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 109 毫秒
1.
The self-diffusion of 44Ti has been measured both parallel to and perpendicular to the c axis in rutile single crystals by a serial-sectioning technique as a function of temperature (1000–1500°C) and oxygen partial pressure (10?14 ? 1 atm). The oxygen-partial-pressure dependence of. D1Ti indicates that cation selfdiffusion occurs by an interstitial-type mechanism and that both trivalent and tetravalent interstitial titanium ions may contribute to cation self-diffusion. At po2 = 1.50 × 10?7 atm where impurity-induced defects are unimportant,
D1Ti(∥c)=6.50+1.33?1.11exp?(66.11±0.56 kcalmoleRTcm2S
and
D1Ti(⊥c)= 4.55+1.78?1.28exp?(64.08±0.99)kcalmoleRTcm2S.
In the intrinsic region, the ratio D1Ti (⊥c)/D1Ti(∥c) was found to increase from 1.2 to 1.6 as the temperature decreased from 1500 to 1000°C. Computations based upon the defect model of Kofstad (involving the atomic defects Ti...iTi....iand V..o), of Marucco etal. (Ti....i and V..o), and of Blumenthal etal. (Ti...i and Ti....i) are compared with the experimental data on deviation from stoichiometry, electrical conductivity, cation self-diffusion and chemical diffusion in TiO2?x. These comparisons provide values of the defect concentrations, cation-defect diffusivities, electron mobility and reasonable values of the correlation factor for cation diffusion by the interstitialcy mechanism. Only the model of Kofstad is inconsistent with the data.  相似文献   

2.
Self-diffusion studies have been performed in the orientationally disordered or the so-called plastic phase of pivalic acid. Single crystals of high purity (99.9999%) containing 109?1010 dislocations m?2 have been used. Thin layers of pivalic acid labelled with 14C or tritium were deposited on sample surfaces. Concentrationpenetration curves were established by serial sectioning. Lattice self-diffusion coefficients D, were measured from 281 to 304.75K. At 281K, the value of D is independent of time. From 281 to 301K, D is given by: D(m2S?1) = (4.9 ± 0.3)10?4 exp [? (59± 1) kJ mole?1/RT].The activation enthalpy of the lattice self-diffusion is roughly equal to the heat of sublimation (Ls = 57 kJ mole?i) and in good agreement with values obtained from NMR. The mass factorf ΔK, where f is the correlation factor and ΔK is a correction factor, has been measured using isotope effect studies. Between 281 and 301K the value obtained is fΔK* 0.1+0.2?0.1  相似文献   

3.
The adsorption and nucleation of indium on clean (111) silicon surfaces are studied by a UHV molecular beam mass-spectrometric technique. The thermal accommodation of the adatoms on the surface is complete. At very low surface coverages θ, an adsorption energy of 57 kcalmole and a preexponential term τ0 of the Frenkel relation equal to 8 × 10?13 s are found from transient response measurements. The isosteric heat of adsorption Ea varies very slowly with θ, Ea is equal to 59 kcalmole for θ ~ 10?3 and 57 kcalmole for θ = 0.9. The nucleation occurs without supersaturation in an adsorbed layer near a monolayer.  相似文献   

4.
The proton spin-lattice relaxation time, T1, is measured as a function of temperature in α -(COOH)2·-2H2O, K2HgCl4· H2O and LiCHO·H2O. The relaxation is caused by 180° flips of the water molecules about their 2-fold axes and good agreement is obtained between calculated and observed values of T1. Empiricly the flip rate follows a classical Arrhenius equation: P· exp (? ΔH(RT)). A literature survey of values of P and ΔH obtained from similar investigations on other hydrates is given. The survey shows that the preexponential factor, P, is a function of the activation enthalpy, ΔH. P increases from 1012 to 1017 Hz when ΔH changes from 2 to 17 kcalmole. Using a dynamical rate theory as formulated by Feit, we find the flip rate is given by: K2· √(ΔH)· exp (K1ΔH)· exp (?ΔH(RT>)). This expression can be fitted to the observed data using K1 = 0.69 molekcal and K2 = 2 × 1011 Hz · (kcalmole)?12. Thus both the frequency factor, K2√ (ΔH), and the entropic factor, exp (K1ΔH), have been obtained for flipping water molecules in hydrates. The values of K1 and K2 are shown to be physically reasonable.  相似文献   

5.
Cation self-diffusion D1Fe, parallel to the c axis has been measured as a function of temperature (1100–1300°C) and oxygen partial pressure po2 (2 × 10?3-1 atm) in the same single crystals of Fe2O3 as those used by Chang and Wagner. Whereas the po2 dependence of D1Fe, observed by Chang and Wagner has been confirmed, the absolute value of D1Fe and the activation enthalpy for self-diffusion are much higher than those reported by them. The various diffusion studies indicate that cation self-diffusion occurs by an interstitial-type mechanism. However, the sample-to-sample variations in D1Fe, suggest that all diffusion measurements may have been performed on samples where the defect concentrations are impurity controlled. Impurity diffusion of 60Co, 51Cr, and 88Y has also been measured as a function of po2 at 1200°C. The results indicate that these impurities diffuse by an interstitialcy mechanism in Fe2O3.  相似文献   

6.
Auger electron spectroscopy (AES) has been employed to determine the relative coverage of oxygen on polycrystalline tungsten at high temperatures (1200 ?T ? 2500 K) and low O2 pressures (5 × 10?9 ?po2 ?5 × 10?6 Torr). We believe that this is the first demonstration that chemical analysis of solid surfaces by AES is possible even at temperatures as high as 2500 K. It is assumed that the relative oxygen coverage is directly proportional to the peak-to-peak amplitude of the first derivative of the 509 eV oxygen Auger peak. The experimental results illustrate the dependence of coverage on temperature and pressure, and it is shown that the results for low coverages may be described reasonably well by a simple first-order desorption model plus a semi-empirical expression for the equilibration probability (or sticking coefficient). On the basis of this approximate model, the binding energy of oxygen on tungsten is estimated as a function of coverage, giving a value of ~ 140 kcalmole in the limit of zero coverage.  相似文献   

7.
The diffusion of 48V in disordered VOx, crystals has been measured by a serial-sectioning technique as a function of temperature (1100–1500°C) over the homogeneity range 0.78 < x < 1.28. The temperature dependence of the cation tracer diffusivity at each fixed composition is characterized by an Arrhenius behavior in the temperature range 1100–1500°C. The Arrhenius parameters decrease rather sharply near the stoichiometric composition as the composition increases from the metal-rich to the metal-deficient regime; the activation energy for diffusion decreases from ~71 kcalmole to ~48 kcalmole, and the frequency factor decreases by nearly two orders of magnitude (from ~5 cm2s to ~0.05 cm2s). It is concluded that the significant difference in the cation self-diffusion behavior between the metal-rich and the metal-deficient VOx may be attributed to the significant differences in the defect structures of the two regimes. The various possible diffusion mechanisms are explored, and comparisons of the cation diffusion behavior in VOx, with that of the related transition-metal monoxides TiOx and Fe1?δO are made. It is concluded that the experimental results for the entire composition range are consistent with the process of diffusion occurring by the migration of monovacancies in equilibrium with defect clusters, the nature of the clusters being different for x < 1 and x #62;; 1.  相似文献   

8.
Isotope labelling experiments have established that the adsorption of O2 on the W(110) plane at 20 K leads first to the formation of a dissociated atomic layer. A weakly bound molecular species, α-O2, forms only when the atomic layer is essentially complete (O/W = 0.6). The desorption of α-O2 was found to be first order with an activation energy of E = 1.9 kcalmole and a frequency factor γ = 3 × 109 s?1. The activation energy is shown to be less than the enthalpy of desorption and the meaning of this result is discussed.  相似文献   

9.
Oxide and silver paste were applied on the BaO-doped Bi2O3 electrolyte and their behavior was studied as a function of temperature and oxygen partial pressure. Interface resistance of most oxide/electrolyte were of the same order of magnitude with those of Ag paste/electrolyte in air (300–500°C). A high electrode capacitance of (0.8–1.7)×10?2 F/cm2 was observed for the silver electrode at 450°C in the PO2 region of 1–10?5 atm.  相似文献   

10.
The diffusion of sulfur in nickel oxide single crystals has been investigated over the temperature range from 1000 to 1250°C. The measured data were found to deviate markedly from the error function complement dependence for diffusion from a constant source. The deviation is attributed to the migration of sulfur by the “double mode simultaneous diffusion mechanism.” The faster mode diffusion is suggested to be via nickel vacancies, and the slower mode diffusion is suggested to be via oxygen vacancies. The diffusivities for faster mode are given by Df = 2.94 exp[? 86.6 kcal/RT] cm2 sec?1 and, the slower mode, Ds = 1.08 × 10?9 exp [?32.8 kcal/RT]cm2sec?1.  相似文献   

11.
The adsorption and desorption of nitrogen on a platinum filament have been studied by thermal desorption techniques. Nitrogen adsorption becomes significant only after any carbon contamination is removed from the surface by heating the platinum filament in oxygen, and after the CO content in the background gas is reduced substantially. At room temperature nitrogen populates an atomic tightly bound β-state, E = 19 kcal mole?1. The saturation coverage of the (3-state is 4.5 × 1014 atoms cm?2. Formation of the (β-state is a zero order process in the pressure range studied. At 90 K two additional α1- and α2-desorption peaks are observed. The activation energy for desorption for the α2-state is 7.4 kcal mole?1 at low coverage decreasing to 3 kcal mole?1 at saturation of this state, 6 × 10 molecules cm?2. The maximum total coverage in the α-states was 1.2 × 1015 molecules cm?2. A replacement process between the β- and α-states has been observed where each atom in the (β-state excludes two molecules from the α-state.  相似文献   

12.
The defect structure of lanthanum-doped polycrystalline calcium titanate was investigated by measuring the oxygen partial pressure (100–10?18 atm.) dependence of the electrical conductivity at 1000° C and 1050°C. Two types of charge compensation were observed, namely electronic and ionic. For P02 < 10?15 atm. the carrier concentration was fixed by the amount of lanthanum (donor) added and the conductivity was found to be independent of oxygen partial pressure (electronic compensation). For higher oxygen partial pressure conditions (P22 > 10?13 atm.) the extra charge of the lanthanum was compensated by doubly ionized calcium vacancies (ionic compensation). In the ionic compensation region, a model involving a shear structure is discussed.  相似文献   

13.
In Co-doped TiO2?δ oxide films deposited on SrTiO3(100) substrates, a room-temperature ferromagnetism is found to occur only in a limited charge-carrier concentration interval from 2×1018? 5×1022 cm?3. This indirectly testifies that ferromagnetism in the aforementioned n-type semiconductor is associated with the exchange interaction of magnetic ions via conduction electrons rather than with the formation of Co clusters in the material. The magnetic moment per Co atom is 0.8μB in the TiO cubic phase and 0.5μB in the anatase tetragonal phase of TiO2.  相似文献   

14.
Abstract

The electrical conductivity of CaTi1?x Fe x O3-δ (x = 0.1) was measured by an alternating current van der Pauw technique versus oxygen partial pressure (10?30-1 atm) and temperature (450–1200°C). The results were interpreted to reflect n-type, ionic and p-type conductivity at respectively low, intermediate and high oxygen partial pressures. The apparent activation enthalpy for the ionic conductivity, interpreted to reflect the mobility of oxygen vacancies, was 0.87 eV. The enthalpy of intrinsic formation of electronic defects (apparent band gap E g) was 3eV. The results are compared with literature data for CaTi0.8Fe0.2O3-δ and with Fe-substituted SrTiO3 and discussed in terms of iron-oxygen vacancy association and ordering.  相似文献   

15.
The chemical diffusion coefficient of Cu2O has been obtained for an oxygen partial pressure near 5 10?4 atm as a function of the temperature in the range 700–900°C D? = 1 62 10?4 exp(?5140 ± 600 cal mol ?1)/RT cm2s?1 This was easily achieved according to the electrochemical method used for the preparation of gaseous mixtures whose Po2; is lower than 10?5 atm The slight difference observed with the previously published results by Maluenda, and obtained for Po2 values which increase with T between 10?4 and 0.21 atm, may be due to an oxygen partial pressure effect already observed in the case of CoO. An ambipolar treatment of the chemical diffusion, in the case of p-type semiconductor MaOb, oxides, has allowed us to express the chemical diffusion coefficient as a function of the concentration of the prevailing defects and of their diffusion coefficient In the case where the prevailing defects are cationic vacancies α times ionized we have shown that the expression D? = (1 + α)Dvα can be generalized to the A2O compounds This set of results has allowed us, according to the copper self diffusion data obtained recently by Peterson etal, to estimate the apparent enthalpy of formation of the catiomc vacancies ΔHf 23 ± 0 8 kcal mol?1.  相似文献   

16.
Electrical conductivity measurements on nickel oxide have been performed at high temperatures (1273 K<T< 1673 K) and in partial pressures of oxygen ranging from Po2 = 1.89 × 10?4 atm to Po2 = 1 atm. The po21n dependence of the conductivity decreases from about 14 for Po2 = 1 atm to smaller values for lower partial pressures of oxygen. The activation enthalpy for conduction increases for decreasing oxygen partial pressures (from 22.5 kcal mol?1 at Po2 = 1 atm to 26.0 kcal mol?1 for Po2 = 1.89 × 10?4 atm). This behaviour can be explained by the simultaneous presence of singly and doubly ionized nickel vacancies, with different energies of formation.Furthermore, chemical diffusion coefficient measurements have been performed in the same temperature range, using the conductivity technique, and leading to the result:
D? = 0.244 exp (?36,600RT) cm2 s?1
.  相似文献   

17.
The time evolution of the KLL Auger spectrum of carbon as a function of temperature is used to derive the kinetics of the surface diffusion and bulk-to-surface precipitation of carbon on polycrystalline nickel. The results show that the activation energy for the surface diffusion of carbon atoms on polycrystalline nickel is 6.9 ± 0.6 kcalmole, and the activation energy for bulk-to-surface precipitation is 9.4 ± 0.6 kcalmole. The dependence on the surface diffusion coefficient Ds (cm2s?1), on the absolute temperature T can be represented, over the experimental temperature range, 350–425° C, by: ln Ds = 10.27 ? 3568T.  相似文献   

18.
Quantitative high-resolution absorption spectroscopy was applied to the (0,0) violet band of CN. The CN radical was prepared in a furnace at 1421°K containing pure cyanogen gas. Since the calculated CN concentration is dependent on the controversial CN heat of formation, only the relationship, fυ = 6·84 X 10-3exp (0·354δ), where fυ is the excess over the initially assumed ΔH0f(CN) = 100·8 kcal/mole, could be directly determined in this study with an estimated error in fυ of ±20%. For δ = 0, our fυ is a factor of 4·8 smaller than an average value of 0·033±0· derived from other measurements. If this latter value of fυ is assumed, our relationship yields ΔH0f(CN) = 105·3±1· kcal/mole or D0(CN) = 7·66±0·05 eV. The rotational temperature and line widths for this band were also measured.  相似文献   

19.
A series of nano-crystalline ceria-based solid solution electrolyte, Ce0.8La0.2?x MgxO2?δ (x?=?0.0, 0.05, 0.10, 0.15, and 0.2), were synthesized via the polyvinyl alcohol (PVA) assisted combustion method, and then characterized to the crystalline structure, powder morphology, sintering micro-structure, and electrical properties. Present study showed that Ce0.8La0.2?x Mg x O2?δ was exceedingly stable as a cubic phase in all temperature range and exhibited fine crystals ranging from 15 to 20 nm. After sintering at 1,400 °C, the as-prepared pellets exhibited a dense micro-structure with 96 % of theoretical density. The electrical conductivity was studied using AC impedance spectroscopy and it was observed that the composition Ce0.8La0.1?Mg0.1O2?δ showed higher electrical conductivity of 0.020 S?cm?1 at 700 °C. The thermal expansion was measured using dilatometer technique in the temperature range 30–1,000 °C. The average thermal expansion coefficient of Ce0.8La0.1?Mg0.1O2?δ was 12.37?×?10?6 K?1, which was higher than that of the commonly used SOFC electrolyte YSZ (~10.8?×?10?6 K?1).  相似文献   

20.
The diffusion of 59Fe and 60Co has been measured in pure CoO and dilute iron-doped CoO, (Co1?cFecO, as a function of temperature (1000–1400°C) and oxygen partial pressure Po2), (10?7Po2 ≦ 0 21 atm) The enhancement factors for the diffusivities of iron and cobalt are nearly identical, which suggests that the primary cause of the enhancement is the increased concentration of charge-compensating cation vacancies with the addition of iron. The Fe ions dissolved in CoO appear to exist as a mixture of Fe2+ and Fe3+ ions, the fraction of iron ions in the three-plus state decreases with decreasing Po2 The simultaneous diffusion of 52Fe and 59Fe has been measured as a function of (itpo; at 1200°C The correlation factor for Fe impurity diffusion determined from the isotope-effect measurements is about the same as that for self-diffusion in CoO at high (itPo2 (2 × 10?3po2 ≦ 0 21 atm), but increases slightly with decreasing pO2 Both the enhancement-effect and isotope-effect experiments suggest that the nearestneighbor interactions between Fe ions and vacancies is small, and that the dissolved Fe ions do not have strongly bound electron holes.  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