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.
Using the re-equilibration kinetic method the chemical diffusion coefficient in nonstoichiometric chromium sesquisulfide, Cr2+yS3, has been determined as a function of temperature (1073–1373 K) and sulphur vapour pressure (10?104 Pa). It has been found that this coefficient is independent of sulphur pressure and can be described by the following empirical equation: D?Cr2+yS3=50.86 exp(-39070 cal/mole/RT) (cm2s?1). It has been shown that the mobility of the point defects inCr2+yS3 is independent of their concentration and that the self-diffusion coefficient of chromium in this sulfide has the following function of temperature and sulphur pressure: DCr=2.706×102P?14.85S2exp(-56070 cal/mole/RT). (cm2s?1).  相似文献   

2.
The self-diffusion coefficient of manganese in manganous sulphide has been calculated as a function of temperature and sulphur vapour pressure. It has been shown that near the Mn/MnS phase boundary Mn self diffusion occurs by means of interstitial or interstitialcy mechanism and DMn is the following function of temperature and sulphur vapour pressure: DMn = 0.252 PS2?16exp (?269 kJ/mol/RT). At higher sulphur pressures manganese diffuses via doubly ionized cation vacancies and analogous pressure and temperature dependence can be described by the following empirical equation: DMn = 6.70 × 10?4 PS216exp(?121 kJ/mol/RT).  相似文献   

3.
Diffusional transport of manganese through MnO scales has been studied at temperatures from 900–1200°C by means of a novel diffusion/evaporation method and by the two-stage oxidation (Rosenburg) method. In the phase field of Mn1?vO near the MnO3O4 phase boundary the manganese self-diffusion coefficient and the concentration of defects is concluded to be proportional to p12o2. The concentration of defects increases with decreasing temperature at constant partial pressure of oxygen. In this non-stoichiometry range Mn1?yO has properties similar to wüstite, Fe1?yO, and it is concluded that the important cation defects are defect clusters. The results may be explained assuming that the defect clusters consist of four vacancies on octahedral sites and one interstitial on a tetrahedral site. Values for the concentration of defects, the self-diffusion of manganese ions and of the defects have been calculated. For the phase field near the M/MnO phase boundary it is concluded that Mn-interstitials predominate.  相似文献   

4.
A method for studying metal ion self-diffusion in oxides (or other inorganic compounds) is described. The method involves oxidation of an appropriate metal to form a dense, single-layered scale of the lowest valent oxide (e.g. MnO on Mn). The specimen is then treated in high vacuum, and the evaporation of metal diffusing through the scale is measured. From the rates of metal diffusion/evaporation as a function of scale thickness information about the defect structure is obtained. The metal ion self-diffusion coefficient is determined from the rate of metal transport (evaporation) through a scale with known thickness. The requirements and limitations of the method are discussed. The use of the method is illustrated for Mn self-diffusion in MnO at 1100°C. The self-diffusion coefficient of Mn in MnO is proportional to the square root of the oxygen pressure, DMn ∞ pO212, in t MnO phase field near the MnO/Mn3O4 phase boundary. It is also tentatively concluded that the predominating defects near the Mn/MnO phase boundary are manganese interstitials.  相似文献   

5.
6.
We report the results of a search for all integrable hamiltonian systems of type H = (12)px2 + (12)py2 + V(x,y), where V is a polynomial in x and y of degree 5 or less and the second invariant is a polynomial in px and py of order 4 or less. Both classical and quantum integrability are discussed.  相似文献   

7.
Lithium-containing oxide bronzes LixV2?yMyO5-β (M=Mo or W) have been studied in wide range of temperature (25–550°C) and composition (0.22?x?0.47, 0?y?0.25) as prospective ionic-electronic conductors. By coulometric titration tecknique, ΔGLi, ΔHLi and ΔSLi have been found and ordering of the lithium ions within the channel structure LixV2O5-β has been discussed depending on composition and temperature. The chemical diffusion coefficient for lithium D?Li has been measured by potentiometric technique on hot-pressed ceramic samples. The lithium self-diffusion coefficient DLi and partial lithium-ionic conductivity have been calculated.  相似文献   

8.
We report the first piezomodulation spectrum of Cd1?xMnxTe. Signatures of the exciton transition and the internal transition of Mn++ are identified. The opposite signs of these signatures are attributed to the positive pressure coefficient of the direct interband transition as contrasted to a negative pressure coefficient for the 6A14T1 transition causing the Mn++ feature.  相似文献   

9.
The activation barrier ΔE1ABfor dissociation AB → A + B on transition-metal surfaces is analyzed within an additive Morse-type approach based on the bond-order conservation. It is shown that ΔE1AB = DAB?(QA + QB + QAQB/(QA + QB) where DAB is the gas-phase dissociation energy and QA(QB) is the heat of atomic chemisorption. Estimates of ΔE1 for H2, N2, O2, and NO are shown to be in reasonable agreement with experiment. The two-dimensional potential diagram of the metal-AB interactions is defined analytically and discussed in some detail.  相似文献   

10.
Using a solution to the inverse scattering problem we have generated phase-equivalent separable potentials in the 1S0 and 3S1?3D1 states, which have nearly the same singlet UPA form factors and deuteron parameters (ED, PD, QD, AS and ADAS) as the Reid soft-core potential. We compare our results for the binding energy of the triton and the neutron-deuteron doublet scattering length with the corresponding values for the Reid soft-core potential.  相似文献   

11.
Predissociations in the y1Πg and x1Σg? Rydberg states of N2 (configurations u?14pσ and u?13pπ, respectively) and their likely causes, are discussed. Peaking of rotational intensity at unusually low J values, without sharp breaking off, is interpreted as due to case c? or case ci predissociation. Λ doubling in the y state, attributed to interactions with the x1Σg? state and with another, 1Σ+, state of the same electron configuration as x, is analyzed. From this analysis the location of the (unobserved) 1Σg+ state, here labeled x′, is obtained. It is concluded that the predissociation in the Π+ levels of the y state is an indirect one mediated by the interaction with x′ coupled with predissociation of x′ by a 3Σg? state dissociating to 4S + 2P atoms: combined, however, with perturbation of the y state by the k1Πg Rydberg state (configuration g?14dπ), whose Π+ levels are completely predissociated.  相似文献   

12.
The 5D3?5D3 cross-relaxation of YAG:Tb3+ has investigated by analysis of the 5D3 decay curve as a function of Tb3+ concentration. This technique is a more sensitive test of the mechanism than relative intensity measurements alone. It is shown that the relaxation is predominantly dipole-dipole in character (R0 ~ 1.3nm), and insensitive to temperature. The population of the 5D4 state following high energy excitation occurs almost entirely via cross-relaxation from the 5D3 state.  相似文献   

13.
Departure from stoichiometry in vapor grown FeCr2S4 was studied using Mössbauer spectroscopy. The paramagnetic Mössbauer spectra give evidence of two singlets and two doublets which correspond respectively to A site Fe2+ ? Fe3+, FeII in Td symmetry and in symmetry lower than Td. The following ionic distribution has been deduced:
(Fe2+1?yFe3+y)|Cr3+2?xx|S4?zz
Compounds in the system Fe1+xCr2?xS4 have been studied for 0 ? x ? 0.1. The spectra are solved assuming FeII in A site with Td symmetry, A site FeII with lower symmetry and B site Fe3+. No Fe2+ appears in B site. These features are discussed in terms of schematic band structures implying single electron narrow bands. The non-affinity of Fe2+ for B sites of iron thiochromites is discussed in relation with B site Cr2+ level.  相似文献   

14.
The wavenumbers of the rotation-vibration lines of 14N16O are reported for the (2-0) and (3-0) bands. The full set of spectroscopic constants for the three bands (1-0), (2-0), and (3-0) has been determined with the method developed by Albritton, Schmeltekopf, and Zare for merging the results of separate least-squares fits. The vibrational constants ωe, ωexe, ωeye, and the vibrational dependence of the rotational constants have been deduced. The apparent spin-orbit constant A?v and its centrifugal correction A?D (including the spin-rotation constant) have a vibrational dependence of the following form: A?v = A?e ? αA(v + 12) + γA(v + 12)2 and A?Dv = A?De ? βA(v + 12) + δA(v + built+12)2; the values of the constants in these two equations have been determined.  相似文献   

15.
Diffusion of 54Mn in Mn1?δO single crystals has been measured by a serial sectioning technique as a function of temperature (1000–1500°C) and deviation from stoichiometry (0.00003 < δ < 0.12). The value of m in the expression D = D0(T)pO21m varies from about 6 at low Po2 at all temperatures to a value approacing 2 at high Po2 and high temperatures, thus suggesting that diffusion occurs by doubly charged vacancies at low Po2 with increasing contributions from singly charged and neutral vacancies as Po2 (and vacancy concentration) increases. For δ near 0.1, the values of D fall below the values extrapolated from smaller defect concentrations. The isotope effect for cation self-diffusion was measured by simultaneous diffusion of 52Mn and 54Mn in Mn1?δO (0.0004 < δ < 0.116) at 1300 and 1500°C. The measured values of fΔK are independent of temperature within experimental error, and decrease from a value of 0.70 at low defect concentrations to 0.37 for large values of δ. The isotope-effect results suggest that diffusion occurs by single non-interacting vacancies at low defect concentrations; defect-defect interactions become important for δ ? 0.01. The defect-defect interactions may involve essentially individual defects or may result in defect clusters; the similarity between the present isotope-effect results and those for Fe1?δ0 suggests that defect clustering plays a significant role in mass transport in Mn1?δO at large values of δ.  相似文献   

16.
S. Fujita 《Physica A》1977,89(1):127-138
The self-diffusion (spin-diffusion) coefficient D of an electron gas is investigated by means of a proper connected diagram expansion which treats a collision in a medium (rather than in vacuum) in a natural manner. By calculating real and virtual interaction processes to the order e4 (second order in interaction strength), we obtain
D = 0.3597 ?12 e-2(kBT)54 n0-12 M-34
for a non-degenerate gas where n0 represents the density. The square-root density (n0) dependence is noteworthy. In the calculation, no cut-off limiting low momentum transfer in collision is introduced; the screening effect which is important at higher densities, is neglected.  相似文献   

17.
Self-diffusion of 59Fe parallel to the c axis in single crystals of Fe2O3 has been measured as a function of temperature (1150–1340°C) and oxygen partial pressure (2 × 10?3 ? pO2 ? 1 atm) The temperature dependence of the cation diffusivity in air is given by the expression
DFe1 = (1.9?1.4+5.2 × 109exp(?141.4 ± 4.0 kcal/moleRT) cm2/s
.The unusually large value of D0 is interpreted in terms of the values of the preexponential terms in the reaction constants for the creation of defects in Fe2O3. The oxygen-partial-pressure dependence of the diffusivity indicates that cation self-diffusion occurs by an interstitial-type mechanism The simultaneous diffusion of 52Fe and 59Fe has been measured in Fe2O3. The small value of the isotope effect suggests that iron ions diffuse by an noncollinear interstitialcy mechanism, which is consistent with the crystal structure of Fe2O3.  相似文献   

18.
Three-, two-, and one-dimensional disordered systems with randomly distributed, purely repulsive scattering centers, known as Lorentz models, are studied in the low energy limit [1]. Using a functional integral representation and a version of the “replica trick”, we have found in the D-dimensional system the density of electronic levels of the form
n(E)=b0exp(?b1E?(D2)+b2E?(D2)+1+·+bDE?(12))(1+O(E))
and the constants b0, b1,…, bD, and γ have been determined.  相似文献   

19.
The diffusion of water into additively colored potassium iodide has been studied in the range 15–45°C. Penetration depths, measured by decrease in the F-band absorption, increase with t12. The diffusion coefficient, D = 0·58 exp (?6496/T) cm2 sec?1 agrees very well with that determined by other workers. The Henry's law constant, K = C0pw = 1·3 × 109exp (+4882/T) cm?3 torr?1 implies a water concentration of C0 ? 1017 molecules per cm3 in the surface of KI crystals in equilibrium with an environment at 25°C and 35 per cent relative humidity. The large C0 makes penetration very rapid. Diffusion occurs by interstitial migration of water molecules with an entropy of activation of 9.4 cal/mol deg and an enthalpy of activation of 12·9 kcal/mol.  相似文献   

20.
Two electrochemical methods - involving the application of a long-time galvanostatic current pulse and a small potentiostatic voltage step to a M/MxSSE cell - are presented. From the overvoltage, respectively current response the chemical diffusion coefficient (DM+) and the thermodynamic factor (? ln a/? ln c) are obtained. The methods have been applied to the cells: Li/1M·LiClO4 in propylenecarbonate/LixTi1.03S2 0.05 < x < 0.95, T = 20°C; and LixCoO2 0.10 < × < 1, T = 20°C. From the application of the current pulse/voltage decay method it followed: DLi+(LixTi1.03S2) = 1?4 × 10?8cm2s?1, with a slight tendency to increase with decreasing x; DLiC(LixCoO2) = 2?40 × 10?9cm2s?1, decreasing with decreasing x. These values are among the highest found for solid state Li+-ion diffusion, and will be closely evaluated and compared with data reported by other workers. The x-dependence of the thermodynamic factor, determined from kinetic data, for LixTi1.03S2 (x = 0.05-0.95) and LixCoO2 (x = 0.60-1.00) is in accordance with a simple thermodynamic model. Unlike for LixTi1.03S2, the thermodynamic factor for LixCoO2, determined from the EMF-x relation, cannot be accounted for by this model. Furthermore, a fast, but crude method to determine the average chemical diffusion coefficient in LixTi1.03S2 and LixCoO2 is discussed.  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