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 265 毫秒
1.
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   

2.
Microstructural evolution adn densification kinetics of Al2O3-SiC powder composites were studied using two different SiC powders. Examination of the microstructural evolution of Al2O3-fine SiC powder composites showed three well-defined stages of densification: the first was characterized by constant pore size and no grain growth; the second involved rapid pore coarsening and grain growth; the third was characterized by pore shrinkage and slow grain growth. Studies of the densification kinetics of Al2O3-coarse SiC powder composites exhibited two stages of densification: in the first stage there were no significant differences in densification rate between pure Al2O3 compacts and composites; in the second stage, however, differences in densification behavior between pure Al2O3 compacts and composites became pronounced.  相似文献   

3.
Stable Al2O3–SiC–YAG hybrid composites were successfully fabricated by reaction of Al2O3 and Y2O3 and incorporation of SiC. The hot-pressed bodies consisted of uniformly dispersed grains of microsized YAG particulates and nanosized SiC particulates in an Al2O3 matrix. Although the grain size of monolithic A12O3 increases markedly with increased temperature, the grain size of the Al2O3–SiC–YAG hybrid composites was effectively restrained due to grain-boundary pinning by the particulates.  相似文献   

4.
The composite sol—gel (CSG) technology has been utilized to process SiC—Al2O3 ceramic/ceramic particulate reinforced composites with a high content of SiC (up to 50 vol%). Alumina sol, resulting from hydrolysis of aluminum isopropoxide, has been utilized as a dispersant and sintering additive. Microstructures of the composites (investigated using TEM) show the sol-originating phase present at grain boundaries, in particular at triple junctions, irrespective of the type of grain (i.e., SiC or Al2O3). It is hypothesized that the alumina film originating from the alumina sol reacts with SiO2 film on the surface of SiC grains to form mullite or alumina-rich mullite-glass mixed phase. Effectively, SiC particles interconnect through this phase, facilitating formation of a dense body even at very high SiC content. Comparative sinterability studies were performed on similar SiC—Al2O3 compositions free of alumina sol. It appears that in these systems the large fraction of directly contacting SiC—SiC grains prevents full densification of the composite. The microhardness of SiC—Al2O3 sol—gel composites has been measured as a function of the content of SiC and sintering temperature. The highest microhardness of 22.9 GPa has been obtained for the composition 50 vol% SiC—50 vol% Al2O3, sintered at 1850°C.  相似文献   

5.
The mechanical properties of various 2D ceramic matrix fiber composites were characterized by tension testing, using the gripping and alignment techniques development in this work. The woven fabric composites used for the test had the basic combinations of Al2O3 Fabric/Al2O3, SiC fabric/SiC, and SiC minofilament uniweave fabric/SiC. Tension testing was performed with strain gauge and acoustic emission instrumentation to identify the first-matrix cracking stress and assure a valid alignment. The peak tensile stresses of these laminate composites were about one-third of the flexural strengts. The SiC monofilament uniweave fabric (14 vol%)/SiC composites showed a relatively high peak stress of 370 MPa in tension testing.  相似文献   

6.
Composites of silicon carbide (SiC) with up to 30 vol% of dysprosia (Dy2O3) were fabricated by hot pressing and hot isostatic pressing. The effects of Dy2O3 dispersions on the microstructure and on selected mechanical properties of the composites were investigated. When 10-15 vol% of Dy2O3 was dispersed in the SiC matrix, the fracture toughness increased by ∼40%, whereas the flexural strength was comparable to that of unreinforced SiC. The increased fracture toughness was due to crack deflection, in conjunction with crack-interface grain bridging, and was not related to a phase transformation of Dy2O3 in the matrix.  相似文献   

7.
Al2O3-5 vol% SiC particle composites doped with 800 ppm rare-earth impurities (Y3+, Nd3+, and La3+) were fabricated by hot-pressing at a temperature of 1550°C. Doping of rare-earth impurities in Al2O3-SiC composites led to a fracture- mode change from transgranular in dopant-free composites to intergranular in rare-earth-doped composites. The fracture mode change obviously increased the crack deflection so that the fracture toughness of rare-earth-doped composites was higher than that of the composites without dopants, especially for the Nd3+- and La3+-doped composites. It was found that the fracture-mode change originated from a weak grain-boundary bonding caused by co-segregation of the rare-earth dopants and Si4+ ions dissolved from the SiC particle surfaces.  相似文献   

8.
Flexural creep studies were conducted in a commercially available alumina matrix composite reinforced with SiC particulates (SiCp) and aluminum metal at temperatures from 1200° to 1300°C under selected stress levels in air. The alumina composite (5 to 10 μm alumina grain size) containing 48 vol% SiC particulates and 13 vol% aluminum alloy was fabricated via a directed metal oxidation process (DIMOX(tm))† and had an external 15 μm oxide coating. Creep results indicated that the DIMOX Al2O3–SiCp composite exhibited creep rates that were comparable to alumina composites reinforced with 10 vol% (8 (μm grain size) and 50 vol% (1.5 μm grain size) SiC whiskers under the employed test conditions. The DIMOX Al2O3–SiCp composite exhibited a stress exponent of 2 at 1200°C and a higher exponent value (2.6) at ≥ 1260°C, which is associated with the enhanced creep cavitation. The creep mechanism in the DIMOX alumina composite was attributed to grain boundary sliding accommodated by diffusional processes. Creep damage observed in the DIMOX Al2O3-SiCp composite resulted from the cavitation at alumina two-grain facets and multiple-grain junctions where aluminum alloy was present.  相似文献   

9.
The effect of SiC concentration on the liquid and solid oxide phases formed during oxidation of ZrB2–SiC composites is investigated. Oxide-scale features called convection cells are formed from liquid and solid oxide reaction products upon oxidation of the ZrB2–SiC composites. These convection cells form in the outermost borosilicate oxide film of the oxide scale formed on the ZrB2–SiC during oxidation at high temperatures (≥1500°C). In this study, three ZrB2–SiC composites with different amounts of SiC were tested at 1550°C for various durations of time to study the effect of the SiC concentration particularly on the formation of the convection cell features. A calculated ternary phase diagram of a ZrO2–SiO2–B2O3 (BSZ) system was used for interpretation of the results. The convection cells formed during oxidation were fewer and less uniformly distributed for composites with a higher SiC concentration. This is because the convection cells are formed from ZrO2 precipitates from a BSZ oxide liquid that forms upon oxidation of the composite at 1550°C. Higher SiC-containing composites will have less dissolved ZrO2 because they have less B2O3, which results in a smaller amount of precipitated ZrO2 and consequently fewer convection cells.  相似文献   

10.
Al2O3 and SiC composite materials have been produced from mixtures of aluminosilicates (both natural minerals and synthetic) and carbon as precursor materials. These composites are produced by heating a mixture of kaolinite (or synthetic aluminosilicates) and carbon in stoichiometric proportion above 1550°C, so that only Al2O3 and SiC remain as the major phases. A similar process has also been used for synthesizing other composite powders having mixtures of Al2O3, SiC, TiC, and ZrO2 in different proportions (all compounds together or selective mixtures of some of them), as desired. The microstructure of hot-pressed dense compacts, produced from these powders, revealed that the SiC phase is distributed very homogeneously, even occasionally within Al2O3 grains on a nanosize scale. The homogeneous distribution of SiC particles within the system produced high fracture toughness of the hot-pressed material (KIC∼ 7.0 MPa · m1/2) and having Vicker's hardness values greater than 2000 kgf/mm2.  相似文献   

11.
One possible approach to improving the densification of powder composites containing a major crystalline phase which densifies (e.g., Al2O3) and a difficult-to-sinter phase (e.g., SiC) is to accommodate the matrix volume shrinkage with a "disappearing" polymer coating. A polymer coating prevents contact between the nonsinterable particles and the surrounding matrix. The coating can be burned off before sintering, allowing the matrix phase to "shrink-fit" around the nonsinterable particles during sintering. The effects of a polymer coating on the densification of a two-phase particle system were tested using SiC powder dispersed in an Al2O3 matrix. The composites processed with a polymer coating showed more densification during equivalent firing cycles than did those processed without a polymer coating. Densification during sintering was approximately proportional to the amount of polymer adsorbed on SiC, suggesting that the Al2O3 matrix did shrink-fit into the gaps between the SiC particles and the surrounding Al2O3 matrix. Differences in the pore-size distributions of polymercoated green compacts and uncoated compacts indicated a perturbation of the green microstructure by the gaps. The estimated average thickness of the gap is approximately 20 nm, ∼8% of the average radius of the SiC powder used in this study.  相似文献   

12.
Mullite composites toughened with ZrO2 (with or without a MgO or Y2O3 stabilizer) and/or SiC whiskers (SiC( w )) were fabricated by hot-pressing powders prepared from Al, Si, Zr, and Mg(Y) alkoxide precursors by a sol–gel process. Micro-structures were studied by using XRD. SEM, and analytical STEM. Pure mullite samples contained prismatic, preferentially oriented mullite grains. However, the addition of ZrO2, as well as the hot-pressing temperature, affected the morphology and grain size in the composites; a fine, uniform, equiaxed microstructure was obtained. The effect of SiC( W ) was less pronounced than that of ZrO2. Glassy phases were present in mullite and mullite/SiC( W ) composites, but were rarely observed in Al2O3-rich or ZrO2-containing samples. The formation of zircon due to the reaction between ZrO2 and SiO2 and the considerable solid solution of SiO2 in ZrO2 prevented the formation of the glassy phase, whereas the reaction between Al2O3 and MgO in MgO-containing samples formed a spinel phase and also deprived the ZrO2 phase of the stabilizer. Intergranular ZrO2 particles were either monoclinic or tetragonal, depending on size and stabilizer content; small intragranular ZrO2 inclusions were usually tetragonal in structure.  相似文献   

13.
Oxidation studies were conducted on Al2O3-SiC and mullite-SiC composites at 1375° to 1575°C in O2 and in Ar-1% O2. The composites were prepared by hot-pressing mixtures of Al2O3 or mullite and SiC powders. The reaction products contained alumina, mullite, an aluminosilicate liquid, and gas bubbles. The parabolic rate constants were about 3 orders of magnitude higher than those expected for the oxidation of SiC. Higher rates are caused by higher oxygen permeabilities through the reaction products than through pure silica. Our results suggest that oxygen permeabilities are comparable in the three condensed phases observed in the reaction products.  相似文献   

14.
During high-temperature oxidation in air, SiC-whisker-reinforced Al2O3—ZrO2 composites degrade by the formation of a whisker-depleted mullite-zirconia scale. The reaction kinetics have been studied as a function of time and temperature for composites with whiskers preoxidized for different times. The evolution of the microstructure has been investigated by optical, scanning and transmission electron microscopy. Possible reaction mechanisms have been discussed. A model compatible with our observations on Al2O3—ZrO2—SiC and the results reported in the literature for Al2O3—SiC whisker composites is proposed: The oxidation occurs at an internal reaction front. Oxygen diffuses along dislocations and grain boundaries through the mullite scale to react at this front with silicon carbide, thereby forming amorphous silica and graphite. Silica penetrates grain boundaries and further reacts with alumina and zirconia to form mullite and zircon, while the second reaction product, graphite, is oxidized into carbon monoxide when the reaction front moves deeper into the sample.  相似文献   

15.
Model laminated composites were fabricated with porous-Al2O3 interfaces between SiC bars. The porous Al2O3 was deposited using an aerosol spray deposition technique, and the sandwich specimen was fabricated by hot pressing. Residual thermal stresses were present in the interface because of the difference in the coefficients of thermal expansion of SiC and Al2O3. Crack deflection was observed with measured interfacial fracture resistances that were considerably higher than the deflection threshold predicted by the He–Hutchinson criterion. Examination of the fracture surface revealed a tortuous crack path and significant crack–flaw interaction.  相似文献   

16.
An aluminum/MgAl2O4 in situ metal matrix composite has been synthesized using silica gel containing ∼98% SiO2 in an Al–5Mg alloy. The thermodynamics and kinetics of MgAl2O4 formation have been discussed in detail. A transition phase of composition between MgO and MgAl2O4 has been detected in the SEM-EDS analysis of the particles extracted from the composite by a 25% NaOH solution. This confirms the gradual transformation of MgO to MgAl2O4 by the reaction 3SiO2( s )+2MgO( s )+4Al( l )→2MgAl2O4( s )+3Si( l ). The stoichiometry, n , of MgAl2O4 has been found to sustain close to 1 and the crystallite growth of MgAl2O4 has been stopped at D ∼30 nm in the composites held at 750°C up to 10 h.  相似文献   

17.
Based on the RBAO technology, low-shrinkage mullite/SiC/ Al2O3/ZrO2 composites were fabricated. A powder mixture of 40 vol% Al, 30 vol% A12O3 and 30 vol% SiC was attrition milled in acetone with TZP balls which introduced a substantial ZrO2 wear debris into the mixture. The precursor powder was isopressed at 300–900 MPa and heattreated in air by two different cycles resulting in various phase ratios in the final products. During heating, Al oxidizes to Al2O3 completely, while SiC oxidizes to SiO2 only on its surface. Fast densification (at >1300°C) and mullite formation (at 1400°C) prevent further oxidation of the SiC particles. Because of the volume expansion associated with the oxidation of Al (28%), SiC (108%), and the mullitization (4.2%), sintering shrinkage is effectively compensated. The reaction-bonded composites exhibit low linear shrinkages and high strengths: shrinkages of 7.2%, 4.8%, and 3%, and strengths of 610, 580, and 490 MPa, corresponding to compaction pressure of 300, 600, and 900 MPa, respectively, were achieved in samples containing 49–55 vol% mullite. HIPing improved significantly the mechanical properties: a fracture strength of 490 MPa and a toughness of 4.1 MPa.m1/2 increased to 890 MPa and 6 MPa.m1/2, respectively.  相似文献   

18.
Al2O3–SiC particulate composites were fabricated by hot-pressing mixtures of 5–30 vol% SiC with either α-Al2O3, γ-Al2O3, or boehmite (γ-AlOOH) to determine whether grain growth or the α-alumina phase transformation could be used to fabricate intragranular particulate composites. Samples starting with α-alumina resulted in primarily intergranular SiC of 0.3 μ and an alumina grain size of 1.5–4.1 μm. Heat treatments resulted in SiC coarsening but no entrapment of SiC by grain boundary breakaway. The α-alumina transformation in the samples starting with γ-alumina resulted in the entrapment of ∼48% of the 5 vol% of SiC added whereas 79% of the SiC was entrapped in the α-alumina grains in samples starting with boehmite. Only SiC particles ≤0.2 SmUm were entrapped in the α-alumina grains during the phase transformation. With increasing SiC content, the relative volume of intragranular SiC decreased, but the amount of intragranular SiC was constant and independent of the amount of SiC added before transformation. The formation of intragranular composites from γ-alumina and boehmite samples was explained with a model that attributes particle entrapment to the vermicular growth of α-alumina into the transition alumina matrix during the α-alumina phase transformation. Seeding the boehmite-based samples did not affect the concentration of entrapped SiC, but did lower the hot-pressing densification temperature by as much as 150°C.  相似文献   

19.
Advanced Alumina Composites Reinforced with Titanium-Based Alloys   总被引:1,自引:0,他引:1  
New (inter)metallic-ceramic composites for high-temperature structural and functional applications are prepared via high-energy ball milling. During compaction by pressureless sintering, dense Al2O3/Ti-based alloy composites are formed that consist of inter-connected networks of the ceramic and the (inter)metallic phases. Ti-Al-V/Al2O3 and Ti-Al-Nb/Al2O3 composites show enhanced damage tolerance over monolithic Al2O3, i.e ., fracture toughnesses up to 5.6 MPa·m0.5 and bending strengths up to 527 MPa. The resistance against abrasive wear is almost doubled with respect to monolithic Al2O3 ceramic. Electrical resistivity scales with the ceramic volume fraction and ranges between 0.3 mΩ·cm and 55.1 mΩ·cm, with only a weak temperature dependence ≤700°C.  相似文献   

20.
SiC( w )/Al2O3 composites were made from an AlCl3-H2-CO2 mixture by a thermal-gradient chemical vapor infiltration (CVI) method. Al2O3 was deposited from the reaction of AlCl3 and H2O, which was produced from the oxidation of H2 by CO2. The densification rate was measured at various reactant compositions and total pressures. When the reaction rate or total pressure increased, the rate-controling step shifted from H2O production to AlCl3 diffusion, which led to premature pore closing. To obtain dense composites in a short infiltration time, the diffusion rate of AlCl3 had to be increased by decreasing the total pressure.  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