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 343 毫秒
1.
In order to improve the hot corrosion resistance of yttria-stabilized zirconia (YSZ), an Al2O3 overlay has been deposited on the surface of YSZ by electron-beam physical vapor deposition. Hot corrosion tests have been performed on the YSZ coatings with and without an Al2O3 overlay in the molten salt mixture (Na2SO4+0–15 wt% V2O5) at 950°C. The presence of V2O5 in the molten salt exacerbates degradation of both the monolithic YSZ coating and the composite YSZ/Al2O3 system. The formation of a low-melting Na2O–V2O5–Al2O3 liquid phase is responsible for degradation of the Al2O3 overlay. The Al2O3 overlay acts as a barrier against the infiltration of the molten salt into the YSZ coating during exposure to the molten salt mixture with <5 wt% vanadate.  相似文献   

2.
Crack Bifurcation in Laminar Ceramic Composites   总被引:2,自引:0,他引:2  
Crack bifurcation was observed in laminar ceramic composites when cracks entered thin Al2O3 layers sandwiched between thicker layers of Zr(12Ce)O2. The Al2O3 layers contained a biaxial, residual, compressive stress of ∼2 GPa developed due to differential contraction upon cooling from the processing temperature. The Zr(12Ce)O2 layers were nearly free of residual, tensile stresses because they were much thicker than the Al2O3 layers. The ceramic composites were fabricated by a green tape and codensification method. Different specimens were fabricated to examine the effect of the thickness of the Al2O3 layer on the bifurcation phenomena. Bar specimens were fractured in four-point bending. When the propagating crack encountered the Al2O3 layer, it bifurcated as it approached the Zr(12Ce)O2/ Al2O3 interface. After the crack bifurcated, it continued to propagate close to the center line of the Al2O3 layer. Fracture of the laminate continued after the primary crack reinitiated to propagate through the next Zr(12Ce)O2 layer, where it bifurcated again as it entered the next Al2O3 layer. If the loading was stopped during bifurcation, the specimen could be unloaded prior to complete fracture. Although the residual stresses were nearly identical in all Al2O3 layers, crack bifurcation was observed only when the layer thickness was greater than ∼70 μm.  相似文献   

3.
High reflectance thermal barrier coatings consisting of 7% Yittria-Stabilized Zirconia (7YSZ) and Al2O3 were deposited by co-evaporation using electron beam physical vapor deposition (EB-PVD). Multilayer 7YSZ and Al2O3 coatings with fixed layer spacing showed a 73% infrared reflectance maxima at 1.85 μm wavelength. The variable 7YSZ and Al2O3 multilayer coatings showed an increase in reflection spectrum from 1 to 2.75 μm. Preliminary results suggest that coating reflectance can be tailored to achieve increased reflectance over a desired wavelength range by controlling the thickness of the individual layers. In addition, microstructural enhancements were also used to produce low thermal conductive and high hemispherical reflective thermal barrier coatings (TBCs) in which the coating flux was periodically interrupted creating modulated strain fields within the TBC. TBC showed no macrostructural differences in the grain size or faceted surface morphology at low magnification as compared with standard TBC. The residual stress state was determined to be compressive in all of the TBC samples, and was found to decrease with increasing number of modulations. The average thermal conductivity was shown to decrease approximately 30% from 1.8 to 1.2 W/m-K for the 20-layer monolithic TBC after 2 h of testing at 1316°C. Monolithic modulated TBC also resulted in a 28% increase in the hemispherical reflectance, and increased with increasing total number of modulations.  相似文献   

4.
This paper reports ionic conductivity of yttria-stabilized zirconia (YSZ)–Al2O3 composite membranes. The tape cast specimens were subjected to binder burnout (500°C) and sintering (1550°C) processes to obtain 200–300 μm thick membranes. The ionic conductivity and microstructure of the membranes were characterized and are discussed in this paper. The ionic conductivity of the composite specimens was enhanced and was correlated with the number of charge carrier and their mobility. The solubility of Al2O3 in YSZ was minimal and nanosize Al2O3 of the batch sintered into microsize and existed as a distinct phase. The scanning electron microscopy micrographs revealed that YSZ and Al2O3 grains were strained.  相似文献   

5.
The feasibility of preparing a thin layer of α-Al2O3 on the surface of a single-crystal, Ni-based superalloy was examined using a chloride-based chemical vapor deposition (CVD) process previously developed for cutting tool applications. A coating directly deposited by this method on the alloy surface consisted of ∼1 μm α-Al2O3 crystals in a matrix of amorphous Al2O3. When the alloy surface was predeposited with an electroplated Pt layer, the coating was mostly α-Al2O3, but with the presence of fine microcracks on the coating surface. In comparison to the results observed for pure Pt substrate, the role of the Pt interlayer was apparently to promote the rapid formation of κ-Al2O3 nuclei, which subsequently transformed to α-Al2O3 during the CVD growth process.  相似文献   

6.
Alumina Dissolution into Silicate Slag   总被引:1,自引:0,他引:1  
Dissolution of commercial white fused and tabular Al2O3 grains into a model silicate slag was investigated after 1 h at 1450° and 1600°C. Formation of CA6 and hercynitic spinel layers was observed at all Al2O3/slag interfaces. The spinel layer was not always continuous, and so, compared with the CA6 layer, it had a less-significant effect on the dissolution process. The CA6 layer that formed adjacent to the tabular Al2O3 was incomplete at both temperatures, so that its dissolution was not a totally indirect process. These incomplete CA6 and spinel layers meant that slag penetrated into the tabular Al2O3 grains, which, thus, were corroded and disintegrated by the penetrating slag. There was evidence of liquid in the CA6 layer adjacent to the fused Al2O3 after 1 h at 1450°C, which also enabled direct dissolution. After 1 h at 1600°C, fused Al2O3 revealed a thick (∼60 μm), continuous and unpene-trated CA6 layer, indicating fully indirect dissolution at this temperature.  相似文献   

7.
Mechanisms and Kinetics of Reaction-Bonded Aluminum Oxide Ceramics   总被引:1,自引:0,他引:1  
Reaction-bonded Al2O3 (RBAO) ceramics were fabricated starting from mechanically alloyed Al2O3/Al, Al2O3/ Al/ZrO2, and Al2O3/Al/ZrO2/Zr mixtures. Isopressed compacts were heat-treated in air up to 1550°C. Reaction-bonding mechanisms, kinetics, and the influence of ZrO2 and Zr additions are investigated. Independent of additive, oxidation of Al proceeds both as solid/gas and liquid/gas reaction, and the reaction kinetics follow a parabolic rate law. The reaction rate depends strongly on the particle size of Al. The activation energy of the reaction depends essentially on green density. Below the melting temperature of Al, in samples containing 45 vol% Al and 55 vol% Al2O3, it is 112 and 152 kJ/mol at ∼64% and ∼74% TD, respectively, while above the melting temperature, it lies in the range ∼ 26–33 kJ/mol. Zr additions reduce the activation energy to some extent. Samples with only ZrO2 additions exhibit nearly the same activation energies as ZrO2-free samples, though ZrO2 has a very positive effect on the microstructural development in RBAO ceramics. Microstructure evolution and some strength data of RBAO bodies are also reported.  相似文献   

8.
In part I of this work, it was found that titanium (Ti) wire encapsulated within mechanically milled alumina powder and sintered at 1350°C forms potentially useful microcavities due to the consolidation of Kirkendall porosity. Here a series of samples sintered at 1350°C in the range 0–24 h has shown the remarkable way in which these cavities form. The cavity has already started in samples quenched from the top of the heating ramp (0 min at 1350°C). It is surrounded by a diffusion zone ∼300 μm in diameter, which does not change size throughout the firing process although the contents change markedly. The diffusion zone microstructure is initially complex with phase sequence TiO2/Al2O3/TiO2+Al2O3/Al2TiO5. Microstructure evolution may be summarized as outward growth of the cavity accompanied by inward growth of the Al2TiO5 resulting in a ∼190-μm-diameter cavity surrounded by a 50-μm-thick layer of Al2TiO5. The formation of the cavity and surrounding microstructure is discussed although some features, such as the nucleation of Al2TiO5 in the part of the diffusion zone furthest from the Ti source and the ring of Al2O3, which persists in between Ti-rich parts of the diffusion zone are still poorly understood.  相似文献   

9.
A two-step ion-exchange technique was developed for introducing compressive stresses on the surface of ZrO2–Al2O3 composites. In the first step, a thin layer (∼250 μm) of Na-β"-Al2O3 was formed on the surface of the composite by a vapor-phase process at ∼1400°C. In the second step, Na+ ions were replaced by K+ ions by a heat treatment at ∼385°C for 2 h in a molten KNO3 bath. Replacement of sodium by potassium led to the creation of surface compressive stresses. The flexural strength and Weibull modulus of ZrO2–Al2O3 composite were ∼915 MPa and 10, respectively, for the as-sintered samples. By contrast, the flexural strength and Weibull modulus were ∼1140 MPa and 26, respectively, for the ion-exchanged samples. A residual surface compressive stress of ∼480 MPa was measured by a strain-gauge technique in K+-ion-exchanged samples. The presence of surface compressive stresses also was confirmed using an indentation technique. The technique developed here can be used to introduce compressive stresses on components of virtually any shape.  相似文献   

10.
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   

11.
Al2O3–ZrO2–SiC whisker composites were prepared by surface-induced coating of the precursor for the ZrO2 phase on the kinetically stable colloid particles of Al2O3 and SiC whisker. The fabricated composites were characterized by a uniform spatial distribution of ZrO2 and SiC whisker phases throughout the Al2O3 matrix. The fracture toughness values of the Al2O3–15 vol% ZrO2–20 vol% SiC whisker composites (∼12 MPa.m1/2) are substantially greater than those of comparable Al2O3–SiC whisker composites, indicating that both the toughening resulting from the process zone mechanism and that caused by the reinforced SiC whiskers work simultaneously in hot-pressed composites.  相似文献   

12.
The scavenging of a resistive siliceous phase via the addition of Al2O3 was studied, using imaging secondary-ion mass spectroscopy (SIMS), given the improved grain-boundary conductivity in 8-mol%-yttria-stabilized zirconia (8YSZ). The grain-boundary resistivity in 8YSZ decreased noticeably with the addition of 1 mol% of Al2O3. Strong SiO2 segregation at the grain boundaries was observed in a SIMS map of pure 8YSZ that contained 120 ppm of SiO2 (by weight). The addition of 1 mol% of Al2O3 caused the SiO2 to gather around the Al2O3 particles. The present observations provided direct and visual evidence of SiO2 segregation at the grain boundaries (which had a deleterious effect on grain-boundary conductivity) and the scavenging of SiO2 via Al2O3 addition.  相似文献   

13.
Anorthite-glass films were grown on basal Al2O3 substrates using pulsed-laser deposition. The substrates were cleaned and annealed in air at 1400°C to produce crystallographically flat (0001) terraces. The films were deposited in an oxidizing environment. X-ray microanalysis confirmed the composition of the glass films to be close to that of anorthite (CaO·Al2O3·2SiO2). Although anorthite usually has triclinic symmetry, subsequent crystallization of these films in air at 1200°C resulted in the formation of pseudo-orthorhombic CaAl2Si2O8 ( o -anorthite), a known metastable form of the mineral. Microstructural characterization was performed using visible-light microscopy, scanning electron microscopy, and transmission electron microscopy. The films dewetted the substrate either before or after crystallization to form o -anorthite islands which had strong orientation relationships to the Al2O3 substrate. The epitaxy of the o -anorthite islands was accompanied by a small lattice mismatch parallel to the substrate plane. The formation of three orientational variants is consistent with the symmetry of the basal Al2O3 surface. The dislocation network observed at the o -anorthite/Al2O3 interface indicates that nucleation and growth of the anorthite occurs directly on the substrate surface without an intervening interfacial amorphous layer. The study of anorthite-glass films is important because they are present in liquid-phase-sintered Al2O3, and may be devitrified by postsintering heat treatments.  相似文献   

14.
Poly(acrylic acid) (PAA) dispersant concentration, suspension pH, and Al2O3 solids loading effects on PAA adsorption onto Al2O3 nanoparticles were studied; the stability and rheology of the Al2O3 nanoparticle suspensions were examined. The most desirable suspension conditions were 7.5–9.5 for pH and 2.00–2.25 wt% of Al2O3 for the PAA concentration. Electrical double-layer thickness and PAA adsorption layer thickness comparison showed that electrosteric stabilization was dominant. 45.0 vol% Al2O3 solids loading can be achieved for freeze casting. The maximum solids loading was predicted to be 50.7 vol%. The freeze-cast sample showed that pre-rest before freezing was critical for achieving desirable microstructures.  相似文献   

15.
An electrophoretic deposition and sintering route was used to prepare YSZ/Al2O3 composites with a compositional gradient. The YSZ content was continuously decreased from the YSZ-rich surface to the Al2O3-rich surface, Microstructural and Vickers hardness (16–24 GPa) evidence tracked the compositional development, and the indentation fracture toughness was found to vary across the section (10–3 MPa·m1/2).  相似文献   

16.
The sintering behavior and electrical conductivity of high-purity 8-mol% Y2O3-stabilized ZrO2 (8YSZ) with Al2O3 additions were investigated. The addition of 1 wt% AI2O3 to 8YSZ provided dense, sintered samples with 9.1% relative density at 1400°C without a holding time. Addition of 1 wt% SiO2 enhanced the sinterability of 8YSZ. Na2O addition of 0.1 wt% remarkably lowered it. Electrical conductivity at 1000°C in air increased slightly with increased Ai2O3 content up to 1 wt% and then monotonously decreased. 8YSZ with 1 wt% AI2O3 showed the maximum conductivity of 0.16 S/cm at 1000°C.  相似文献   

17.
Polyacrylic acid (PAA) is known to be an effective dispersant for Al2O3 powder in aqueous media. However, at high solid loading (>55 vol%), the dispersion of the Al2O3 suspensions became difficult with only PAA as a dispersant. In this paper, ethylenediaminetetraacetic acid, tetrasodium salt, dihydrate (EDTA-4Na) was introduced to improve the dispersion of the Al2O3 suspensions. With the aid of EDTA-4Na, the adsorption amount of sodium polyacrylic acid (PAA-Na) increased, while the apparent viscosity of 60 vol% Al2O3 slurries decreased significantly. Particle size measurements showed that EDTA-4Na could help to reduce larger agglomerates, possibly by modifying the adsorbed layer thickness. The interactions between EDTA-4Na and PAA-Na were studied using Fourier-transform infrared spectroscopy analysis. Results showed that it was possible to introduce EDTA-4Na as the second dispersant to improve the dispersion of high solid content Al2O3 slurries.  相似文献   

18.
The solid solution range of melt-grown mullite was examined by crystal-chemical methods. The maximum Al2O3 content as determined by EDX was ∼83.6 wt%, 75 mol%, or the nominal composition 3Al2O3.SiO2. For samples of overall composition 81 to 83 wt% Al2O3, extra lines indicating crystallographic superstructure appeared in Guinier X-ray patterns. The corresponding TEM microstructure consisted of a mullite matrix finely twinned on (001), the twins being 0.02 to 0.10 μm wide, with oriented exsolution of α-A12O3, often twinned, also being present. The analogy between mullite superstructure and that of plagioclase feldspars, as well as the relevance of these findings to the SiO2-Al2O3 metastable phase equilibria are discussed.  相似文献   

19.
Two different zirconia-alumina composites, ZTA-30 (70 wt% Al2O3+30 wt% ZrO2) and ZTA-60 (40 wt% Al2O3+60 wt% ZrO2), with potential for orthopedic applications, were processed in aqueous media and consolidated by slip casting (SC), hydrolysis-assisted solidification (HAS), and gelcasting (GC) from suspensions containing 50 vol% solids loading. For comparison purposes, the same ceramic compositions were also consolidated by die pressing of freeze-dried granules (FG). In the HAS process, 5 wt% of Al2O3 in the precursor mixture was replaced by equivalent amounts of AlN to promote the consolidation of the suspensions. Ceramics consolidated via GC exhibited higher green (three-point bend) strengths (∼17 MPa) than those consolidated by other techniques. Further, these ceramics also exhibited superior fracture toughness and flexural strength properties after sintering for 1 h at 1600°C in comparison with those consolidated by other techniques, including conventional die pressing (FG).  相似文献   

20.
The orientation and grain boundary microstructure of alumina in reactive metal penetration Al/Al2O3 composites are studied using orientation imaging microscopy and the results are compared with those of sintered polycrystalline Al2O3. The interconnected Al2O3 in the composite material is separated by Σ3 boundaries (twins) with a 60° rotation around the [0001] direction. A high frequency (∼100%) of Σ3 coincidence boundaries in composite alumina is remarkable since only ∼12% of boundaries in a sintered polycrystalline Al2O3 are of special nature. The coincidence boundaries in the in situ alumina grow in a coherent and faceted manner.  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