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.
By the method of helium thermal desorption from submicrocrystalline palladium presaturated in the gaseous phase, the diffusion coefficient D eff and solubility coefficient C eff of helium are measured in the range P=0–3 MPa and T=293–508 K. The pressure dependence of C eff flattens at high pressures. At low saturation pressures, the temperature dependences of the diffusion and solubility coefficients may be divided into (1) high-temperature (400–508 K) and (2) low-temperature (293–400 K) ranges described by the exponentials D 1, 2=D 0exp (−E 1, 2 D /kT) and C 1, 2=C 0exp (−E 1, 2 S /kT). The energies of diffusion activation are E 2 D =0.0036±0.0015 eV and E 1 D =0.33±0.03 eV, and the solution energies are E 2 S =−0.025±0.008 eV and E 1 S =0.086±0.008 eV in the low-and high-temperature ranges, respectively. Mechanisms behind the diffusion and solution of helium are discussed.  相似文献   

2.
We studied adsorption and desorption of Xe and deuterium on Ni(111) using an optical differential reflectance technique. The main findings are: (i) the differential reflectance varies almost linearly with the surface densities of deuterium and Xe adatoms, and the signals can be described well with a three-layer model and the known dielectric responses of the surface layers: (ii) the adsorption of deuterium atT = 120 K follows the Langmuir kinetics, while the adsorption of Xe atT = 38 K follows the zeroth-order kinetics; (iii) nearT = 70 K, the rate of Xe desorption is almost coverage-independent with an activation energy ofE des = 4.4 ± 0.2 kcal/mol. Our analysis suggests that the Xe desorption is likely to be dominated by the escape rate from the corners of two-dimensional Xe islands.  相似文献   

3.
The diffusion of Au in dislocation-free or plastically deformed Si (1011 to 1013 dislocations/m2) was measured with the aid of the spreading-resistance technique. The Au profiles produced indislocation-free Si slices by in-diffusion from both surfaces possess nonerfc-type U shapes as predicted by the so-called kick-out diffusion model. This model is used to calculate the contribution of self-interstitials to the (uncorrelated) Si self-diffusion coefficient,D I SD =0.064×exp(–4.80 eV/kT)m2 s–1, from the present and previous data on the diffusivity and solubility of Au in Si in the temperature range 1073–1473 K. Inhighly dislocated Si the diffusion of Au is considerably faster than in dislocation-free Si. From the erfc-type penetration profiles found in this case, effective Au diffusion coefficients were deduced and combined with data on the solubility of Au in Si. ThusC i eq D i=0.0064 ×exp(–3.93 eV/kT)m2 s–1 was obtained in the temperature range 1180–1427 K, whereC i eq andD i are the solubility and diffusivity of interstitial Au in Si.  相似文献   

4.
《X射线光谱测定》2003,32(5):363-366
The diffusion of Ag in superconducting YBa2Cu3O7 ceramic was studied over the temperature range 700–850 °C by the energy‐dispersive X‐ray fluorescence (EDXRF) technique. For the excitation of silver atoms, an annular Am‐241 radioisotope source (50 mCi) emitting 59.543 keV photons was used. The temperature dependences of Ag diffusion coefficients in grains (D1) and over the grain boundaries (D2) are described by the equations D1 = 1.4 × 10?2exp[?(1.18 ± 0.10)/kT] and D2 = 3.1 × 10?4exp[?(0.87 ± 0.10)/kT]. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

5.
The mean residence time, τi, of potassium ions on “clean” and oxygenated tungsten has been measured, together with the ionization efficiency, as a function of surface temperature T by using incident K beams of low intensity (109–1012 atoms cm?2 s?1). For T higher than ~900 K the observed τi followed Frenkel's equation τi = τi0exp (QikT) as usual and the agreement of the ionic desorption energy Qi and of the pre-exponential factor τi0 with the corresponding values of previous experiments was quite satisfactory. Below 830 ~ 910 K, where a steep drop of ionization efficiency began to be noticeable, Arrhenius plots of τi deviated considerably from linearity. The apparent increment of the desorption energy was shown to be nearly equal to the decrement of thermionic work function of tungsten as obtained from the ionization efficiency and Saha-Langmuir equation. The increase of surface coverage by potassium was accordingly taken as the main cause of the departure of Arrhenius plots from linearity. Under certain conditions of incident beam intensity and surface temperature τi was observed to make an abrupt change from a higher to a lower value — a difference expressed as 100–140 meV in terms of the difference in ionic desorption energy. This peculiar phenomenon was attributed to the phase change of adsorbed potassium on tungsten.  相似文献   

6.
Proton T1 of FeSiF6·6H2O is proportional to exp (Δ/kT)at liquid helium temperatures, with Δ ≈ D - 3E. We find the crystal field splitting of the Fe2+ ion in this salt to be D = (12.2 ± 1.0) cm-1 using E = 0.54 cm-1.  相似文献   

7.
The (Bi1/9Na2/3)(Mn1/3Nb2/3)O3 ceramics with perovskite structure were sintered. The XRD test proved that the samples are cubic (a?=?3.920?±?0.001?Å). Microstructure and atomic composition were determined with a SEM (JSM-5410) equipped with energy dispersion X-ray analyser (ISIS-300). The fluctuation in the chemical composition was found indicating on local disorder. Broadband dielectric spectroscopy in the range 10?1–3?·?107?Hz was applied within the range of 100–650?K. The real, ?′(f,?T), and imaginary, ?″(f,?T), parts of complex dielectric permitivity characteristics, both in the temperature and frequency domain, show relaxation processes partially covered by electric conductivity. At high temperatures the electric conductivity exhibits a thermally activated behaviour σ(f,?T)?∝?exp(?E a/kT) but the variable range hopping (VRH) dependence σ?∝?exp[?(T 0/T)1/4] is manifested at low temperatures. The derivatives technique in the frequency (??log??/??log?ω) and temperature (??log??/?T) domain enabled various relaxation processes to be distinguished. The data converted to electric modulus representation, M*(f,?T)?=?1/?*, exhibited clearly resolved relaxation peaks. The relaxation times obtained from the peaks position show a slightly non-Arrhenius temperature behaviour with the activation energy varying in 0.4–0.6?eV range and characteristic time of the electric conductivity relaxation of the order of 10?12?s. The relaxation times can be fitted at better accuracy with the VRH dependence where T 0 are of the order of 108?K. It is shown that the low frequency ac-conductivity converges to dc-conductivity and the relation σ(0)?~?ωm?~?τm ?1 typical for the disordered solids applies. The conduction current relaxation relationship behaves in accord with the VRH system: σdc?∝?(T/T 0)q (e 2/kT) ωc, where ωc?=?νph exp[?(T 0/T)1/4] is valid for the locally disordered (Bi1/9Na2/3)(Mn1/3Nb2/3)O3 compound.  相似文献   

8.
Proton spin-lattice relaxation times (T 1) have been measured for triethylene diamine, ethylene oxide, and tetrahydrofuran as clathrate deuterates. The results are interpreted in terms of anisotropic rotation of the guest molecules. Triethylene diamine is thought to be undergoing rotation about its C 3 axis with a correlation time given by τc/s = 4·87 × 10-14 exp (1680 K/T) at temperatures between 120 K and the decomposition point (308 K). Between 77 K and 120 K, T 1 is dominated by conformational distortions of the guest molecule. Ethylene oxide and tetrahydrofuran rotate about at least two axes in the deuterate at rates sufficient to produce some motional narrowing. At high temperatures the relaxation is caused in both cases by rotation about an axis perpendicular to the C 2 axis, and at lower temperatures by rotation about the C 2 axis itself. The correlation times are for ethylene oxide τc/s = 6·76 × 10-14 exp (450 K/T), T < 160 K; and for tetrahydrofuran τc/s = 4·79 × 10-14 exp (470 K/T), T < 140 K.

The free induction decay shapes indicate that, in each case, low frequency motion is occurring about all axes throughout the temperature range studied (77 K to the decomposition temperature in each case). From the lack of an observable signal from the clathrate deuterates of hexamethylene tetramine and dioxan, it is deduced that there is no reorientational motion of these guests at frequencies greater than their rigid-lattice linewidths.  相似文献   

9.
Mutual mass diffusion and thermal diffusion has been investigated in poly(dimethylsiloxane)/ poly(ethylmethylsiloxane) (PDMS/PEMS) polymer blends of equal weight fractions. Molar masses ranged from below 1 to over 20 kg/mol. Both the mutual mass (D) and the thermal diffusion (DT) coefficient contain a thermally activated factor with an activation temperature of 1415 K. The molar mass dependence of DT is due to an end-group effect of the local friction coefficient. The thermal diffusion coefficient in the limit of long chains and infinite temperature is DT0, = - 1.69×10-7cm2(sK)-1. The Soret coefficient ST of blends far enough away from a critical point is proportional to the static structure factor S(q = 0).  相似文献   

10.
ZnO single crystals were doped with Mn and Co by diffusion. In the temperature range from 1400–1600 K the Mn and Co-diffusion-constants were determined:D Mn=3.2 · 10?3 exp (?2.87 eV/kT) cm2 sec?1 andD Co=1·10exp(?3.98 eV/kT) cm2 sec?1. The Mn doped ZnO crystals show a characteristic colour due to an absorption near the intrinsic absorption edge. The corresponding absorption spectra were measured forE⊥c andE∥c. A discussion of different absorption mechanism shows that a charge-transfer transition is responsible for this absorption.  相似文献   

11.
By comparing diffusion coefficientsD of bivalent cations Ba2+, Ca2+, Sr2+ in NaCl crystals it was shown that in the temperature range above 550 °CD (Ba2+)>D (Sr2+)>D (Ca2+) is valid. Temperature dependences of jump frequenciesw 2 of these cations are described byw 2 (Ba2+)=(2·15±0·55) × 1012 × exp {?(0·817±0.007)/kT};w 2 (Sr2+)=(2·9±1·1) × 1012 × exp {?(0·84±0.02)/kT} andw 2 (Ca2+)=(5·5±6·5) × 1010 × exp {?(0·51±0·07)/kT}. It was demonstrated that in NaCl crystals the activation enthalpy and the preexponential factor of the jump frequencyw 2 increase with increasing ionic radius and mass of the bivalent alkaline earth cation.  相似文献   

12.
A pulsed field gradient proton spin-echo NMR self-diffusion study of organic glasses COANP, MBANP, PNP and NPP in their liquid and weakly supercooled states was performed. The NMR phase diagrams, based on the proton NMR transverse relaxation time (T 2) temperature hysteresis data of these materials, clearly give evidence of the onset of a glass phase on cooling the isotropic liquids below their respective melting temperatures. The self-diffusion data exhibit in the supercooled glassy state a non-Arrhenius behaviour and can be described in terms of the Vogel-Fulcher modification of the Arrhenius law,D=D exp{–E a /[k B (T–T VF )]}. The activation energiesE a and Vogel-Fulcher temperaturesT VF are 83.2 meV and 239 K for COANP, 66 meV and 249 K for MBANP, and 85 meV and 245 K for PNP, respectively. The flow viscosity data obtained for COANP in the same temperature region as well conform to the Vogel-Fulcher behaviour, exp{E a () /[k B (T–T VF )]}, withE a () =80.4 meV andT VF =239 K. In case of COANPD was found to increase with decreasing diffusion time in the supercooled (glassy) melt just belowT M whereas no such behaviour was found aboveT M .  相似文献   

13.
In water-molecule-doped barium peroxide and barium oxide, a step increase in the dynamic conductivity to ∼10−3−1 cm−1 was found. The increase is observed when water molecules are present in two nonequivalent states in the lattice, with concentrations of the molecules n t of ≥2.2×1021 cm−3. At n>n t , the conductivity does not depend on the number of molecules in the lattice but is temperature-dependent, obeying the law σ(T = C 1exp(−E 1/kT) + C 2exp(−E 2/kT. The run of the σ(n, T) curve is explained by trapping electrons that result from H2O dissociation and by two sorts of carrier jumps between localized and delocalized states.  相似文献   

14.
The luminescence response of xylene-based liquid scintillator (scintillation grade), xylene + 1 g/l PPO + 0.1 g/l BBOT, has been studied as a function of temperature in the range 200 to 242 K. It has been observed that under gamma-ray excitation the light output increases with the decrease of temperature. The data are well described by the Arrhenius relation,I lowI=I 0 exp(−E/kT), whereI is the count rate at temperatureT.I low is a constant equal to 5600 counts per min, theI 0 factor is 3.3 × 106 counts/min,k is the Boltzmann constant. The activation energyE was found to be equal to 0.20 eV. It is typical for thermally activated diffusion controlled process.  相似文献   

15.
Author index     
The mean adsorption lifetimes of F, Cl and Br on (100) and (111) Mo surfaces have been obtained from the first order desorption kinetics observed in low converage conditions (θ < 10?2 of a monolayer) using a pulsed ionic beam method. The mean adsorption lifetimes τ fit a general expression τ = τ0exp(EkT) in a large temperature range (1700–2400 K) allowing the determination of the binding energies E. The main results of this study are (1) the binding energies decrease from F through Br; (2) the binding energies on both (100) and (111) orientations are similar, E(F)~4.65 eV, E(Cl)~4.15 eV, and E(Br)~3.65 eV. These results are discussed and compared with those previously reported on (100) and (111) Nb and W surfaces. The close binding energies of F, Cl and Br on (100) and (111)?Nb and Mo surfaces suggest that halogens have a different chemisorption behaviour with respect to O and N.  相似文献   

16.
We have measured the nuclear spin lattice relaxation time in liquid indium from 130°C to 300°C to be: 1/T 1=(1.98 × 0.0082T) × 103 sec-1. The relaxation rate consists of two significant parts: (1/T 1) K from the nuclear magnetic hyperfine interaction, and (1/T 1) Q from the nuclear quadrupole interaction. We calculate (1/T 1) K from the the modified Korringa relation using a correction factor of order unity for electron-electron interactions. The hyperfine term is linear in T and accounts for the second term in 1/T 1. Within experimental error the remaining rate, (1/T 1) Q , is temperature independent, and theoretically varies as the product of the square of the electric field gradient, q, and τc, a typical time between field gradient fluctuations. Making use of the x-ray RDF, we construct a simple model for liquid indium and calculate the ionic and electronic contributions, q I and q E, to the electric field gradient, to be q I=1.4 × 1024/cm3 and q E=8.5 × 1024/cm3. The calculation of q E assumes covalent bonding between nearest neighbours. Taking q I and q E to be of opposite sign, we find that the correlation time τc is 1.6 × 10-13 sec. When we further identify τ c with the correlation time for diffusion in a three-dimensional random walk, we are able to calculate the r.m.s. jump distance, Δr D, involved in self-diffusion, Δr D=0.38 Å. This value is consistent with the x-ray peak width of 0.38 Å which we used earlier to calculate the electric field gradient.  相似文献   

17.
The electroluminescent characteristics of an InGaAsSb/GaAlAsSb heterostructure LED emitting at 1.85 μm are studied in the temperature range 20–200°C. It is shown that the emission power exponentially drops as P ≅ 0.4exp(2.05 × 103/T) with a rise in temperature primarily because of an increase in the Auger recombination rate. It is found that band-to-band radiative recombination goes in parallel with recombination through acceptor levels, the latter causing the emission spectrum to broaden. With a rise in temperature, the activation energy of the acceptor levels decreases by the law ΔE≅ 32.9 − 0.075T and the maximum of the LED’s emission spectrum shifts toward the long-wavelength range ( max = 0.693 − 4.497 × 10−4 T). Based on the dependence E g = max − 0.5kT and experimental data, an expression is derived for the temperature variation of the bandgap in the In0.055Ga0.945AsSb active area, E g ≅ 0.817 − 4.951 × 10−4 T, in the range 290 K < T < 495 K. The resistance of the heterostructure decreases exponentially with rising temperature as R 0 ≅ 5.52 × 10−2exp(0.672/2kT), while cutoff voltage U cut characterizing the barrier height of a p−n junction decreases linearly with increasing temperature (U cut = −1.59T + 534). It is found that the current through the heterostructure is due to the generation-recombination mechanism throughout the temperature interval.  相似文献   

18.
In a dc glow discharge in oxygen, the concentrations of minor components of O2(a1Δg), O2(b1 Σg), O3, O(1D), as well as nagative ions and electrons have been measured. Balance equations have been derived which describe satisfactorily the stationary concentrations of these components as functions of gas pressure and discharge current. For the first time, the rate constants of important aeronomical reactions (a) O? + O2(a1Δg) → O3 + e, (b) O2? + O2(a1Δg) → 2O2 + e and (c) e + O3 → O2? +O have been measured as functions of gas temperature T and mean energies of ions Ei and electron E6: Ka = (2.5 ± 0.5) · 10?9 · (T/300)4 ± 0.4· (Ei/0.04)?2.6 ± 0.4 cm3/s for T = 385?605 K and Ei = 0.10 ? 0.66 eV; Kb = (1.0 ± 0.3) · 10?10 · (T/300)?2 ± 0.5 · (Ei/0.04)0.23 ± 0.05 cm3/s for T = 330?605 K and Ei = 0.09 + 1.5 eV; Kc for Ee = 0.8÷5 eV.  相似文献   

19.
The temperature dependence of a zero-bias anomaly in the tunneling conductance of an Al/δ-GaAs tunneling structure with a two-dimensional electron density in the δ-layer of 3.5 × 1012 cm?2 has been investigated. It has been shown that the respective drop Δρ(?, T) in the tunneling density of states ρ near the Fermi level E F of the two-dimensional electron system depends logarithmically on the energy ? within the range of 2.7kT < |?| < ?/τ, where ? is measured with respect to E F and τ is the momentum relaxation time of two-dimensional electrons. It has been found that the drop depth Δρ(0, T)/ρ is also proportional to ln(kT/?0) in the temperature range T = 0.1–20 K and saturates below 0.1 K.  相似文献   

20.
Using a surface ionisation ion microscope the desorption parameters and the diffusion constant of potassium were measured on stepped W(100) surfaces. The activation energy of ionic desorption as well as the corresponding prefactor do not depend on the step density; the mean adsorption lifetime τ can be expressed as τ=1.6×10?14s exp(2.44 eV/kT).Whereas the surface diffusion of potassium on “flat” W(100) and on W(S)-[9(100)×(110)] was found to be isotropic, on W(S)- [5(100)×(110)] and W(S)-[3(100)×(110)] it occurs preferentially parallel to the step direction. The diffusion constant D for this direction has roughly the same value for all investigated surfaces: D=7.8×10?2 cm2s?1 exp(?0.42 eV/kT). For the direction perpendicular to the steps D⊥ depends on the step density, whereby the activation energy as well as the prefactor increase with increasing step density.  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