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.
Summary Secondary Neutral Mass Spectrometry (SNMS) and X-Ray Diffraction (XRD) were used to find optimum parameters for the in-situ pulsed laser deposition of ZrO2/Y2O3 (YSZ) buffer layers on silicon (100) substrates. Homogeneous and nearly stoichiometric concentration depth profiles were found by SNMS for the laser deposited YSZ films. A peak of the SiO intensity during profiling of the YSZ/Si interface points to a SiO2 intermediate layer. An increasing Y-deficit of the YSZ films was found by decreasing the laser energy density at the target. Epitaxial growth of the YSZ thin films was observed at an oxygen partial pressure lower than 10–3 mbar, a substrate temperature of 600–800°C and a laser energy density at the target of about 8 J/cm2.  相似文献   

2.
A two-stage packing of a 30 wt.% Ga2O3/Al2O3 catalyst and a mixture catalyst of 30 wt.% Ga2O3/Al2O3 with 10 wt.% of Mn2O3 is shown to be quite efficient to NO selective reduction by a lower hydrocarbon under lean conditions. A high NO reduction conversion was observed in the temperature range of 200-600°C.  相似文献   

3.
A strong Al+ and a minor Ti+ peak without a proportional increase of the O+ signal in SNMS high-frequency sputtering mode (HFM) time profiles of an insulating μm-thick oxide layer on Ti-48Al-2Cr-2Nb led us to check for a possible contribution of positive secondary ions (SI+). SI+ and SI (negative secondary ions) can be detected in ion energy spectra. This is shown using Al+, O, AlO, and AlO2 ions sputtered from massive Al2O3. Similarly, and depending on stoichiometry, also Ti+ from mixed sintered, microscopically inhomogeneous Al2O3-TiO2-SiO2 pellets has been identified to be partly SI+. The subtraction of an assumed contribution of ionized secondary neutrals (SN+) suggests that SI+ may form several 10% of the detected ions obtained in the HFM sputtering and plasma processes. However, the positive surface potential of some 10 V being necessary to cause detectable SI+ contributions does not build up on μm-thin insulating layers. Therefore, we have to conclude that the Al+ and Ti+ peaks in the sputter time profiles of the μm-thick oxide layer on Ti-48Al-2Cr-2Nb which are accompanied by an O+ deficiency cannot have been caused by SI+. Instead, their more probable origin is the inhomogeneous Al2O3 interlayer itself. Together with the residues of a topmost TiO2 layer which has strongly been depleted in O by preferential sputtering, the relative O+ deficiency may be explained without assuming SI+ contributions. Received: 22 February 1999 / Revised: 1 July 1999 / Accepted: 6 July 1999  相似文献   

4.
The Er3+-doped Al2O3 nanopowders have been prepared by the sol-gel method, using the aluminium isopropoxide [Al(OC3H7)3]-derived γ-AlOOH sols with addition of the erbium nitrate [Er(NO3)3·5H2O]. The five phases of γ-(Al,Er)2O3, θ-(Al,Er)2O3, α-(Al,Er)2O3, ErAlO3, and Al10Er6O24 were detected with the 0–20 mol% Er3+-doped Al2O3 nanopowders at the different sintering temperature of 600–1200°C. The average grain size was increased from about 5 to 62 nm for phase transformation of undoped γ-Al2O3→α-Al2O3 at the sintering temperature from 600 to 1200°C. At the same sintering temperature, average grain size was decreased with increase of the Er3+ doping concentration. Infrared absorption spectra of γ-Al2O3 and θ-Al2O3 nanopowders showed the two broad bands of 830–870 and 550–600 cm−1, the three broad bands of 830–870, 750–760, and 550–600 cm−1, respectively. The infrared absorption spectra for the α-Al2O3 nanopowder showed three characteristic bands, 640, 602, and 453 cm−1. The two characteristic bands of 669 and 418 cm−1 for Er2O3 clusters were observed for the Er3+-doped Al2O3 nanopowders when Er3+ doping concentration was increased up to 2 mol%. The 796, 788, 725, 692, 688, 669, 586, 509, 459, and 418 cm−1 are the characteristic bands of Al10Er6O24 phase.  相似文献   

5.
A strong Al+ and a minor Ti+ peak without a proportional increase of the O+ signal in SNMS high-frequency sputtering mode (HFM) time profiles of an insulating μm-thick oxide layer on Ti-48Al-2Cr-2Nb led us to check for a possible contribution of positive secondary ions (SI+). SI+ and SI (negative secondary ions) can be detected in ion energy spectra. This is shown using Al+, O, AlO, and AlO2 ions sputtered from massive Al2O3. Similarly, and depending on stoichiometry, also Ti+ from mixed sintered, microscopically inhomogeneous Al2O3-TiO2-SiO2 pellets has been identified to be partly SI+. The subtraction of an assumed contribution of ionized secondary neutrals (SN+) suggests that SI+ may form several 10% of the detected ions obtained in the HFM sputtering and plasma processes. However, the positive surface potential of some 10 V being necessary to cause detectable SI+ contributions does not build up on μm-thin insulating layers. Therefore, we have to conclude that the Al+ and Ti+ peaks in the sputter time profiles of the μm-thick oxide layer on Ti-48Al-2Cr-2Nb which are accompanied by an O+ deficiency cannot have been caused by SI+. Instead, their more probable origin is the inhomogeneous Al2O3 interlayer itself. Together with the residues of a topmost TiO2 layer which has strongly been depleted in O by preferential sputtering, the relative O+ deficiency may be explained without assuming SI+ contributions. Received: 22 February 1999 / Revised: 1 July 1999 / Accepted: 6 July 1999  相似文献   

6.
V2O5/Al2O3 solids of varying compositions were prepared, dried at 100°C and calcined in air at 400–1000°C. The solid-solid interactions between the mixed oxides were investigated by means of DTA, TG and XRD techniques.The results revealed that ammonium metavanadate and aluminium hydroxide decomposed at 260 and 290°C, respectively, to yield an ammonium vanadium intermediate compound and Al2O3 as solids. The intermediate compound readily decomposed at 360°C to give V2O5.Solid V2O5 generally catalyses the crystallization of Al2O3 to an extent proportional to its amount present. The solid-solid interactions between Al2O3 and V2O3 to produce AlV2O4 and AlVO4 took place at 750 and 900°C. These solids decomposed entirely at 1000°C, producing V2O5 and alpha-corundum. The pure Al2O3 samples employed existed as amorphous solids even when heated in air at 750°C, but in the presence of V2O5 (7–18 wt.%) they crystallized to thetaalumina at 600°C. The pure solid alumina crystallized at 1000°C to a mixture of theta and kappa-alumina. In the presence of V2O5, alpha-corundum together with kappa and theta phases was obtained on heating at 900°C.  相似文献   

7.
Summary Improved methods for Al2O3 metallization by Cu are described. Good adhesion between Cu and Al2O3 substrate depends on the formation of chemical bonds between the substrate and the metallic layer. The temperature needed for the formation of a CuAl2O4 spinel interface is reduced from 1050°C to 900°C by the addition of various oxides. The adhesion between the CuAl2O4 interface and deposited Cu is stronger then the tensile strength of pure Cu. Plasma techniques for the formation of a Cu containing interface are also described. Bombardment of a Cu film with Xe+ ions in a rf-glow discharge implants Cu atoms into the substrate to a depth of 5 nm, as determined by SIMS depth profiling. Methods for reduction of the CuAl2O4 surface for subsequent metallization are also presented.  相似文献   

8.
Series of compositions Bi2(M′xM1−x)4O9 with x=0.0, 0.1,…, 1.0 and M′/M=Ga/Al, Fe/Al and Fe/Ga were synthesized by dissolving appropriate amounts of corresponding metal nitrate hydrates in glycerine, followed by gelation, calcination and final heating at 800 °C for 24 h. The new compositions with M′/M=Ga/Al form solid-solution series, which are isotypes to the two other series M′/M=Fe/Al and Fe/Ga. The XRD data analysis yielded in all cases a linear dependence of the lattice parameters related on x. Rietveld structure refinements of the XRD patterns of the new compounds, Bi2(GaxAl1−x)4O9 reveal a preferential occupation of Ga in tetrahedral site (4 h). The IR absorption spectra measured between 50 and 4000 cm−1 of all systems show systematic shifts in peak positions related to the degree of substitution. Samples treated in 18O2 atmosphere (16 h at 800 °C, 200 mbar, 95% 18O2) for 18O/16O isotope exchange experiments show a well-separated IR absorption peak related to the M-18Oc-M vibration, where Oc denotes the common oxygen of two tetrahedral type MO4 units. The intensity ratio of M-18Oc/M-16Oc IR absorption peaks and the average crystal sizes were used to estimate the tracer diffusion coefficients of polycrystalline Bi2Al4O9 (D=2×10−22 m2s−1), Bi2Fe4O9 (D=5×10−21 m2s−1), Bi2(Ga/Al)4O9 (D=2×10−21 m2s−1) and Bi2Ga4O9 (D=2×10−20 m2s−1).  相似文献   

9.
Zusammenfassung Beim Erhitzen gefällter Kieselsäure zeigen sprunghafte Änderungen in der Emanationsabgabe Umwandlungen bei etwa 5700 bzw. 8700 C an.Calciumoxyd reagiert mit SiO2 erst nach 9000 C.Die Umwandlung eines -Al2O3 in Korund erfolgt zwischen 8400 und 9500 C. Diese Umwandlang schreitet von außen nach dem Inneren des Al2O3-Partikels vor; daraus wird auf einen größeren Energieinhalt der Atome an der Oberfläche der Partikeln gegenüber jenen in deren Innerem geschlossen.Kobaltcarbonat zeigt auf Aluminiumhydroxyd bis etwa 4000 C keinen Einfluß; zwischen 4000 und 8000 C tritt eine den Austritt der Emanationsatome stark blockierende Wirkung ein. Die Oberfläche des entstehenden -Al2O3 wird stark verdeckt bis nach 8000 C, also beim Beginn der —-Umwandlung, ein E.V.-Anstieg einsetzt: Das Kobaltoxyd reagiert mit dem Al2O3 unter Bildung von Thenardsblau und der Beginn der Reaktion ist mit einer starken Auflockerung verbunden.CaCO3 wirkt auf die Oberfläche von Al2O3 nach 6000 C stark ein. Zu Beginn der Reaktion von CaO mit Al2O3 tritt eine starke Auflockerung der Al2O3-Oberfläche ein.Die Bildung von Calciumferrit erfolgt zwischen 5500 und 6000 C unter erheblicher Erhöhung der Dispersität beim Beginn der Reaktion.Für Unterstützungen zu großem Dank verpflichtet bin ich den Herren Professoren H.Mark und St.Meyer.  相似文献   

10.
The structure of Ga2O3–Al2O3 supports and Pd/Ga2O3–Al2O3 catalysts and the performance of these catalysts in liquid-phase acetylene hydrogenation have been investigated. The deposition of Ga(NO3)3 onto Al2O3 by impregnation followed by calcination of the impregnated support at 600°C yields γ-Ga2O3–Al2O3 solid solutions containing up to 50 wt % Ga2O3. X-ray diffraction characterization of model palladium catalysts and their temperature-programmed reduction with hydrogen have demonstrated that, while palladium in Pd/Ga2O3 is in the form of a Pd2Ga alloy, in the Pd/γ-Ga2O3–Al2O3 catalyst there is no direct interaction between PdО and Ga2O3 particles and palladium is in the monometallic state. The introduction of 10–20 wt % gallium oxide into Al2O3 lowers the activity of the supported palladium catalyst relative to that of the initial Pd/Al2O3 but increases the ethylene yield by enhancing the ethylene formation selectivity.  相似文献   

11.
A Ga2O·11Al2O3 nanonet was synthesized by using Ga2O3 powder as the precursor to generate Ga2O vapor in H2 atmosphere which further reacted with Al2O3 at 730 °C to form Ga2O·11Al2O3 at the interfaces of a porous anodic aluminum oxide (AAO) template. The prepared Ga2O·11Al2O3 nanonet then served as a Ga2O-stablizing reservoir to fabricate single crystal GaN nanowires. The residual Ga2O3 powder at the surface of the produced Ga2O·11Al2O3 nanonet and the metallic Ga or Ga2O from the Ga2O·11Al2O3 decomposition reacted with ammonia to yield GaN nanowires at 780 °C. The reaction mechanisms were investigated.  相似文献   

12.
Phase transitions in MgAl2O4 were examined at 21-27 GPa and 1400-2500 °C using a multianvil apparatus. A mixture of MgO and Al2O3 corundum that are high-pressure dissociation products of MgAl2O4 spinel combines into calcium-ferrite type MgAl2O4 at 26-27 GPa and 1400-2000 °C. At temperature above 2000 °C at pressure below 25.5 GPa, a mixture of Al2O3 corundum and a new phase with Mg2Al2O5 composition is stable. The transition boundary between the two fields has a strongly negative pressure-temperature slope. Structure analysis and Rietveld refinement on the basis of the powder X-ray diffraction profile of the Mg2Al2O5 phase indicated that the phase represented a new structure type with orthorhombic symmetry (Pbam), and the lattice parameters were determined as a=9.3710(6) Å, b=12.1952(6) Å, c=2.7916(2) Å, V=319.03(3) Å3, Z=4. The structure consists of edge-sharing and corner-sharing (Mg, Al)O6 octahedra, and contains chains of edge-sharing octahedra running along the c-axis. A part of Mg atoms are accommodated in six-coordinated trigonal prism sites in tunnels surrounded by the chains of edge-sharing (Mg, Al)O6 octahedra. The structure is related with that of ludwigite (Mg, Fe2+)2(Fe3+, Al)(BO3)O2. The molar volume of the Mg2Al2O5 phase is smaller by 0.18% than sum of molar volumes of 2MgO and Al2O3 corundum. High-pressure dissociation to the mixture of corundum-type phase and the phase with ludwigite-related structure has been found only in MgAl2O4 among various A2+B3+2O4 compounds.  相似文献   

13.
A green BaZr0.1Ce0.7Y0.2O3−δ (BZCY) electrolyte layer was deposited on porous anode substrate (BZCY:NiO = 35:65, in weight ratio) by a suspension spray. In this process, the suspension was prepared by directly ball-milling the mixed BaCO3, CeO2, ZrO2 and Y2O3 powders in ethanol for 24 h. Then the bi-layers were co-sintered at 1400 °C for 5 h in air to obtain dense and uniform electrolyte membrane in the thickness of 10 μm. With Nd0.7Sr0.3MnO3−δ cathode, a fuel cell was assembled. It was tested from 600 °C to 700 °C using humid hydrogen as fuel and air as oxidant. The cell at 700 °C exhibited 1.02 V for open circuit voltage (OCV), 450 mW/cm2 for peak output and 0.18 Ω cm2 for electrode polarizations under open circuit conditions, respectively. The results indicate that it is feasible to fabricate thin electrolyte membrane for solid oxide fuel cells (SOFCs) by this simple, cost-effective and efficient technique.  相似文献   

14.
Based on powder X-ray diffraction and 31P Magic Angle Spinning Nuclear Magnetic Resonance (MAS NMR) investigations of mixed phosphate Al0.5Ga0.5PO4, prepared by co-precipitation method followed by annealing at 900 °C for 24 h, it is shown that Al0.5Ga0.5PO4 phase crystallizes in hexagonal form with lattice parameter a=0.491(2) and c=1.106(4) nm. This hexagonal phase of Al0.5Ga0.5PO4 is similar to that of pure GaPO4. The 31P MAS NMR spectrum of the mixed phosphate sample consists of five peaks with systematic variation of their chemical shift values and is arising due to existence of P structural units having varying number of the Al3+/Ga3+ cations as the next nearest neighbors in the solid solution. Based on the intensity analysis of the component NMR spectra of Al0.5Ga0.5PO4, it is inferred that the distribution of Al3+ and Ga3+ cations is non-random for the hexagonal Al0.5Ga0.5PO4 sample although XRD patterns showed a well-defined solid solution formation.  相似文献   

15.
Compounds A3+Te6+M33+X25+O14 (A = Na, K; M = Ga, Al, Fe; X = P, As, V) with the Ca3Ga2Ge4O14 structure (sp. gr. P321) were prepared by solid-phase synthesis at 600–850°C in air. The compounds melt incongruently or decompose in the solid state.  相似文献   

16.
Ag colloid-containing coatings on soda lime glass and fused silica are prepared via the sol-gel process. To incorporate Ag+-ions in the coatings homogeneously, they are stabilized by a functionalised silane (aminosilane) and then mixed with the basic sol prepared from 3-glycidoxypropyl trimethoxysilane (GPTS) and tetraethoxysilane (TEOS). Crack-free and transparent coatings with a thickness of 0.5 to 1.2 m, are obtained by heat treatment between 120°C and 600°C. The Ag-colloid formation was monitored by UV-VIS spectroscopy as a function of temperature. The investigations reveal that the substrate has a deciding influence on the Ag-colloid formation caused by alkali diffusion from the substrate into the coating. High resolution transmission electron microscopy (HRTEM) investigations prove that poly-crystalline AgxOy-nanoparticles are formed during thermal densification in the coatings and that this change is accompanied by a vanishing of the yellow colour of the coatings. A post-heat treatment in a reducing atmosphere (90% N2, 10% H2) turns back the yellow colour and single-crystalline Ag-colloids can be detected by HRTEM. A suitable choice of the temperature and time conditions allows the control of the colloid size during heat treatment in a reducing atmosphere. For comparison, ion-exchange experiments have been carried out which showed that a spontaneous Ag-colloid formation was achieved in the soda lime substrate at 400°C. Since Ag containing SiO2-coatings remained colourless after thermal treatment between 400°C and 600°C in air, on soda lime substrates, a remarkable diffusion of Ag+ into the substrate was excluded.  相似文献   

17.
Novel γ-Al2O3 supported nickel (Ni/Al2O3) catalyst was developed as a functional layer for Ni–ScSZ cermet anode operating on methane fuel. Catalytic tests demonstrated Ni/Al2O3 had high and comparable activity to Ru–CeO2 and much higher activity than the Ni–ScSZ cermet anode for partial oxidation, steam and CO2 reforming of methane to syngas between 750 and 850 °C. By adopting Ni/Al2O3 as a catalyst layer, the fuel cell demonstrated a peak power density of 382 mW cm?2 at 850 °C, more than two times that without the catalyst layer. The Ni/Al2O3 also functioned as a diffusion barrier layer to reduce the methane concentration within the anode; consequently, the operation stability was also greatly improved without coke deposition.  相似文献   

18.
A series of Bi2(GaxAl1−x)4O9 solid solutions (0≤x≤1), prepared by mechanochemical processing of Bi2O3/Ga2O3/Al2O3 mixtures and subsequent annealing, was investigated by XRD, EDX, and 27Al MAS NMR. The structure of the Bi2(GaxAl1−x)4O9 solid solutions is found to be orthorhombic, space group Pbam (No. 55). The lattice parameters of the Bi2(GaxAl1−x)4O9 series increase linearly with increasing gallium content. Rietveld refinement of the XRD data as well as the analysis of the 27Al MAS NMR spectra show a preference of gallium cations for the tetrahedral sites in Bi2(GaxAl1−x)4O9. As a consequence, this leads to a far from random distribution of Al and Ga cations across the whole series of solid solutions.  相似文献   

19.
The pressure dependences of the self-diffusion coefficients of deuterium oxide in 4.5m solutions of LiCl–D2O and CsCl–D2O (also 7m) and 3.06m CaCl2–D2O have been measured by the NMR spin-echo method at 30°C, 60°C, and 90°C. Shear viscosities and densities of these solutions have also been determined over the same range of experimental conditions. The experimental data show that the diffusion constantD decreases with the increasing structure-making ability of the electrolyte cation Ca+2>Li+. In contrast, the diffusion coefficient for D2O in the 4.5 and 7m CsCl solutions is equal to that for pure D2O at 30°C but lower at 60°C and 90°C. It has been found that the Stokes-Einstein equation relates well the diffusion coefficients to shear viscosity in these concentrated electrolyte solutions.  相似文献   

20.
By precipitation with ammonia of ethanolic solutions containing the appropriate proportions of gallium and aluminium nitrate, following by calcination of the resulting gels at 773 K, mixed Ga2O3/Al2O3 oxides having Ga:Al ratios of 9:1, 4:1, 1:1, 1:4 and 1:9 were obtained. Powder X‐ray diffraction showed that these mixed metal oxides form a series of solid solutions having the spinel‐type structure; also shown by γ‐Al2O3 and γ‐Ga2O3. The specific surface area (determined by nitrogen adsorption at 77 K) was found to range from 160 m2 g?1 for the mixed oxide having Ga:Al = 9:1 up to 370 m2 g?1 for that having Ga:Al = 1:9. High resolution MAS NMR showed that Ga3+ and Al3+ ions occur at both tetrahedral and octahedral sites in the spinel‐type structure of the mixed metal oxides, although there is a preferential occupation of tetrahedral sites by Ga3+ ions. A proportion of penta‐coordinated Al3+ ions was also found. IR spectra of carbon monoxide adsorbed at 77 K showed that the mixed metal oxides have a considerable Lewis acidity, related mainly to tetrahedrally coordinated metal ions exposed at crystal surfaces. The characteristic infrared absorption band of coordinated (adsorbed) CO appears in the range 2205–2190 cm?1, and its peak wavenumber is nearly independent of Ga:Al ratio in the mixed gallia‐alumina oxides.  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