relatively high hardness H increasing from ∼15 GPa to ∼20 GPa, (ii) enhanced elastic recovery We increasing from ∼67% to ∼76% with increasing Cu content from ∼5 to ∼16 at.% Cu and (iii) low values of Young's modulus E* satisfying the ratio H/E* > 0.1 at ≥5 at.% Cu, and (3) highly elastic Al–Cu–O films with H/E* > 0.1 exhibit enhanced resistance to cracking during indentation under high load.  相似文献   

17.
Homogeneous glasses in the Na2O–B2O3–Al2O3–SiO2 system doped with proper amount of AgCl were obtained by melting at a temperature of 1450 °C. Then, with several steps of treatment, including crystallization, elongation and reduction, the glass with oriented arrangement of needle-like micro–nano silver particles was produced. The microstructure and the optical properties of the glass samples in different stages were studied by SEM-EDAX, FE-SEM and UV–Vis spectrum. The results showed that the glass after elongation and reduction exhibits excellent polarization performance in the wavelength range from 600 nm to 900 nm, with an extinction ratio larger than 45 dB. The glass only elongated shows also slight polarizing performance, which may result from the formation of filament structure of Ag during elongation processing.  相似文献   

18.
The glasses with the composition of 37.5Li2O–(25 − x)Fe2O3xNb2O5–37.5P2O5 (mol%) (x = 5,10,15) are prepared, and it is found that the addition of Nb2O5 is effective for the glass formation in the lithium iron phosphate system. The glass–ceramics consisting of Nasicon-type Li3Fe2(PO4)3 crystals with an orthorhombic structure are developed through conventional crystallization in an electric furnace, showing electrical conductivities of 3 × 10− 6 Scm− 1 at room temperature and the activation energies of 0.48 eV (x = 5) and 0.51 eV (x = 10) for Li+ ion conduction in the temperature range of 30–200 °C. A continuous wave Nd:YAG laser (wavelength: 1064 nm) with powers of 0.14–0.30 W and a scanning speed of 10 μm/s is irradiated onto the surface of the glasses, and the formation of Li3Fe2(PO4)3 crystals is confirmed from XRD analyses and micro-Raman scattering spectra. The crystallization of the precursor glasses is considered as new route for the fabrication of Li3Fe2(PO4)3 crystals being candidates for use as electrolyte materials in lithium ion secondary batteries.  相似文献   

19.
Radioluminescence (RL) and optically stimulated luminescence (OSL) results of LiAlO2 were compared with Al2O3:C. For blue (470 nm) optical stimulation, RL + OSL signal in LiAlO2 exhibited a sharp initial increase followed by a decay within the first 20 s of continuous wave (CW)-OSL and a very slow increase thereafter. The RL + OSL signal was about 1.25 times of its RL signal in LiAlO2 as compared to 2.5 times in Al2O3:C. With the continued beta irradiation, the RL signal exhibited a faster growth in LiAlO2 than that in Al2O3:C. Emission spectrum of LiAlO2 exhibited multiple emission peaks in the range from 320 to 380 nm as against 410 nm of Al2O3:C. RL and OSL emission spectra were similar in both LiAlO2 and Al2O3:C. The intense RL in LiAlO2 (about 300 times of that of its background signal) for a beta ray dose rate of 4.6 mGy/s appears attractive for radiation dosimetry including real time/online dosimetry and dose mapping.  相似文献   

20.
We examine the oscillator strengths and the intensity parameters Ω t (t = 2, 4, 6) of yttrium-aluminum, scandium-containing, and gallium garnet crystals doped with Er3+ ions. A comparative analysis of the oscillator strengths and the intensity parameters Ω t (t = 2, 4, 6) of garnets with different contents of Al3+ and Sc3+ ions (Gd2.4Er0.5Sc1.8Al3.3O12, Gd2.4Er0.5Sc1.9Al3.2O12, Gd2.4Er0.5Sc2.0Al3.1O12) is performed, as a result of which the oscillator strengths and the intensity parameters Ω t (t = 2, 4, 6) of these crystals are shown to have close values. We find that Ca3(NbGa)5O12 crystals doped with Er3+ ions are characterized by highest values of the oscillator strengths for hypersensitive transitions and of the intensity parameter Ω2 of Er3+ ions compared to the values of these quantities in the examined garnet crystals, which is determined by the fact that the symmetry of the local environment of Er3+ ions in these crystals is C 1, C 2, or C . We reveal that, as the concentration of Er3+ ions in these crystals increases from 1 to 39 at %, both the oscillator strength of the hypersensitive transition 4 I 15/22 H 11/2 of Er3+ ions and their intensity parameter Ω2 tend to decrease, which can be related to an increase in the relative fraction of Er3+ ions with higher symmetry of the local environment.  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