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 530 毫秒
1.
The knowledge of the phase relations and solubilities in the Y–Ba–Cu–O and Nd–Ba–Cu–O systems are of fundamental importance for crystal growth and liquid-phase epitaxy of YBa2Cu3O7−δ (YBCO) and Nd1+xBa2−xCu3Oδ (NdBCO). The determination of the solubility curve of YBCO and NdBCO in a BaO/CuO flux containing 31 mol% BaO was done by observation of the formation and dissolution of crystals on the surface of the high-temperature solution. The heat of the solution of YBCO at 1000°C was found to be 34.7 kcal/mol, and for NdBCO at 1060°C, it was found to be 28.1 kcal/mol. The determination of the solubility curves requires special care, and the problems of the time-dependent shift of the solution composition due to the corrosion of the crucible is discussed. The scatter of the solubility data published by different authors could be due to the use of solutions with different Ba : Cu ratios, different determination methods, i.e. different crystallization mechanisms, different crucibles and starting chemicals.  相似文献   

2.
This paper reports the growth and spectral properties of 3.5 at% Nd3+:LaVO4 crystal with diameter of 20×15 mm2 which has been grown by the Czochralski method. The spectral parameters were calculated based on Judd–Ofelt theory. The intensity parameters Ωλ are: Ω2=2.102×10−20 cm2, Ω4=3.871×10−20 cm2, Ω6=3.235×10−20 cm2. The radiative lifetime τr is 209 μs and calculated fluorescence branch ratios are: β1(0.88μm)=45.2, β2(1.06μm)=46.7, β3(1.34μm)=8.1. The measured fluorescence lifetime τf is 137 μm and the quantum efficiency η is 65.6%. The absorption band at 808 nm wavelength has an FWHM of 20 nm. The absorption and emission cross sections are 3×10−20 and 6.13×10−20 cm2, respectively.  相似文献   

3.
We have used in situ scanning tunneling microscopy (STM) to study the facet formation in the selective growth of pyramidal Si nanocrystals on Si(0 0 1) windows in ultrathin 0.3-nm-thick SiO2 films. Broad (0 0 1) surfaces developed as the top of the crystals, and {1, 1, (2n+1)} (n=1–6) facets formed the sidewalls. As growth continued, the slope angle of sidewall facets increased, and {1, 1, 9} and {1, 1, (2m+1)} (0 <m < 4) facets often came to coexist on the sidewalls. On well-oriented Si(0 0 1) surfaces, layer-by-layer growth in the [0 0 1] direction was dominant. On vicinal Si(0 0 1) surfaces, lateral step growth took place in the initial stage, and the layer-by-layer growth was suppressed until after a large (0 0 1) surface had formed as the top of the crystal.  相似文献   

4.
Interconnecting cage-like porous structures of several halide compounds were prepared by the selective leaching of one eutectic phase method. The binary eutectic precursors were prepared by directional solidification using the Bridgman crystal growth technique. Porous NaMgF3 (40% pore volume), CaF2 (57% pore volume) and BaF2 (43% pore volume) crystals were obtained after water leaching the NaF component of the directionally solidified NaF/NaMgF3, NaF/CaF2 and NaF/BaF2 eutectics with the appropriate entangled microstructure. The growth conditions for eutectic-coupled growth and the morphology of the eutectics have been determined. In the coupled growth regime, the size of the eutectic phases “λ” is fairly uniform and varies with the eutectic growth rate “v” as λ2v=constant, which allows us to control the pore size within the 0.5–10 μm range. The simplicity and versatility of the eutectic growth also allows us to fabricate highly aligned porous structures at relatively high production rates.  相似文献   

5.
Phase diagrams of 1,2,4,5-tetrachlorobenzene–β-naphthol and 1,2,4,5-tetramethylbenzene–succinonitrile systems which are organic analogues of a nonmetal–nonmetal and a nonmetal–metal system, respectively, show the formation of a simple eutectic (melting point 103.7°C) with 0.71 mole fraction of β-naphthol in the former case and a monotectic (melting point 76.0°C) with 0.07 mole fraction of succinonitrile and a eutectic (melting point 52.5°C) with 0.97 mole fraction of succinonitrile in the latter case. The growth behaviour of the pure components, the eutectics and the monotectic studied by measuring the rate of movement of the solid–liquid interface in a capillary, suggests that the data obey the Hillig–Turnbull equation, v=uT)n, where v is the growth velocity, ΔT is the undercooling and u and n are constants depending on the nature of the materials involved. From the values of enthalpy of fusion determined by the DSC method using Mettler DSC-4000 system, entropy of fusion, interfacial energy, enthalpy of mixing and excess thermodynamic functions were calculated. The optical microphotographs of pure components and polyphase materials show their characteristic features.  相似文献   

6.
A micro-pulling-down process, using Ir crucibles and RF heating, has been used to grow single-crystal fiber and bulk crystals of Tb3Ga5O12 garnet (TGG). Single crystals ranging up to 450 mm in length have been produced. The crystals were 1–4 mm in diameter and were seeded-grown in the direction close to 1 1 1. The maximal crystal diameter achieved was 10 mm. Dependence of behavior of the solid–liquid interface on the growth parameters (temperature and pulling-rate) is discussed in detail.  相似文献   

7.
High-quality epitaxial YBa2Cu3O7−δ (YBCO) superconducting films with thicknesses between 0.2 and 2 μm were fabricated on (0 0 l) LaAlO3 with direct-current sputtering method. The influence of film thickness on the structure and texture was investigated by X-ray diffraction conventional θ–2θ scan and high-resolution reciprocal space mapping (HR-RSM). The films grew with strictly c-axis epitaxial, and no a-axis-oriented growth was observed up to a thickness of 2 μm. Lattice parameters of the YBCO films with different thicknesses were extracted from symmetry and asymmetry HR-RSMs. The X-ray lattice parameter method was used to determine the residual stress in YBCO films by measuring the a-, b-, c-axis strains, respectively. The results showed that YBCO films within thinner than 1 μm were under compressive stress, which was relieved increasing of film thickness. However, beyond 1 μm in thickness, YBCO films exhibited a tensile stress. Based on the experimental analysis, the variety of residual stresses in the films is mainly attributed to oxygen vacancies with thickness of YBCO film increasing.  相似文献   

8.
Micro-pulling-down (μ-PD) growth apparatus was modified for fluoride crystals. PrF3 was grown with various concentrations of Ce3+ from 0–100%. The crystals were transparent and colorless (CeF3) or greenish and 3 mm in diameter and 15–50 mm in length. Neither visible inclusions nor cracks were observed. Radioluminescence spectra and decay kinetics were measured for the sample set at room temperature. In comparison to the Czochralski or Bridgman method, the μ-PD method allows to produce single crystalline material in a faster thus more economic way. Once it is established for the fluoride crystals, it is an efficient tool for exploring the field of new functional fluorides.  相似文献   

9.
Calcium phosphate glass ceramics with incorporation of small additions of two nucleating agents, MgO and K2O were prepared in the metaphosphate and pyrophosphate region, using an appropriate two-step heat treatment of controlled crystallization defined by differential thermal analysis results. Identification and quantification of crystalline phases precipitated from the calcium phosphate glass were performed using X-ray diffraction and Rietveld analysis. The β-Ca2P2O7 (β-DCP), KCa(PO3)3, β-Ca(PO3)2 and Ca4P6O19 phases were detected in the glass ceramics. In order to evaluate the degradation of the glass ceramics prepared, degradation studies were carried out during 42 days in Tris-HCl solution at 37 °C, pH 7.4, using granules in the range of 355–415 μm. The materials presented a weight loss ranging up to 12%. The ions leached during the immersion mainly originated from the KCa(PO3)3 phase, probably due to the presence of K+ ion in the calcium metaphosphate, and the residual glassy phase. The structural changes at the surface of materials during degradation have been analyzed by Fourier transform infrared spectroscopy and X-ray diffraction. Results showed that significant surface changes occurred with immersion time, with the decrease of KCa(PO3)3, β-Ca2P2O7 and β-Ca(PO3)2 phases occurring at different periods of immersion. This study has demonstrated an easy way to prepared calcium phosphate materials with specific calcium phosphate phases and crystallization, and therefore specific degradation rates.  相似文献   

10.
Single crystals of ruby have been obtained from fluxed melts based on the systems Li2O–MoO3, Li2O–WO3, Na2O–WO3, 2PbO–3V2O5, PbO–V2O5–WO3, PbF2–Bi2O3 and Na3AlF6 by both the TSSG method and spontaneous crystallization at the temperatures 1330–900 °C. Al2O3 solubility has been measured for the flux composition of 2Bi2O3–5PbF2 in the temperature range 1200–1000 °C and dissolution enthalpy has been defined as 29.4 KJ/Mol. The composition of grown crystals was studied by electron microprobe analysis. The synthetic ruby contains from 0.51 to 6.38 at% of chromium admixture depending on the crystal growth conditions. Experimental results on growth conditions, composition and morphology of grown crystals are presented for each flux and temperature interval.  相似文献   

11.
We present for the first time, direct and clear experimental evidence of Al–O–Al and Si–O–Si linkages in charge-balanced aluminosilicate glasses with Si/Al=1, such as NaAlSiO4 (nepheline) and LiAlSiO4 (β-eucryptite) compositions using 17O triple quantum MAS (3QMAS) NMR and quantify the extent of disorder in framework cations (Si/Al). The degree of Al avoidance in NaAlSiO4 glass is 0.942 at 1050 K, and in LiAlSiO4 glass is 0.928 at 930 K (0.902 at 1050 K), which demonstrates the effect of cation field strength on the extent of disorder. In addition, we find a remarkable similarity between the ordering state of the liquid and that of the first, disequilibrium phase to crystallize.  相似文献   

12.
In this study, single-crystal γ-MnO2 nanowires have been successfully synthesized at room temperature in the absence of catalysts or templates, the diameter was found to be ca. 10–20 nm and the characteristic lengths up to several micrometers. The crystal phase of nanowires was confirmed by XRD and TEM measurements. Further, a dissolution– condensation–recrystallization process was proposed for the formation of nanowires under the room temperature condition.  相似文献   

13.
Long-wavelength vertical cavity surface emitting lasers (VCSELs) are considered the best candidate for the future low-cost reliable light sources in fiber communications. However, the absence of high refractive index contrast in InP-lattice-matched materials impeded the development of 1.3–1.5 μm VCSELs. Although wafer fusions provided the alternative approaches to integrate the InP-based gain materials with the GaAs/AlAs materials for their inherent high refractive index contrast, the monolithic InP-based lattice-matched distributed Bragg reflectors (DBRs) are still highly attractive and desirable. In this report, we demonstrate InP/InGaAlAs DBRs with larger refractive index contrast than InP/InGaAsP and InAlAs/InGaAlAs DBRs. The switching between InP and InGaAlAs layers and growth rate control have been done by careful growth interruption technique and accurate in situ optical monitoring in low-pressure metal organic chemical vapor deposition. A 35 pairs 1.55 μm centered InP/InGaAlAs DBRs has the stopband of more than 100 nm and the highest reflectivity of more than 99%. A VCSEL structure incorporating 35 pairs InP/InGaAlAs DBR as the bottom mirror combined with a 2λ thick periodic gain cavity and 10 pairs SiO2/TiO2 top dielectric mirrors was fabricated. The VCSELs lased at 1.56 μm by optical pumping at room temperature with the threshold pumping power of 30 mW.  相似文献   

14.
III–V semiconductor Indium Arsenide (InAs) nanocrystals embedded in silica glasses was synthesized by combining the sol–gel process and heat treatment in H2 gas. The size of InAs nanocrystals can be easily controlled via changing the In and As content in the starting materials and the heating temperature in a H2 gas atmosphere. Absorption measurements indicate a blue shift in energy with a reduction on the In and As content in the SiO2 gel glasses as a result of quantum confinement effects. A near-infrared photoluminescence with peak at 3.40 μm was observed at 6 K under 514.5 nm Ar+ laser excitation from InAs nanocrystals embedded in the silica gel glasses.  相似文献   

15.
Drastic effect of additional Mo on diamond nucleation was observed in the system of Mc60Cc40 (mixture of 60 mol% MgCO3 and 40 mol% CaCO3)–graphite at 7.7 GPa, 1700 and 1800°C for 1 h. The decrease of temperature required for nucleation of diamond by addition of Mo is 200°C, compared with the temperature in the system of Mc60Cc40–graphite (Sato et al., Diamond Related Mater. 8 (1999) 1900). These results indicate that a lower energy barrier is expected for the nucleation in the Mc60Cc40–Mo–graphite system. Therefore, this nucleation process in the system of Mc60Cc40–graphite–Mo can be considered to be different from the spontaneous nucleation in the system of Mc60Cc40–graphite. Two probable models for nucleation can be suggested: (1) heterogeneous nucleation in which nucleation is affected by MoC that is formed by the reaction between the carbonate and Mo and (2) achievement of enough concentration of carbon for formation of nuclei by the reaction between the carbonate and Mo.  相似文献   

16.
The single-crystalline β-wollastonite (β-CaSiO3) nanowires were prepared via a simple hydrothermal method, in the absence of any template or surfactant using cheap and simple inorganic salts as raw materials. Xonotlite [Ca6(Si6O17)(OH)2] nanowires were first obtained after hydrothermal treatment at a lower temperature of 200 °C for 24 h, and after being calcinated at 800 °C for 2 h, xonotlite nanowires completely transformed into β-wollastonite nanowires and the wire-like structure was preserved. The synthesized β-wollastonite nanowires had a diameter of 10–30 nm, and a length up to tens of micrometers, and the single-crystalline monoclinic parawollastonite structured β-wollastonite was identified by XRD with the space group of P21/a and cell constants of a=15.42 Å, b=7.325 Å, c=7.069 Å and β=95.38°. A possible growth mechanism of β-wollastonite nanowires was also proposed. The advantages of this method for the nanowire synthesis lie in the high yield, low temperature and mild reaction conditions, which will allow large-scale production at low cost.  相似文献   

17.
Er3+-doped and Er3+–Yb3+ co-doped yttrium aluminum borate (YAB) single crystals have been grown by the top-seeded solution growth method using a new flux system, namely NaF–MoO3–B2O3. The Er3+ concentrations were 1.3 mol% for both single doped and co-doped crystals and the Yb3+ concentration in the Er3+–Yb3+ co-doped crystal was 20.0 mol% in the raw materials. The distribution coefficients of Er3+ single doped and Er3+–Yb3+ co-doped crystals were measured. The polarized absorption and fluorescence spectra of Er3+–Yb3+ co-doped crystal were recorded and compared with those of Er3+ single doped crystal. The results demonstrate that Er3+–Yb3+ co-doped YAB crystal is a potential candidate for 1.55 μm laser materials.  相似文献   

18.
Using a perfect single crystal sample of CdTe grown using PVT method, the electronic charge transfer in the II–VI compound semiconductor CdTe at 200 and 300 K has been evaluated using two different approaches: (1) by solving a quadratic equation involving the observed structure factors of h+k+l=4n+2 type reflections; and (2) by a graphical approach in which the observed and calculated atomic form factors are extrapolated to sinθ/λ=0, to determine the transferred charge. Precise X-ray structure factors collected using MoK radiation have been used for the analysis. The results obtained are reasonable and clearly indicate the ionicity by which charge is transferred from Cd to Te in CdTe.  相似文献   

19.
Silica-based sol–gel glasses activated by Er3+ ions are attractive materials for integrated optics (IO) devices such as frequency upconverters and optical amplifiers. Monolithic erbium-activated silica xerogels with erbium content ranging from 0 up to 40 000 ppm were prepared by the sol–gel technique. Samples were densified by thermal treatment in air at 950°C for 120 h. The densification degree and the relative content of hydroxyl groups were studied by Raman spectroscopy. Refractive indices were measured at 632.8 and 543.5 nm by a prism coupling technique. Green to blue and violet upconversion luminescence upon continuous-wave excitation at 514.5 nm was observed for all samples. Emission at 1.5 μm, characteristic of the 4I13/24I15/2 transition of Er3+ ions, was observed at room temperature for all samples upon continuous-wave excitation at 980 nm. For the 5000 Er/Si ppm-doped xerogel, a photoluminescence was observed and a lifetime of 8 ms for the metastable 4I13/2 level was measured.  相似文献   

20.
Melts with the basic compositions 10Na2O · 10MgO · xAl2O3 · (80−x)SiO2 (x=0, 5, 10, 15 and 20), 10Na2O · xMgO · 10Al2O3 · (80−x)SiO2 (x=5, 10, 15 and 20) and xNa2O · 10MgO · 10Al2O3 · (80−x)SiO2 (x=5, 10 and 15) all doped with 0.25 mol% Fe2O3 were studied using square-wave voltammetry. The temperatures applied were in the range of 1000–1600 °C. The square-wave voltammograms recorded show peaks caused by the reduction of Fe3+ to Fe2+. The attributed peak potentials measured decreased linearly with decreasing temperatures. Increasing the MgO-concentration led to more negative peak potentials. Introducing alumina in the melt first resulted in less negative peak potentials. If the molar Al2O3-concentration is equal to that of Na2O (=10 mol%) the peak potentials are least negative. Further increase of the Al2O3-concentration led to more negative peak potentials. The variation of the Na2O-concentration led to a maximum in the peak potentials at an Na2O-concentration of 10 mol%. An empirical formula which allows the calculation of standard potentials from the chemical composition is proposed. Furthermore, a structural explanation for the effect of the chemical composition is given. Especially, the incorporation of Al2O3 as AlO4-tetrahedra at [Al2O3] < [Na2O] and as network modifier at larger concentrations was structurally explained by the similarities of Fe2+ and Mg2+, with respect to cation radii and metal–oxygen bond lengths.  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