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 46 毫秒
1.
Diffusion of 57Co isotope on NiO (110) and AI2O3 (0001) surfaces was investigated by using the edge source method. The surface diffusion parameter α D ,Δ, where α is the segregation factor, D , the surface diffusion coefficient, and Δ the thickness of the high-diffusivity layer, was determined for the temperature range 750°–1200°C. For calculation of experimental results, the Whipple solution was used. The Arrhenius plot shows a break at ∽900°C for the surface diffusion of 57Co isotope on AI2O3 (0001) plane. Above this temperature, vapor transport seems to be the overriding diffusion mechanism. Below this temperature, ionic transport predominates. The apparent activation energy for the ionic transport was calculated to be 120 ± 12 kj/mol.
Ionic transport predominated in the surface diffusion of 57Co on NiO over the entire investigated temperature range. This can be explained by the weak bond between Co-vapor species and the NiO surface. The results obtained suggest that the surface diffusion of Co2+ ion on NiO at 750°C is ∼7 orders and at 1200°C ∼5 orders of magnitude faster than volume diffusion. Activation energies are 139 and 227 kJ/mol, for surface and volume diffusion, respectively.  相似文献   

2.
Diffusion of 51Cr isotope on MgO (100), Al2O3 (0001, and MgAl2O4 (111) surfaces was investigated by using the edgesource method. The surface diffusion parameter, αD,δ, where α is the segregation factor, D , the surface diffusion coefficient, and δ the thickness of the high-diffusivity layer, was determined for the temperature region 650° to 1250°C. For calculation of experimental results Whipple's solution was used. Arrhenius plots show a break at ∼1050°C for MgO and at ∼900°C for MgAl2O4. Above these temperatures vapor transport seems to be the overriding diffusion mechanism. Below these temperatures ionic transport predominates. Ionic transport is the predominant mechanism for surface diffusion of 51Cr on Al203 over the entire investigated temperature region. This can be explained by the weak bond between Cr-vapor species and Al203 surface. The apparent activation energies for ionic transport of 51Cr are 110 ± 12, 121 ± 12, and 119 ± 12 kJ/mol on MgO, Al2O3, and MgAl2O4 surfaces, respectively. They include enthalpy of motion and binding enthalpy and are 2.5 to 2.8 times smaller than the activation energies for volume diffusion. Investigations of surfaces by LEED, Auger, and SIMS indicated structural nonperiodicity and surface segregation of impurities.  相似文献   

3.
Diffusion of the 57Co isotope on Fe3O4 (110) and NiO (100) surfaces was investigated by the edge-source method. The surface diffusion parameter, α D sδ, where α is the segregation factor, D s the surface diffusion coefficient, and δ the thickness of the high-diffusivity layer, was determined at 750°C for different partial pressures of oxygen. In both cases point defects strongly influenced surface diffusion. For Fe3O4 the vacancy mechanism is dominant at high oxygen activities and interstitial (or interstitialcy) mechanism dominates at low oxygen activities. For NiO the surface diffusion at low oxygen activities is influenced by aliovalent impurities and at high oxygen activities the strong influence of the intrinsic defects, nickel vacancies, on surface diffusion was observed. It was concluded that similar mechanisms operate during surface diffusion in the near surface layer and during diffusion in the lattice.  相似文献   

4.
This paper reports the transport kinetics of Mg in cubic yttria-stabilized zirconia (containing 10% mol of Y2O3 (10YSZ)) involving the bulk and the grain boundary diffusion coefficients. The diffusion-controlled concentration profiles of Mg were determined using secondary ion mass spectrometry (SIMS) in the range 1073–1273 K. The determined bulk diffusion coefficient and the grain boundary diffusion product may be expressed as the following functions of temperature, respectively: D = 5.7 exp[(−390 kJ/mol)/ RT ] cm2·s−1 and D 'αδ= 3.2 × 10−15 exp[(−121 kJ/mol)/ RT ] cm3·s−1, where α is the segregation enrichment factor and δ is the boundary layer thickness. The grain boundary enhancement factor decreases with temperature from 105 at 1073 K to 103 at 1273 K.  相似文献   

5.
Grain-boundary diffusion of 60Co and 51Cr in bicrystals and polycrystals of NiO was measured at temperatures <0.6 Tm at various equilibrium oxygen partial pressures, using a tracer-sectioning technique. Values of D 'δ were obtained using the Whipple and Suzuoka analyses. The results show that, for both 60Co and 51Cr, grain-boundary-enhanced diffusion is observed and the temperature dependence of D 'δ is about the same as that of the respective bulk diffusivity,/). The effects of equilibrium oxygen partial pressure and the type of gram boundary on D 'δ are not detectable within experimental error.  相似文献   

6.
The thermal expansion of the hexagonal (6H) polytype of α-SiC was measured from 20° to 1000°C by the X-ray diffraction technique. The principal axial coefficients of thermal expansion were determined and can be expressed for that temperature range by second-order polynomials: α11= 3.27 × 10–6+ 3.25 × 10–9T – 1.36 × 10–12 T 2 (1/°C), and ş33= 3.18 × 10–6+ 2.48 × 10–9 T – 8.51 × 10–13 T 2 (1/°C). The σ11 is larger than α33 over the entire temperature range while the thermal expansion anisotropy, the δş value, increases continuously with increasing temperature from about 0.1 × 10–6/°C at room temperature to 0.4 × 10–6/°C at 1000°C. The thermal expansion and thermal expansion anisotropy are compared with previously published results for the (6H) polytype and are discussed relative to the structure.  相似文献   

7.
The diffusive transport of chromium in both pure and Y-doped fine-grained alumina has been investigated over the temperature range 1250°–1650°C. From a quantitative assessment of the chromium diffusion profile in alumina, as obtained from electron microprobe analysis, it was found that yttrium doping retards cation diffusion in the grain-boundary regime by over an order of magnitude. The Arrhenius equations for the undoped and Y-doped samples were determined to be: δ D b=(4.77±0.24) × 10−7 exp (−264.78±47.68 (kJ/mol)/RT)(cm3/s) and δDb=(6.87±0.18) × 10−8 exp (−284.91±42.57 (kJ/mol)/RT)(cm3/s), respectively. Finally, to elucidate the mechanism for this retardation, the impact of yttrium doping on diffusion activation energies and prefactors was examined.  相似文献   

8.
The grain-boundary diffusion product, D'δ , of 51Cr in MgO and Cr-doped MgO as a function of grain-boundary orientation and point-defect concentration was determined at T =1200° to 1450°C. A large degree of anisotropy was found in the grain-boundary diffusion behavior in MgO. The ratio of D'δ|| parallel to D'δ perpendicular to the growth direction, D'||/D' , is 102 for a 5° (100) tilt boundary, decreased to ∼2 in boundaries with tilt angles > 10°. The decrease in D'||/D' is due to a large increase in D' with increasing tilt angle. The results indicate that grain-boundary diffusion in MgO is connected to the orientation of dislocations and the mechanism is one of dislocation pipe diffusion. The grain-boundary diffusion product D'δ increases with increasing Cr concentration in MgO and is ∼4 times larger for MgO containing 0.56 at. % Cr than for the undoped MgO. For all bicrystals studied, the activation energies are within 180 ± 20 kJ/mol which is 60% of the activation energy for 51Cr diffusion in undoped MgO.  相似文献   

9.
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   

10.
Alpha uranium dicarbide disks coated with pyrolytic carbon migrate, relative to the carbon, up a temperature gradient, leaving behind a layer of rejected graphite. A model for the migration process based on solution of carbon at the hot UC2-C interface, transport across the UC2 phase by thermal diffusion, and rejection as graphite at the cool UC2-C interface was developed. The rate of migration may be described by: The apparent activation energy for migration is 76.5±9 kcal/mol, compared to a literature value for carbon self-diffusion in α-UC2 of 90.2 kcal/mol. This difference in temperature dependence may be considered preliminary evidence that the heat of transport decreases with increasing temperature. The average value of the heat of transport, calculated using literature values for carbon self-diffusion in α-UC2, is 92±27 kcal/mol.  相似文献   

11.
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   

12.
The vapor pressure of SmC2 in equilibrium with graphite was measured by the Mnudsen effusion technique. Rates of weight loss from the cells were measured with an automatic recording balance. The apparent pressures varied with orifice size, and equilibrium pressures were calculated by extrapolation to zero orifice area. This work was combined with other studies to obtain log10 P(atm) = - 13.869 × 103/T + 3.752 (1300°-2050°K) for the Sm vapor pressure above SmC2-C. Estimates of S°298 and cp were made for SmC2, and δH°298 was calculated to be 72.0 ± 2 kcal/mol for the reaction SmC2(s) = Sm(g) + 2C(s). This value combined with δH°v, 298= 48.6 kcal/mol for Sm gives a δ°f298 for SmC2 of 23.4 ± 2 kcal/mol.  相似文献   

13.
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   

14.
Bi2Sr2Ca2Cu2O8±δ-type compound thick films were exposed to oxygen-argon-gas mixtures (1% to 20% oxygen gas) at elevated pressures (up to 207 MPa) and temperatures (500° to 940°C) for times ranging from 5 to 96 h. At a sufficiently high oxygen fugacity and temperature, Bi2Sr2Ca1Cu2O8±δ decomposed via a solid-state reaction. Room-temperature X-ray diffractometry and electron probe microanalysis of decomposed films revealed the presence of Bi2(Sr,Ca)2-Cu1O6±θ ro-type compound, Bi2Sr2,Ca1O8±δ-type compound, and CuO. Bi2Sr2Ca1Cu2O8±δ decomposition was accompanied by a modest weight gain, which was consistent with an oxidation reaction. The solid-state decomposition reaction could be reversed by heat treatment of decomposed films at 860°C in pure, flowing oxygen at ambient pressure.  相似文献   

15.
Thermal decomposition of silicon diimide, Si(NH)2, in vacuum resulted in very-high-purity, fine-particle-size, amorphous Si3N4 powders. The amorphous powder was isothermally aged at 50° to 100° intervals from 1000° to 1500°C for phase identification. Examination of ir spectra and X-ray diffraction patterns indicated a slow and gradual transition from an amorphous material to a crystalline α-phase occurring at 1200°C for >4 h and/or 1300° to 1400°C for 2 h. As the temperature was increased to ≥1450°C for 2 h, the crystalline β-phase was observed. Phase nucleation and crystallite morphology in this system were studied by electron microscopy and electron diffraction combined with TG as functions of temperature for the inorganic polymer starting materials. Powders prepared in this manner with 4 wt% Mg3N2 added as a sintering aid were hot-pressed to high-density fine-grained bodies with uniform microstructures. The optimum hot-pressing condition was 1650°C for 1 h. Silicon concentration steadily increased as the hot-pressing temperature or time was increased. A method for chemical etching for high-density fine-grained Si3N4 is described. Electrical measurements between room temperature and ∼500°C indicated dielectric constant and tan δ values of 8.3±0.03 and 0.65±0.05×10−2, respectively.  相似文献   

16.
In this paper we map out the 1300°C isothermal section in the Ti–In–C ternary system. Two ternary compounds exist: Ti2InCα and Ti3InCδ. At 1300°C TiC x is in equilibrium with all phases, Ti3InCδ is in equilibrium with all the phases except C, and Ti2InC is in equilibrium with all phases except Ti and C. The range of δ in Ti3InCδ varies from 0.95 to 0.8. A correlation was found between δ and the lattice parameters of this phase. The maximum solubilities of In and C in Ti are 14 and 4 at.%, respectively. Similarly, α in Ti2InCα varies from ≈0.85 to 1. The dissolution of ∼4±0.3 at.% In in TiC x reduces the C concentration to ≈24 at.%. In the In-rich corner, a liquid region exists.  相似文献   

17.
Vaporization of Lead Zirconate-Lead Titanate Materials   总被引:2,自引:0,他引:2  
A thermogravimetric investigation of the vaporization of cold-pressed Pb(Zr0.65Ti0.35)O3 in vacuum from 690° to 1130°C shows that the vaporization occurs in two steps. An initial loss of 1 to 7 wt% proceeds by a mechanism with a logarithmic time dependence and represents the vaporization of the unreacted PbO that remains after the initial calcining. The second step proceeds by a slower, diffusion-controlled mechanism with a parabolic time dependence. It is shown that this rate-determining step is bulk diffusion across a thickening lead-depleted layer in the cold-pressed pellet, despite pore volumes of 24 to 42%. The rate constant, g/cm2 sec½, for the parabolic weight loss process is expressed by: log K = (3.37 ± 0.14) - (8.46 ± 0.16) (103/ T K). The activation energy is 38.7 ± 2.0 kcal/mole.  相似文献   

18.
The dissipation factors and dielectric constants of alumina ceramics containing less than 100 ppm of impurities and of specimens doped with Si, Ti, Ca, Mg, Fe, and Cr ions were measured in the region 102 to 8.5 × 109 cps and 25° to 875° C. Multiple regression analysis of the data at 500° C and 106 cps showed a linear relation between impurity concentration and tan δ, with a correlation coefficient of 0.93. Si ions caused the greatest rise of tan δ, Mg and Ti were second, Ca third, and Cr and Fe had no significant influence. These effects diminished with rising frequency and became negligible in the microwave region. Activation energies of conduction for pure and doped alumina were estimated from measurements of δ at 105 cps and 500°C. Values between 1.2 and 1.6 ev were calculated for all compositions except the one containing Mg2+, for which 2.0 ev was obtained. At low frequencies, the dielectric constant (k') rose exponentially with temperature, reflecting a similar rise in the number of free charge carriers contributing to interfacial polarization. At higher frequencies the temperature variation of k' fell to a shallow positive slope of about 120 ppm per °C. This coefficient was not influenced by low concentrations of impurities but could be effectively compensated without excessive loss by additions of 10 to 20% SrTi03.  相似文献   

19.
Temperature dependence of lattice parameters of bismuth sesquioxide (Bi2O3) has been obtained between 26° and 778°C by the Rietveld method using neutron powder diffraction data. Lattice parameters     and unit-cell volume of α and δ phases increased, while the β angle of α phase decreased with an increase of temperature. Here the     denotes the lattice parameter a of the α phase on the basis of pseudo-fluorite lattice. The thermal expansion coefficients of α phase were 26.7, 6.6, and 9.0 (× 10−6°C−1) for     and     axes, respectively, indicating a large anisotropy. The α- to -δ phase transition of Bi2O3 was confirmed between 721° and 760°C on heating. At the α–δ transition point, the lattice parameters and unit-cell volume discontinuously changed, indicating that the transition is of the first order.  相似文献   

20.
The early stages of the reduction kinetics of single-crystal rutile, TiO2, were studied by optical absorption techniques for specimens re duced in the range 550° to 760°C at a constant vacuum level of 10−3 mm Hg. When reduced to a room-temperature resistivity value of ∼103 ohm-cm, rutile shows a blue coloration as a re sult of a broad infrared absorption band centered at a wavelength of 1.2μ. The kinetics of growth of the optical absorption coefficient at 1.2μ fit a parabolic relation of the form (Δα)2= K't. The Arrhenius rate constant, K' , exhibited an activation energy of 111 ± 9 kcal/mole which is ∼50% too high for the diffusion of oxygen in rutile. It is postulated that the defect center is likely to be a Ti interstitial and that the kinetics reflect the self-diffusion of Ti as the rate-con trolling mechanism.  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