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 tribological properties of Ti2SC were investigated at ambient temperatures and 550°C against Ni-based superalloys Inconel 718 (Inc718) and alumina (Al2O3) counterparts. The tests were performed using a tab-on-disk method at 1 m/s and 3N (≈0.08 MPa). At room temperature, against the superalloy, the coefficient of friction, μ, was ∼0.6, and at ∼8 × 10−4 mm3·(N·m)−1 the specific wear rate (SWRs), was high. However, against Al2O3, at ∼5 × 10−5 mm3·(N·m)−1 and ∼0.3, the SWRs and μ were significantly lower, which was presumably related to more intensive tribo-oxidation at the contact points. At 550°C, the Ti2SC/Inc718 and Al2O3 tribocouples demonstrated comparable μ's of ∼0.35–0.5 and SWRs of ∼7–8 × 10−5 mm3·(N·m)−1. At 550°C, all tribosurfaces were covered by X-ray amorphous oxide tribofilms. At present, Ti2SC is the only member of a family of the layered ternary carbides and nitrides (MAX phases) that can be used as a tribo-partner against Al2O3 in the wide temperature range from ambient to 550°C.  相似文献   

2.
The electrical conductivity and ion/electron transference numbers in Al3O3 were determined in a sample configuration designed to eliminate influences of surface and gas-phase conduction on the bulk behavior. With decreasing O2 partial pressure over single-crystal Al2O3 at 1000° to 1650°C, the conductivity decreased, then remained constant, and finally increased when strongly reducing atmospheres were attained. The intermediate flat region became dominant at the lower temperatures. The emf measurements showed predominantly ionic conduction in the flat region; the electronic conduction state is exhibited in the branches of both ends. In pure O2 (1 atm) the conductivity above 1400°C was σ≃3×103 exp (–80 kcal/ RT ) Ω−1 cm−1, which corresponds to electronic conductivity. Below 1400°C, the activation energy was <57 kcal, corresponding to an extrinsic ionic condition. Polycrystalline samples of both undoped hot-pressed Al2O3 and MgO-doped Al2O3 showed significantly higher conductivity because of additional electronic conduction in the grain boundaries. The gas-phase conduction above 1200°C increased drastically with decreasing O2 partial pressure (below 10−10 atm).  相似文献   

3.
Sinterability of undoped, MgO-doped, and TiO2-doped Al2O3 has been examined by applying reported sintering equations. The order of sinterability was MgO-doped ∼ undoped≪ TiO2-doped Al2O3 in the initial and intermediate stages of sintering, but a relative sintered density at 1600°C for 1 h occurred in the order undoped < TiO2-doped < MgO-doped AI2O3. The dispersion of thermal grooving angles increased in the order MgO-doped < undoped < TiO2-doped Al2O3, The change of sinterability by the dopants is explained in terms of mobility of mass transfer estimated from a densification rate in the initial- and intermediate-stage sintering and of dispersed driving forces of densification and grain growth qualitatively evaluated from the width of the dispersion of thermal grooving angles.  相似文献   

4.
Solubility of Magnesia in Polycrystalline Alumina at High Temperatures   总被引:1,自引:0,他引:1  
High-purity Al2O3 compacts were doped with 0–350 ppm (by weight) of MgO using a liquid immersion technique and equilibrated at temperatures between 1700° and 2000°C under hydrogen. The solubility limits of MgO in Al2O3 at temperatures of 1720° and 1880°C were very low, ∼75 and 175 ppm, respectively. Variation of MgO solubility with temperature could be represented by the equation, ln Mg/Al = 3.80–2.63 × 104/ T . The small MgO solubilities were understood by the high enthalpy (326 kJ/mol) of solution. The results of this study suggested that previous investigations on sintering and grain-growth mechanisms in MgO-doped Al2O3 were probably not done in single-phase Al2O3 solid solutions. However, the conclusions on sintering and grain-growth mechanisms in prior research work in MgO-doped A2O3 may be correct. The effects of SiO2 impurity and grain size on MgO solubility are discussed. Previous grain-growth experiments in MgO-doped Al2O3 are described that demonstrate the clearest evidence for grain-boundary mobility controlled by a solid-solution mechanism.  相似文献   

5.
The rate of formation of NiAl2O4 by reaction between single crystals of NiO and Al2O3 can be described by k = 1.1 × 104 exp (−108,000 ± 5,000/ RT ) cm2/s. In NiO the behavior of D as a function of concentration supports the Lidiard theory of diffusion by impurity-vacancy pairs. A good fit of the theory to the experimental results was obtained by assuming that Al3+ ions diffuse as [AlNi· VNi]'pairs. The diffusion coefficient of pairs, Dp , obeys the equation 6.6 × 10−2 exp (−54,000 ± 3,000/ RT ) cm2/s. The free energy of association for pairs was calculated to range from 6.5 kcal/mol at 1789°C to 9.0 kcal/mol at 1540°C. The interdiffusion coefficients in the spinel showed a constant small increase with increasing concentration of Al3+ dissolved in the spinel.  相似文献   

6.
The sintering temperature of multilayer ceramic substrates must decrease to 1000° or below to avoid melting the conductors (Pd-Ag, Au, or Cu) during sintering. In this study, SiO2, CaO, B2O3, and MgO were used as additives to Al2O3 to decrease the firing temperature by liquid-phase sintering. Compositions with 18.0 and 22.5 wt% B2O3 were sintered at around 1000° in an air atmosphere to yield dense ceramics with good properties: relative dielectric contant between 6 to 7 (1 MHz), tan δ≤× 3 × 10−4 (1 MHz), insulating resistivity > 1014ω cm, coefficient of thermal expansion ∼ 7.0 × 10−6/°, and thermal conductivity ∼ 4.1 W/(m · K).  相似文献   

7.
The local environment around copper in the Cu2O·Al2O3·4SiO3 and CuO·Al2O3·4SiO2 glasses was investigated using the extended X-ray absorption fine structure (EXAFS) technique. It has been found that each Cu(I) in the former glass is coordinated to two oxygens through covalent Cu(I)-O bonds. This is the primary reason for its low thermal expansion coefficient (∼10 × 10−7 K−1). In the latter glass, which was made by heating the former glass at 600°C in air, each Cu(II) is coordinated to four oxygens and the bonds are ionic in character, contributing to the increase in the thermal expansion coefficient to ∼30 × 10−7 K−1.  相似文献   

8.
Dihedral angles, Ψs, from surface thermal grooves were measured using a metal reference line technique for polycrystalline MgO, undoped Al2O3, and MgO-doped Al2O3. The values of Ψs span the following ranges: 89° to 116° for MgO at 1520 K, 76° to 166° for undoped Al2O3 at 1870 K, and 90° to 139° for MgO-doped Al2O3 at 1870 K. For all three systems, the median Ψs values are 105° to 113°, implying that the median γgbs is 1.1 to 1.2, in contrast to metals where γgbs ranges typically from 1/4 to 1/2. The widths in Ψs distributions were different for the three materials with the width increasing in the following order: MgO, MgO-doped Al2O3, undoped Al2O3. For all three materials, the grainboundary grooves and their corresponding Ψs were not symmetrical with respect to the surface normal. The asymmetry for MgO was due to the pinning of the grain boundaries by the surface thermal grooves. The range of inclination angles of the grain boundary to the surface was a function of Ψs, with the maximum inclination angles of ∼13°, in quantitative agreement with theory.  相似文献   

9.
Preparation and Characterization of Aluminum Borate   总被引:2,自引:0,他引:2  
Aluminum borate, 9Al2O3·2B2O3 or Al18B4O33, was synthesized by the reaction of stoichiometric amounts of α-Al2O3 and B2O3. The Al18B4O33 material was formed into a dense ceramic by pressureless sintering with CaO, MgO, or CaAl2B2O7 additives. The material was characterized by low bulk density, moderate coefficient of thermal expansion (3 × 10−6/°C to 5 × 10−6/°C), moderate strength (210 to 324 MPa), and low dielectric constant.  相似文献   

10.
The formation of ZnAl2O4 spinel in diffusion couples of Al2O3 and ZnO was investigated between 1000° and 1390°C in air and in air containing 4.8 vol% Cl2 by X-ray diffraction, electron probe microanalysis, and scanning electron microscopy. The rate of formation of a spinel layer obeyed a parabolic rate law and was accelerated remarkably by the presence of Cl2. The interdiffusion coefficient, , and the activation energy, E, were calculated to be 10−8 to 10−9 cm2/s and 123 kcal/mol (514 kJ/mol) in air and 10−7 cm2/s and 31 kcal/mol (130 kJ/mol) in air containing 4.8 vol% Cl2, respectively.  相似文献   

11.
In a given batch more than 30%–40% of polycrystalline, MgO-doped Al2O3 tubes were converted into single crystals of sapphire by abnormal grain growth (AGG) in the solid state at 1880°C. Most crystals grew 4–10-cm in length in tubes with wall thicknesses of 1/2 and 3/4 mm and outer diameters of 5 and 7 mm, respectively, and had their c -axes oriented ∼ 90° and 45° to the tube axis. Initiation of AGG was associated with low values of bulk MgO concentration near 50 ppm. The unconverted tubes did not develop centimeter-size crystals but instead exhibited millimeter-size grains. The different grain structures in converted and unconverted tubes may be related to nonuniform concentration of MgO in the extruded tubes. The growth front of the migrating crystal boundary was typically nonuniformly shaped, and the interface between the single crystal and the polycrystalline matrix was composed of many "curved" boundary segments indicative of classical AGG in a single-phase material. The average velocities of many migrating crystal boundaries were quite high and reached ∼1.5 cm/h. The average grain boundary mobility at 1880°C was calculated as 2 × 10−10 m3/(N·s), representing the highest value reported so far in Al2O3 and within a factor of 2.5 of the calculated intrinsic mobility. Under similar experimental conditions sapphire crystals did not grow when a codopant of CaO, La2O3, or ZrO2 was added in concentrations of several hundred ppm.  相似文献   

12.
Interdiffusion coefficients in single-crystal MgO were determined using an MgO-MgAl2O4 diffusion couple. For a concentration of 1 mol% Al2O3 in MgO, the interdiffusion coefficient can be expressed as D =2.0±0.2 exp (−76,000±3,000/ RT ) for the MgO-MgAl2O4 couple. This relation compares well with previous measurements in the MgO-Al2O3 system. The interdiffusion coefficients, which increased with the mol fraction of cation vacancies, were in the range of 10−8 to 10−10 cm2s−1 for the concentrations and temperatures studied. Diffusion was enhanced below 1640°C if powdered MgAl2O4 was used. Self-diffusion coefficients for Al3+ ions in MgO were calculated; Al3+ diffuses faster than Cr3+ in MgO.  相似文献   

13.
The deviation from stoichiometry, δ, in Cr2−δO3 was measured by a tensivolumetric method in the high pO2 range of ≊104 to 104 Pa at 1100°C. The value of δ, or chromium vacancy concentration, was≊9×10−5 mol/mol Cr2O3 in air for Cr2O3 with 99.999% purity. The chemical diffusion coefficient, DT, determined from equilibration data was ≊4.6× cm2·s−1 at 1100°C for pO2= 2.2 ×101 Pa. The self-diffusion coefficient of Cr ions was calculated from and δ and found to be≊1.6×10-17 cm2-s−1, in good agreement with recently measured values.  相似文献   

14.
M-doped zinc oxide (ZnO) (M=Al and/or Ni) thermoelectric materials were fully densified at a temperature lower than 1000°C using a spark plasma sintering technique and their microstructural evolution and thermoelectric characteristics were investigated. The addition of Al2O3 reduced the surface evaporation of pure ZnO and suppressed grain growth by the formation of a secondary phase. The addition of NiO promoted the formation of a solid solution with the ZnO crystal structure and caused severe grain growth. The co-addition of Al2O3 and NiO produced a homogeneous microstructure with a good grain boundary distribution. The microstructural characteristics induced by the co-addition of Al2O3 and NiO have a major role in increasing the electrical conductivity and decreasing the thermal conductivity, resulting from an increase in carrier concentration and the phonon scattering effect, respectively, and therefore improving the thermoelectric properties. The ZnO specimen, which was sintered at 1000°C with the co-addition of Al2O3 and NiO, exhibited a ZT value of 0.6 × 10−3 K−1, electrical conductivity of 1.7 × 10−4Ω−1·m−1, the thermal conductivity of 5.16 W·(m·K)−1, and Seebeck coefficient of −425.4 μV/K at 900°C. The ZT value obtained respects the 30% increase compared with the previously reported value, 0.4 × 10−3 K−1, in the literature.  相似文献   

15.
The initial stages of growth of ordered layers of Al2O3 on NiAl(001) single-crystal surfaces at 1025 K and 10−7 mbar (10−5 Pa) in O2 have been studied using scanning tunneling microscopy (STM) and scanning tunneling spectroscopy (STS). The STM results evidence the formation of elongated strips (26 Å wide and 11 Å high) of Al2O3 oriented along the [100] and [010] directions of the substrate. With longer oxidation, the substrate is increasingly covered by rectangular and striped islands resulting from the vicinal and parallel growth of the strips. On the ultrathin oxide film formed after 500 L (1 L = 10−6 torr·s (∼1.33 × 10−4 Pa·s)) of exposure, STM atomic resolution images have been obtained for the first time. They evidence the [001] orientation of the oxygen sublattice in Bain epitaxy on the substrate. The observation of one-dimensional atomic trenches together with the strips observed on the nanometer scale is consistent with the growth of θ-Al2O3. The STS local measurements evidence the insulating behavior of the oxide layer formed with a gap value ranging from 7 and 8 eV for amorphous and ordered domains, respectively.  相似文献   

16.
Low-Temperature Sintering of Lead-Based Piezoelectric Ceramics   总被引:3,自引:0,他引:3  
The low-temperature sintering of lead-based piezoelectric ceramics has been studied. The sintering temperature of lead zirconate titanate (PZT) ceramics could be reduced from ∼ 1250° to ∼960°C by the addition of a small amount of the lower-melting frit, B2O3–Bi2O3—CdO. It exhibited the following dielectric and piezoelectric properties: Kp= 0.52 to 0.58, Qm= 1000, εT330= 800 to 1000, tan δ= 50 × 10−4, ρ= 7.56 to 7.64 g/cm3. Ceramics with the aid of suitable dopants (CdO, SiO2, and excess PbO) in the Pb-(Ni1/3Nb2/3)O3—PZT family could be sintered at 860° to 900°C. For these materials, Kp= 0.56 to 0.61, Qm= 1000, εT330= 1500 to 2000, tan δ≤ 50 × 10−4, ρ= 7.80 to 8.03 g/cm3. The microstructure, sintering mechanism, and the effects of various impure additions have been analyzed by means of scanning electron microscopy, scanning transmission electron microscopy, electron probe microanalysis, and X-ray photoelectron spectroscopy.  相似文献   

17.
An investigation of the properties of high-purity (>99 wt%) tantalum tungstates (Ta22W4O67, Ta, WO8, and Ta16W18O94) included determination of density (bulk and theoretical), refined lattice constants, maximum use temperatures, micro-hardness, heat capacity, thermal expansion (contraction) and diffusivity, calculated thermal conductivity, and electrical resistivity. Usable to ∼ 1700 K in air or inert atmospheres, these tantalum tungstates have theoretical densities of 7.3 to 8.5 g/cm3, are relatively soft (120 to 655 kg/mm2 hardnesses), and are electrical insulators (6× 103 to 2× 108Ω.cm resistivities). The distinguishing properties of the materials are their thermal expansion (average CTE values from + 0.6×10−8/K to −5.1× 10−6/K at 293 to 1273 K), thermal expansion hysteresis with minimal observable microcracking, and thermal diffusivity  相似文献   

18.
Steady-state creep was studied in hot-forged polycrystalline Al2O3 (3 to 42 μm) of nearly theoretical density doped with≤1 cation % of Fe, Ti, or Cr. Tests were conducted at stresses between 10 and 550 kg/cm2 at 1375° to 1525°C under O2 partial pressures of 0.88 to 10−10 atm. Except in the 10-μm Fe-doped material tested at very small stresses, slightly nonviscous creep behavior was generally observed. The effects of P o2 on the creep rate indicated that increased concentration of a divalent (Fe2+) or quadrivalent (Ti4+) impurity in solid solution enhances the creep rate of polycrystalline Al2O3. The activation energies for the creep of Fe- and Ti-doped Al2O3 samples (148 and 145 kcal/mol, respectively) were significantly higher than that for Cr-doped material (114 kcal/mol). Taking into account the effects of Po2, temperature, and grain size, it was concluded that the steady-state creep of transition-metal-doped Al2O3 is controlled by cation lattice diffusion.  相似文献   

19.
The sintering of a composite of MgO–B2O3–Al2O3 glass and Al2O3 filler is terminated due to the crystallization of Al4B2O9 in the glass. The densification of a composite of MgO–B2O3–Al2O3 glass and Al2O3 filler using pressureless sintering was accomplished by lowering the sintering temperature of the composite. The sintering temperature was lowered by the addition of small amounts of alkali metal oxides to the MgO–B2O3–Al2O3 glass system. The resultant composite has a four-point bending strength of 280 MPa, a coefficient of thermal expansion (RT—200°C) of 4.4 × 10−6 K−1, a dielectric constant of 6.0 at 1 MHz, porosity of approximately 1%, and moisture resistance.  相似文献   

20.
A Cr–Al–C composite was successfully synthesized by a hot-pressing method using Cr, Al, and graphite as starting materials. X-ray diffraction, scanning electron microscopy, and transmission electron microscopy analyses revealed that the composite contained Cr2AlC, AlCr2, Al8Cr5, and Cr7C3. The orientation relationships and atomic-scale interfacial microstructures among Cr2AlC, AlCr2, and Al8Cr5 are presented. This composite displays both excellent high-temperature oxidation resistance in air and hot-corrosion resistance against molten Na2SO4 salt. The parabolic rate constants for the oxidation in air at 1000°, 1100°, and 1200°C are 3.0 × 10−12, 6.2 × 10−11, and 6.2 × 10−10 kg2 (m4·s)−1, respectively, while the linear weight gain rates for the hot corrosion of Na2SO4-coated samples at 900° and 1000°C are, respectively, 1.2 × 10−3 and 4.4 × 10−3 mg (cm2·h)−1. The mechanism of the excellent high-temperature corrosion resistance can be attributed to the formation of a protectively alumina-rich scale.  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