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 oxygen tracer diffusion coefficient (D?) has been measured for 9 mol% scandia 2 mol% yttria co-doped zirconia solid solution, (Y2O3)2(Sc2O3)9(ZrO2)89, using isotopic exchange and line scanning by Secondary Ion Mass Spectrometry, as a function of temperature. The values of the tracer diffusion coefficient are in the range of 10? 8–10? 7 cm2 s? 1 and the Arrhenius activation energy was calculated to be 0.9 eV; both valid in the temperature range of 600–900 °C. Electrical conductivity measurements were carried out using 2-probe and 4-probe AC impedance spectroscopy, and a 4-point DC method at various temperatures. There is a good agreement between the measured tracer diffusion coefficients (D?, Ea = 0.9 eV) and the diffusion coefficients calculated from the DC total conductivity data (Dσ, Ea = 1.0 eV), the latter calculated using the Nernst–Einstein relationship.  相似文献   

2.
F. Christien  R. Le Gall 《Surface science》2011,605(17-18):1711-1718
Phosphorus surface segregation was measured by Auger Electron Spectroscopy on a 17-4 PH martensitic stainless steel at 450, 550 and 600 °C. Surface segregation was shown to be much faster than expected which was attributed to a high contribution of phosphorus diffusion along the former austenitic grain boundaries. A model of surface segregation was developed following the Darken–du Plessis approach and taking account of both bulk and grain boundary solute diffusion. The phosphorus grain boundary diffusion coefficient in 17-4 PH was estimated: DGB< = 6.2 104 exp(? 157 kJ mol? 1/RT)cm2 s? 1. It is found to be more than three orders of magnitude higher in 17-4 PH steel than in α-iron.  相似文献   

3.
The effect of indium-tin oxide (ITO) surface treatment on hole injection of devices with molybdenum oxide (MoO3) as a buffer layer on ITO was studied. The Ohmic contact is formed at the metal/organic interface due to high work function of MoO3. Hence, the current is due to space charge limited when ITO is positively biased. The hole mobility of N, N′-bis-(1-napthyl)-N, N′-diphenyl-1, 1′biphenyl-4, 4′-diamine (NPB) at various thicknesses (100–400 nm) has been estimated by using space-charge-limited current measurements. The hole mobility of NPB, 1.09×10−5 cm2/V s at 100 nm is smaller than the value of 1.52×10−4 cm2/V s at 400 nm at 0.8 MV/cm, which is caused by the interfacial trap states restricted by the surface interaction. The mobility is hardly changed with NPB thickness for the effect of interfacial trap states on mobility which can be negligible when the thickness is more than 300 nm.  相似文献   

4.
Growth rates of sucrose crystallization from pure solutions of initial relative supersaturation levels between 0.094 and 0.181 were studied in agitated crystallizer at 313.13 K. Birth and spread model was applicable for the obtained growth rate data in this range of supersaturation and used to estimate the principal growth parameters. The estimated interfacial free energy varied inversely with supersaturation from 0.00842 to 0.00461 J/m2, respectively. The obtained kinetic coefficient changed with the initial supersaturation from 9.45 × 10? 5 to 2.79 × 10? 7 m/s. The corresponding radius of the 2D (two dimensional) critical nucleus varied from 7.47 × 10? 9to 1.46 × 10? 9 m. Predominance of surface integration or volume diffusion mechanism during the growth process was assessed using the calculated activation free energies of the 2D nucleation process. An acceptable confirmation of the calculated radius of the critical 2D nucleus was found using atomic force microscopy (AFM) technique. The calculated interfacial free energy between the saturated sucrose solution and the crystal surface was found to be 0.02325 J/m2.  相似文献   

5.
《Solid State Ionics》2006,177(37-38):3285-3296
Oxygen nonstoichiometry, structure and transport properties of the two compositions (La0.6Sr0.4)0.99CoO3−δ (LSC40) and La0.85Sr0.15CoO3−δ (LSC15) were measured. It was found that the oxygen nonstoichiometry as a function of the temperature and oxygen partial pressure could be described using the itinerant electron model. The electrical conductivity, σ, of the materials is high (σ > 500 S cm 1) in the measured temperature range (650–1000 °C) and oxygen partial pressure range (0.209–10 4 atm). At 900 °C the electrical conductivity is 1365 and 1491 S cm 1 in air for LSC40 and LSC15, respectively. A linear correlation between the electrical conductivity and the oxygen vacancy concentration was found for both samples. The mobility of the electron-holes was inversely proportional with the absolute temperature indicating a metallic type conductivity for LSC40. Using electrical conductivity relaxation the chemical diffusion coefficient of oxygen was determined. It was found that accurate values of the chemical diffusion coefficient could only be obtained using a sample with a porous surface coating. The porous surface coating increased the surface exchange reaction thereby unmasking the chemical diffusion coefficient. The ionic conductivity deduced from electrical conductivity relaxation was determined to be 0.45 S cm 1 and 0.01 S cm 1 at 1000 and 650 °C, respectively. The activation energy for the ionic conductivity at a constant vacancy concentration (δ = 0.125) was found to be 0.90 eV.  相似文献   

6.
The equilibrated grain boundary groove shapes for solid carbon tetrabromide (CTB) in equilibrium with its melt were directly observed by using a horizontal temperature gradient stage. From the observed grain boundary groove shapes, Gibbs–Thomson coefficient (Γ) and solid–liquid interfacial energy (σSL) and grain boundary energy (σgb) of CTB have been determined to be (7.88 ± 0.8) × 10−8 K m, (6.91 ± 1.04) × 10−3 J m−2 and (13.43 ± 2.28) × 10−3 J m−2, respectively. The ratio of thermal conductivity of equilibrated liquid phase to solid phase for CTB has also been measured to be 0.90 at its melting temperature. The value of σSL for CTB obtained in present work was compared with the values of σSL determined in the previous works for same material and it was seen that the present result is in good agreement with previous works.  相似文献   

7.
In this study, the effects of ultrasound with different ultrasonic frequencies on the properties of sodium alginate (ALG) were investigated, which were characterized by the means of the multi-angle laser light scattering photometer analysis (GPC-MALLS), rheological analysis, circular dichroism (CD) spectrometer and scanning electron microscope (SEM). It showed that the molecular weight (Mw) and molecular number (Mn) of the untreated ALG was 1.927 × 105 g/mol and 4.852 × 104 g/mol, respectively. The Mw of the ultrasound treated ALG was gradually increased from 3.50 × 104 g/mol to 7.34 × 104 g/mol while the Mn of ALG was increased and then decreased with the increase of the ultrasonic frequency. The maximum value of Mn was 9.988 × 104 g/mol when the ALG was treated by ultrasound at 40 kHz. It indicated that ultrasound could induce ALG degradation and rearrangement. The number of the large molecules and small molecules of ALG was changed by ultrasound. The value of dn/dc suggested that the ultrasound could enhance the stability of ALG. Furthermore, it was found that ALG treated by ultrasound at 50 kHz tended to be closer to a Newtonian behavior, while the untreated and treated ALG solutions exhibited pseudoplastic behaviours. Moreover, CD spectra demonstrated that ultrasound could be used to improve the strength of the gel by changing the ratio of M/G, which showed that the minimum ratio of M/G of ALG treated at 135 kHz was 1.34. The gel-forming capacity of ALG was correlated with the content of G-blocks. It suggested that ALG treated by ultrasound at 135 kHz was stiffer in the process of forming gels. The morphology results indicated that ultrasound treatment of ALG at 135 kHz increased its hydrophobic interaction and interfacial activity. This study is important to explore the effect of ultrasound on ALG in improving the physical properties of ALG as food additives, enzyme and drug carriers.  相似文献   

8.
《Solid State Ionics》2006,177(9-10):821-826
The temperature dependence of the spin-lattice relaxation time, T1 and the line width of the 7Li nucleus were measured in delithiated LixCoO2 (x = 0.6, 0.8, 1.0). Two different relaxation behaviors were observed in the temperature dependence of T1 1 in a x = 0.8 sample. These would have arisen from inequivalent Li sites in two coexisting phases; an original hexagonal (HEX-I) and a modified hexagonal (HEX-II) phase in the x = 0.8 sample. We analyzed using a phenomenological non Debye-type relaxation model. Motional narrowing in the line width was observed in each sample, the result revealing that Li+ ions begin to move at low temperature in samples with less Li content. It was found that the activation energy associating with Li+ ion hopping in the HEX-II phase is smaller than that in the HEX-I phase. These results show that the HEX-II phase produced in the Li deintercalation process would be suitable for Li+ ionic diffusion in multi-phase LixCoO2, and it is expected that this would enable fast ionic diffusion. Li+ ionic diffusion related to phase transition is discussed from 7Li NMR results.  相似文献   

9.
《Solid State Ionics》2006,177(19-25):1747-1752
Oxygen tracer diffusion coefficient (D) and surface exchange coefficient (k) have been measured for (La0.75Sr0.25)0.95Cr0.5Mn0.5O3−δ using isotopic exchange and depth profiling by secondary ion mass spectrometry technique as a function of temperature (700–1000 °C) in dry oxygen and in a water vapour-forming gas mixture. The typical values of D under oxidising and reducing conditions at ∼ 1000 °C are 4 × 10 10 cm2 s 1 and 3 × 10 8 cm2 s 1 respectively, whereas the values of k under oxidising and reducing conditions at ∼ 1000 °C are 5 × 10 8 cm s 1 and 4 × 10 8 cm s 1 respectively. The apparent activation energies for D in oxidising and reducing conditions are 0.8 eV and 1.9 eV respectively.  相似文献   

10.
A Nd:Bi12SiO20 crystal was grown by the Czochralski method. The thermal properties of the crystal were systematically studied. The thermal expansion coefficient was measured to be α=11.42×10?6 K?1 over the temperature range of 295–775 K, and the specific heat and thermal diffusion coefficient were measured to be 0.243 Jg?1 k?1 and 0.584 mm2/s, respectively at 302 K. The density was measured to be 9.361 g/cm3 by the buoyancy method. The thermal conductivity of Nd:Bi12SiO20 was calculated to be 1.328 Wm?1 K?1 at room temperature (302 K). The refractive index of Nd:Bi12SiO20 was measured at room temperature at eight different wavelengths. The absorption and emission spectra were also measured at room temperature. Continuous-wave (CW) laser output of a Nd:Bi12SiO20 crystal pumped by a laser diode (LD) at 1071.5 nm was achieved with an output power of 65 mW. To our knowledge, this is the first time LD pumped laser output in this crystal has been obtained. These results show that Nd:Bi12SiO20 can serve as a laser crystal.  相似文献   

11.
Compressive creep tests have been performed on perovskite-type Ba0.5Sr0.5Co0.8Fe0.2O3 ? δ ceramics. The activation energy, stress exponent and inverse grain size exponent of the steady-state creep rates are evaluated at p(O2) = 0.21 ? 105 Pa and 0.01 ? 105 Pa in the stress, temperature and grain size ranges 5–20 MPa, 1078–1208 K and 2.5–17.4 µm, respectively. The results indicate that the creep rate of Ba0.5Sr0.5Co0.8Fe0.2O3 ? δ is controlled by diffusion of cations via both the oxide lattice (bulk diffusion) and along grain boundaries. The creep rate of Ba0.5Sr0.5Co0.8Fe0.2O3 ? δ increases profoundly by more than one order of magnitude at 1153–1178 K, which is tentatively linked with the onset of the hexagonal-to-cubic phase transition in this compound.  相似文献   

12.
Proton diffusion in [(NH4)1 ? xRbx]3H(SO4)2 (0 < x < 1) has been studied by means of 1H spin-lattice relaxation times, T1. The relaxation times were measured at 200.13 MHz in the range of 296–490 K and at 19.65 MHz in the range of 300–470 K. In the high-temperature phase (phase I), translational diffusion of the acidic protons relaxes both the acidic protons and the ammonium protons. Spin diffusion averages the relaxation rate of the two kinds of protons, whereas proton exchange between them are slow. The spin-lattice relaxation times in phase I were analyzed theoretically, and parameters of proton diffusion were obtained. The mean residence time of the acidic protons increases with increase in x for [(NH4)1 ? xRbx]3H(SO4)2 (0  x  0.54). Rb3H(SO4)2 does not obey this trend. The results of NMR well explain the macroscopic proton conductivity.  相似文献   

13.
S. Cohen  N. Shamir  M.H. Mintz  I. Jacob  S. Zalkind 《Surface science》2011,605(15-16):1589-1594
Auger-Electron-Spectroscopy (AES) and Direct-Recoils-Spectrometry (DRS) were applied to study the interaction of O2 with a polycrystalline gadolinium surface, in the temperature range 300–670 K and oxygen pressure up to 2 × 10? 6 Torr. It has been found that initial uptake of oxygen, at coverage measurable by the techniques used here, results in rapid oxide island formation. The subsurface is believed to be a mixture of oxide particles and oxygen dissolved in the Gd metal, the latter being the mobile species, even at relatively low temperatures.Enhanced inward diffusion of oxygen starts as early as 420 K and dictates the surface oxygen concentration and effective thickness of the forming oxide. The oxygen accumulation rate at the near-surface region, as measured by the O(KLL) AES signal intensity, goes through a maximum as a function of temperature at 420 K. This is a result of the combination of still efficient oxygen chemisorption that increases surface occupation and slow inward diffusion. The thickest oxide, ~ 1.7 nm, is formed at 300 K and its effective thickness was found to decrease with increasing temperature (due to oxygen dissolution into the metal bulk).Diffusion coefficients of the oxygen dissolution into the bulk were evaluated for various temperatures utilizing models for infinitely thin oxide layer and thick oxide layer, respectively. The best fit under our experimental procedure was obtained by the thick layer model, and the coefficients that were calculated are D0 = 2.2 × 10? 16m2s? 1 and Ea = 46kJ/mol.  相似文献   

14.
《Solid State Ionics》2006,177(1-2):129-135
LixV2O5 (0.4 < x < 1.4) prepared by solid-state reaction were studied by 7Li and 51V NMR spectroscopy. 7Li NMR spectra showed a narrowing of the line width in relation to Li+ionic diffusion. Analysis of LixV2O5 using a Debye-type relaxation model showed a low activation energy ∼0.07 eV in the sample of x = 0.4 below room temperature, and revealed a Li+ionic diffusion with larger activation energy ∼0.5 eV above 450 K in lithium-rich samples. The latter is ascribed to the existence of a multi-phase system comprising stable ɛ- and γ-phases, resulting from complicated phase transitions at high temperature. These shapes and shifts enable the classification of the β-, ɛ-, δ-, and γ-phases. The ionic diffusion of Li+ ions is discussed in relation to the complicated phase transitions.  相似文献   

15.
《Solid State Ionics》2006,177(37-38):3223-3231
Proton dynamics in (NH4)3H(SO4)2 has been studied by means of 1H solid-state NMR. The 1H magic-angle-spinning (MAS) NMR spectra were traced at room temperature (RT) and at Larmor frequency of 400.13 MHz. 1H static NMR spectra were measured at 200.13 MHz in the range of 135–490 K. 1H spin-lattice relaxation times, T1, were measured at 200.13 and 19.65 MHz in the ranges of 135–490 and 153–456 K, respectively. The 1H chemical shift for the acidic proton (14.7 ppm) indicates strong hydrogen bonds. In phase III, NH4+ reorientation takes place; one type of NH4+ ions reorients with an activation energy (Ea) of 14 kJ mol 1 and the inverse of a frequency factor (τ0) of 0.85 × 10 14 s. In phase II, a very fast local and anisotropic motion of the acidic protons takes place. NH4+ ions start to diffuse translationally, and no proton exchange is observed between NH4+ ions and the acidic protons. In phase I, both NH4+ ions and the acidic protons diffuse translationally. The acidic protons diffuse with parameters of Ea = 27 kJ mol 1 and τ0 = 4.2 × 10 13 s. The translational diffusion of the acidic protons is responsible for the macroscopic proton conductivity, as the NH4+ translational diffusion is slow and proton exchange between NH4+ ions and the acidic protons is negligible.  相似文献   

16.
《Solid State Ionics》2006,177(26-32):2363-2368
The mechanism and kinetics of water incorporation in the double perovskites Ва4Ca2Nb2O11 and Sr6Ta2O11 has been investigated (T = 300÷500 °C and aH2O = 1 · 10 3÷2.2 · 10 2). The formation of hydration products Ba4Ca2Nb2O11·xH2O and Sr6Ta2O11·xH2O (0.2 < x < 0.50) was limited by the diffusion of H2O. It has been found that the concentration dependences of H2O are the same for both samples: small increasing of H2O with increasing x. The temperature dependences of the chemical diffusion coefficients of water for compositions of Ba4Ca2Nb2O11·0.35H2O and Sr6Ta2O11·0.35H2O could be described with close activation energies of Ea = 0.38 ± 0.03 eV and Ea = 0.49 ± 0.03 eV, respectively. The chemical diffusion coefficients of water are nearly one order of magnitude smaller for tantalate Sr6Ta2O11. This result correlates with lower oxygen and proton conductivities in Sr6Ta2O11 as the consequence of lower mobilities.  相似文献   

17.
《Solid State Ionics》2006,177(26-32):2255-2259
Phase inversion spinning technique was employed to prepare dense perovskite hollow fiber membranes made from composition BaCoxFeyZrzO3−δ (BCFZ, x + y + z = 1.0). Scanning electron microscope (SEM) shows that such hollow fibers have an asymmetric structure, which is favored to the oxygen permeation. An oxygen permeation flux of 7.6 cm3/min cm2 at 900 °C under an oxygen gradient of 0.209 × 105 Pa/0.065 × 105 Pa was achieved. From the Wagner Theory, the oxygen permeation through the hollow fiber membrane is controlled by both bulk diffusion and surface exchange. The elements composition of fresh fiber and the fiber after long-term experiments were analyzed by energy-dispersive X-ray spectra (EDXS). Compared to the fresh fiber, sulphur was found on the tested hollow fiber membrane surface exposed to the air side and in the bulk, and Ba segregations occur on the tested hollow fiber membrane surface exposed to the air side. A decrease of the oxygen permeation flux was observed, which was probably due to the sulphur poisoning.  相似文献   

18.
《Solid State Ionics》2006,177(33-34):2873-2880
Proton dynamics in Cs3(HSO4)2(H2PO4) has been studied by means of 1H solid-state NMR as well as thermal analyses. The thermal analysis shows an endothermic peak at 408 K, which corresponds to a superprotonic transition. Above the transition temperature a mass loss is observed in a dry atmosphere, which is easily recovered in a conventional dry atmosphere below the transition temperature. The 1H magic-angle-spinning NMR spectra at room temperature show two peaks at 13.5 and 15.8 ppm, and a shoulder at 11.3 ppm from tetramethylsilane, demonstrating a presence of several inequivalent proton sites. Translational diffusion of protons takes place in both a room-temperature phase (RT) and a high-temperature phase (HT). In both phases reorientation of the SO4/PO4 tetrahedron limits the rate of the proton transport. The 1H mean residence times are estimated as Ea = 33 kJ mol 1 and τ0 = 0.97 × 10 9 s for phase RT from the second moment analysis and as Ea = 20 kJ mol 1 and τ0 = 5.0 × 10 12 s for phase HT from the analysis of the 1H T1 results.  相似文献   

19.
The steady-state oxygen permeation through dense La2NiO4 + δ ceramics, limited by both surface exchange and bulk ambipolar conduction, can be increased by deposition of porous layers onto the membrane surfaces. This makes it possible, in particular, to analyze the interfacial exchange kinetics by numerical modelling using experimental data on the oxygen fluxes and equilibrium relationships between the oxygen chemical potential, nonstoichiometry and total conductivity. The simulations showed that the role of exchange limitations increases on reducing oxygen pressure, and becomes critical at relatively large chemical potential gradients important for practical applications. The calculated oxygen diffusion coefficients in La2NiO4 + δ are in a good agreement with literature. In order to enhance membrane performance, the multilayer ceramics with different architecture combining dense and porous components were prepared via tape-casting and tested. The maximum oxygen fluxes were observed in the case when one dense layer, ~ 60 μm in thickness, is sandwiched between relatively thin (< 150 μm) porous layers. Whilst the permeability of such membranes is still affected by surface-exchange kinetics, increasing thickness of the porous supporting components leads to gas diffusion limitations.  相似文献   

20.
Flexible organic light-emitting devices (FOLEDs) based on multiple quantum well (MQW) structures, which consist of alternate layers of 2,3,5,6-Tetrafluoro-7,7,8,8,-tetracyano-quinodimethane (F4-TCNQ) and 4,4′,4″-tris-(3-methylphenylphe-nylamino)tripheny-lamine (m-MTDATA) have been fabricated. The Alq3-based device with double quantum well (DQW) structure exhibits the remarkable electroluminescent (EL) performances for the brightness of 23,500 cd/m2 at 14 V and the maximum current efficiency of 7.0 cd/A at 300.3 mA/cm2, respectively, which are greatly improved by 114% and 56% compared with the brightness of 10,958 cd/m2 at 14 V and the maximum current efficiency of 4.5 cd/A at 174.0 mA/cm2 for the conventional device without MQW structures. These results demonstrate that the EL performances of FOLEDs could be greatly improved by utilizing the novel MQW structures, and the reason for this improvement has also been explained by the effect of interfacial dipole and interfacial doping between F4-TCNQ and m-MTDATA in this article.  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