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 15 毫秒
1.
Self-diffusion coefficients for the oxygen ion in single-crystal Mn-Zn ferrite were determined by the gas-solid isotope exchange technique. The oxygen volume diffusion coefficients can be expressed as D =6.70 × 10−4 exp (-330 (kJ /mol) /RT)m2/s (>1350°C), D=3.94 × 10−10 exp (−137 (kJ/mol)/RT)m2/s (1100° to 1350°C), and D=7.82 × 104 exp (−507 (kJ/mol)/RT)m2/s (<1100°C).  相似文献   

2.
The oxidation behavior of SiCN–ZrO2 fibers and SiCN at 1350°C are compared. The as-measured parabolic rate constants for the two materials are nearly the same (15–20 × 10−18 m2/s). However, after implementing a correction for the difference in the compositions, the rate constant is 13.2 × 10−18 m2/s for the fiber, and 29.4 × 10−18 m2/s for SiCN. The lower oxidation rate of the fiber is ascribed to the lower carbon content in the fiber material.  相似文献   

3.
Strontium titanate (SrTiO3) is known as a good high-temperature resistive oxygen sensor material; its response time depends on oxygen bulk diffusion and surface exchange processes. In the present work, 18O diffusion has been investigated in lanthanum-doped SrTiO3, single crystals in the temperature range 700° to 900°C by secondary ion mass spectrometry (SIMS). Oxygen tracer diffusivities between 2 × 10−15 and 1 × 10−13 cm2/s have been calculated from the SIMS results. Low surface enrichment of 18O compared to the 18O concentration in the gas atmosphere gives clear evidence for a surface exchange reaction.  相似文献   

4.
Anion self-diffusion coefficients normal to (1102) were obtained for single-crystal Al2O3 in a 1.3 × 10 3 N/m2 (10−5 torr) vacuum at 1585° to 1840°C. Tracer was supplied from an initial 650 to 1300 A Al218O3 layer produced by the oxidation of vapor-deposited Al metal films in an 18O2 atmosphere at 520°C. Concentration gradients extended over depths of 3000 to 5000 A and were measured by mass spectrometry of material sputtered from the samples with a beam of Ar+ ions. Crystals which had not been preannealed to remove surface damage displayed enhanced diffusion. Diffusion coefficients from preannealed crystals may be described by D0 =6.4×105cm2/s, with an activation energy of 188 ± 7 kcal/mol. The diffusion is interpreted as an extrinsic vacancy mechanism.  相似文献   

5.
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   

6.
Compression creep tests were performed on fully dense specimens of UC1.01, UC1.05, UC1.01.+ 4 wt% W, and U0.9Zr0.1C1.01+ 4 wt% W. Steady-state creep rates were measured from 1400° to 1800°C in a vacuum of 1.33 × 10-3 N/m2 (1 × 10-5 torr) at stresses of 4.55 to 69.0 MN/m2 (660 to 10,000 psi). The data for UC1.01 could best be fit by an expression of the form ɛ= 1773σ6.024 exp (106.5/RT) , where σ is the steady-state creep rate (h-l), σ is the applied stress (MN/m2), and the creep activation energy is given in kcal/mol. The stress dependence for creep of UC1.05 decreased with decreasing temperature because of second-phase precipitation; therefore, a unique creep activation energy could not be established for this U/C ratio. At all temperatures, the creep strength of UC1.05 exceeded that of UC1.01. For example, at 1700 ° C steady-state creep rates for UC1.05 are ∼1/4 those for UC1.01, but at 1400°C the creep rates are ∼ 3 orders of magnitude less. At 1700°C, creep rates for UC alloys are ∼4 orders of magnitude lower than those for unalloyed UC1.01.  相似文献   

7.
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   

8.
A diffusion couple of an oxidized molybdenum disk and a glass cylinder was used to measure the solubiiity and effective binary diffusion coefficient of MoO2 in a non-alkali aluminosilicate glass. At 1400°C, the solubility limit was 8.4 mol%; the value of the diffusion coefficient (4.1 × 10−16 m2/s) was significantly lower than that estimated from the Stokes-Einstein relation.  相似文献   

9.
Creep of 9.4-mol%-Y2O3-stabilized cubic ZrO2 has been studied between 1300° and 1550°C. Conventional power-law creep (stress exponent n ∼ 4.5) is found at the higher temperatures, with an activation energy (∼6 eV) corresponding to cation diffusion. Transition to a different creep mechanism occurs at the lower temperatures, as indicated by higher values of the stress exponent ( n ∼ 7) and an activation energy (∼7.5 eV) higher than that for cation self-diffusion. The lower-temperature behavior is caused by a competition between cross-slip-controlled and recovery-controlled creep. Consideration of all the creep and diffusion data now available suggests that the rate-controlling high-temperature mass transport in Y2O3-stabilized ZrO2 can be described by D = 10−3 exp(-5.0 eV/ kT ) m2·s−1.  相似文献   

10.
The mechanisms of the sintering of ZnS were determined by measurement of the rate of growth of the necks between polycrystalline spheres. In vacuum (10−6 mm Hg) at temperatures higher than 600° C the mechanism of sintering is that of volume diffusion with coefficient Dv, = 1.06 × 10−2 exp (-26,400/RT); below 600°C surface diffusion predominates, with coefficient D3, = 9.14 × 10-3 exp (-5700/RT). In Zn vapor (10−2 mm Hg) between 550° and 650°C the predominating mechanism of sintering is surf ace diffusion, D3, = 7.06 × 10−2 exp (-8600/RT). It has been found that in an argon atmosphere the diffusion coefficient is much lower.  相似文献   

11.
The established analysis for the study of oxidation using powder specimens is based on the assumption of monosized particles. The experiments, however, are conducted on powders with a distributed particle size. Here we present a statistical approach for the calculation of the rate constant for oxidation. The results of the analysis are applied to new data on oxidation studies of dense powders of silicon carbonitride amorphous ceramics. The monosized model requires a wide range of values for the rate constant to fit the short term and the long-term data, leading to considerable ambiguity in the estimate of the parabolic rate constant, k p, for oxidation. In contrast the statistical model fits over the entire range of data, yielding a much more reliable value for k p. For example, the monosized approach gave a value in the range 19.7 × 10−18 < k p < 2.7 × 10−18 m2/s. In contrast, the statistical model yields a specific value of 4.5 × 10−18 m2/s.  相似文献   

12.
Electrochemical liquid deposition (ELD) has been used to grow fully dense films of ceria on impervious gadolinia-doped ceria substrates of variable thickness. The process is similar to electrochemical vapor deposition (EVD) except that a liquid is used instead of a vapor to provide precursor material to the growing film. ELD film growth data have been analyzed in a manner similar to that for EVD demonstrated previously. Experimental results verify that the pertinent rate equation is obeyed phenomenologically. Rate constants for the film (CeO2) and the substrate (l-l-mol%-Gd2O3-doped CeO2) at 1200°C have been determined as Kf = 33.1 × 10−12 m2/s and Ks = 51.6 × 10−12 m2/s, respectively.  相似文献   

13.
Optical and electron microscopies are used to analyze the mechanism and kinetics of internal reduction of an Fe2+-doped magnesium aluminosilicate melt. Melt samples are heated to temperatures in the range of 1300°–1400°C under a flowing gas mixture of CO/CO2, which corresponds to a p O2 range of 1 × 10−13–4 × 10−13 atm. The melt experiences an internal reaction in which a dispersion of nanometer-scale iron-metal precipitates forms at an internal interface. The metal precipitates show no signs of coarsening within the samples; however, the crystals at the surface (which formed in the initial part of the reaction) do grow via vapor phase transport. The overall reaction is characterized by parabolic kinetics, which is indicative of chemical diffusion being the rate-limiting step. The diffusion of network-modifier divalent cations—particularly Mg2+ cations—is demonstrated to be the rate-limiting factor, and its diffusion coefficient is calculated to be ∼1 × 10−6 cm2/s within the temperature range of the experiments.  相似文献   

14.
Grain boundary grooving experiments were conducted with Σ5 (210) twist boundaries in Y3Al5O12 (YAG) with the goal of extracting information on diffusion in YAG. Planar boundaries oriented 90° to the surface were annealed in air at various times and temperatures. Atomic force microscopy was used to characterize the subsequent grooves. The Mullins approach leads to the following expression for the diffusion coefficient: D (m2/s) = 3.9 × 10−10 exp[−330 ± 75 (kJ/mol)/ RT ]. The relatively low activation energy agrees well with earlier oxygen tracer diffusion measurements on YAG, suggesting that oxygen is the limiting diffusing species in boundary grooving of YAG.  相似文献   

15.
Pore Drag and Pore-Boundary Separation in Alumina   总被引:2,自引:0,他引:2  
Microdesigned interfacial pore structures were used to study pore drag and pore-boundary separation in Al2O3. This approach allows the creation of pore arrays containing pores of controlled size and spacing at well-defined singlecrystal seed/polycrystalline matrix interfaces, and enables experimental determination of the peak pore velocity. From the peak pore velocity, values of the surface diffusion coefficient pertinent to sintering can be extracted. At 1600°C, the surface diffusion coefficient is ∼1 × 10−7 cm2/s for undoped Al2O3 and ∼4 × 10−7 cm2/s for MgO-doped Al2O3. The values appear to be insensitive to the seed orientation for the two seed orientations studied. The results suggest a strong influence of pore spacing on the separation condition in undoped Al2O3, and a diminished influence in MgO-doped Al2O3. Quantitative agreement between theoretically predicted and experimentally observed separation/attachment conditions was obtained.  相似文献   

16.
Oxygen self-diffusion coefficients for a single-crystal MgO doped with 400 ppm Li were determined at 848° to 1300°C by isotope exchange technique using 18O as the tracer. The diffusion coefficient increased with increasing temperature in the lower temperature range, as described by D = 7.8 × 10−4 exp[−279 (kJ/mol)/RT] cm2/s, which was interpreted to be an extrinsic diffusion due to the oxygen vacancies introduced by substitutional Li ions. The oxygen diffusivity tended to decrease above 1050°C, presumably becauuse of evaporation of Li2O.  相似文献   

17.
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   

18.
19.
The oxidation kinetics were determined for single-crystal SrTiO3 by measuring the time and temperature dependence of the weight gain of reduced crystals. The oxidation can be described as a diffusion-controlled process. The calculated diffusion coefficients between 850° and 1460°C are represented by D = 0.33 exp (-22.5 ± 5.0 kcal/ RT ) cm2/sec. Directly measured oxygen ion diffusion coefficients in the same temperature interval reported earlier are interpreted as being extrinsic and can be represented by D = 5.2 × 10−6 exp (-26.1 ± 5.0 kcal/ RT ) cm2/sec, where the activation energy is for mobility only. Assuming that the calculated diffusion coefficients are for vacancy diffusion and the two activation energies are equivalent within experimental error, a vacancy concentration (fraction of vacant lattice sites), [O□], fixed by impurities in the fully oxidized crystal is calculated to be 1.6 × 10−5 by virtue of the relation between the oxygen self-diffusion coefficient, D02-, and the oxygen vacancy diffusion coefficient, Do□ ; D o2-= [O□] D o□ where the oxygen ion concentration [O2-] is taken as unity.  相似文献   

20.
Crystals of β-Ca2SiO4 (space group P 121/ n 1) were examined by high-temperature powder X-ray diffractometry to determine the change in unit-cell dimensions with temperature up to 645°C. The temperature dependence of the principal expansion coefficients (αi) found from the matrix algebra analysis was as follows: α1= 20.492 × 10−6+ 16.490 × 10−9 ( T - 25)°C−1, α2= 7.494 × 10−6+ 5.168 × 10−9( T - 25)°C−1, α3=−0.842 × 10−6− 1.497 × 10−9( T - 25)°C−1. The expansion coefficient α1, nearly along [302] was approximately 3 times α2 along the b -axis. Very small contraction (α3) occurred nearly along [     01]. The volume changes upon martensitic transformations of β↔αL' were very small, and the strain accommodation would be almost complete. This is consistent with the thermoelasticity.  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