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.
The linearized Peierls equation for the phonon densityN (k λ,r t) is solved by replacing the collision operator in the subspace orthogonal to the collision invariants byk-dependent relaxation rates. For the normal process relaxation time the behaviorτ N (k λ)∝|k|?p for smallk is assumed. Taking into account thisk-dependence ofτ N explicitly and avoiding an expansion with respect toΩτ N () before performing the necessary integration overk yields new, non-analytic, terms in the hydrodynamic equations describing second sound and Poiseuille flow. It is shown that this may lead to a temperature dependence of second sound damping and thermal conductivity in the Poiseuille flow region differing from the usual theoretical predictions and in better agreement with experiments.  相似文献   

2.
Microstructural characteristics and amplitude dependences of the Young modulus E and of internal friction (logarithmic decrement δ) of bio-carbon matrices prepared from beech tree wood at different carbonization temperatures T carb ranging from 600 to 1600°C have been studied. The dependences E(T carb) and δ(T carb) thus obtained revealed two linear regions of increase of the Young modulus and of decrease of the decrement with increasing carbonization temperature, namely, ΔEAΔT carb and Δδ ~ BΔT carb, with A ≈ 13.4 MPa/K and B ≈ ?2.2 × 10?6 K?1 for T carb < 1000°C and A ≈ 2.5 MPa/K and B ≈ ?3.0 × 10?7 K?1 for T carb > 1000°C. The transition observed in the behavior of E(T carb) and δ(T carb) at T carb = 900–1000°C can be assigned to a change of sample microstructure, more specifically, a change in the ratio of the fractions of the amorphous matrix and of the nanocrystalline phase. For T carb < 1000°C, the elastic properties are governed primarily by the amorphous matrix, whereas for T carb > 1000°C the nanocrystalline phase plays the dominant part. The structurally induced transition in the behavior of the elastic and microplastic characteristics at a temperature close to 1000°C correlates with the variation of the physical properties, such as electrical conductivity, thermal conductivity, and thermopower, reported in the literature.  相似文献   

3.
The dependence on temperature of the layer magnetization of a Heisenberg ferromagnetic ultrathin film in presence of magnetocrystalline single-ion anisotropy was theoretically investigated in the framework of a Green's function approach using the random phase approximation (RPA). The effect of surface orientation and of film thickness N on the Curie temperature TC was carefully investigated in the case of face centered cubic (FCC) films: the steepest increase of TC(N) was found in the case of the FCC(1 1 1) orientation and the smoothest in the FCC(1 1 0) one. Our results for TC(N) were successfully fitted by a finite-size scaling relation [TC(∞)−TC(N)]/TC(N)=(N/N0)λ, giving a shift exponent λ≃1.5, irrespectively of the surface orientation. Finally, the temperature evolution of the magnetization profile was analyzed, as well as its limiting shape at TC.  相似文献   

4.
The thermal diffusivity, a, of aluminium has been measured at pressures up to 2.5 GPa at room temperature, and from these results the pressure dependence of the thermal conductivity, λ, has been calculated. Both quantities increase with pressure. The increase in a amounts to 4.6% to 1 GPa and 10.4% to 2.5 GPa. The initial pressure coefficient of the electronic thermal conductivity λe is found to be [λe]-1e/?P = 3.7 × 10-2GPa-1, which agrees very well with a recent theoretical calculation.  相似文献   

5.
Thermal conductivity (λ) of nanocrystalline La0.67(CaxSr1−x)0.33MnO3 (x=0, 0.5, 1) and La0.6Y0.07Ca0.33MnO3 pellets prepared by a novel ‘pyrophoric’ method have been studied between the temperature range 10 and 300 K. Our data show that the magnitude of thermal conductivity is strongly influenced by the ion substitutions at La-site. The analysis of the thermal conductivity data indicates that the thermal transport is governed largely by phonons scattering in these systems and the electronic contribution is as small as 0.2-1% of total thermal conductivity (λtotal). At low temperatures (<90 K) 2D like lattice defects contribute to the phonon scattering dominantly and its strength increases with increasing Sr content and also with partial substitution of La by Y. Depending upon the composition of the samples, the magnon thermal conductivity contributes 2-15% of λtotal close to TC. In the paramagnetic regime the unusual increase in λtotal keeps signature of large dynamic lattice distortion.  相似文献   

6.
7.
The electronic spectrum of a doped semiconductor described by the Anderson-Holstein impurity model and its conductivity derived from the Kubo linear response theory are calculated. Two characteristic temperatures depending on the doping level x are found in the phase diagram, T PG and T λ(x). The pseudogap that opens in the single-particle spectrum at low doping levels and temperatures closes at the lower one, T PG. The pseudogap state of an insulator is attributed to spin fluctuations in a doped compound. At the higher characteristic temperature T λ(x),, spin fluctuations vanish and the doped compound becomes a paramagnetic poor metal. Two distinct metal-insulator crossovers between semiconductor-like and metallic temperature dependence of resistivity are found. An insulator-to-poor-metal transition occurs at T *(x) ≈ T λ(x). A poor-metal-to-insulator transition at a lower temperature is attributed to the temperature dependence of density of states in the pseudogap. It is shown that both transitions are observed in La2?x SrxCUO4.  相似文献   

8.
The thermal conductivity κ (within the range 4–300 K) and electrical conductivity σ (from 80 to 300 K) of polycrystalline Sm3S4 with the lattice parameter a=8.505 Å (with a slight off-stoichiometry toward Sm2S3) are measured. For T>95 K, charge transfer is shown to occur, as in stoichiometric Sm3S4 samples, by the hopping mechanism (σ ~ exp(?ΔE/kT) with ΔE ~ 0.13 eV). At low temperatures [up to the maximum in the lattice thermal conductivity κph(T)], κphT 2.6; in the range 20–50 K, κphT ?1.2; and for T>95 K, where the hopping charge-transfer mechanism sets in, κphT ?0.3 and a noticeable residual thermal resistivity is observed. It is concluded that in compounds with inhomogeneous intermediate rare-earthion valence, to which Sm3S4 belongs, electron hopping from Sm2+ (ion with a larger radius) to Sm3+ (ion with a smaller radius) and back generates local stresses in the crystal lattice which bring about a change in the thermal conductivity scaling of κph from T ?1.2 to T ?0.3 and the formation of an appreciable residual thermal resistivity.  相似文献   

9.
The thermal conductivity of LaAl2 and of two dilute (La, Ce)Al2 alloys was measured in the normal state between 0.4 and 8 K. From the lattice conductivity of LaAl2 a high dislocation densityN d caused by the arc melting process can be inferred. After annealingN d is reduced by an order of magnitude. For the (La, Ce)Al2 samples minima are observed at 5 K in theW e ·T vs.T curves (W e =electronic thermal resistivity). Below 1 K the quantityW e ·T is linear in (— lnT). The electronic Lorenz ratioL e (T)=ρ(T)/W e (T) ·T shows a maximum at 2 K with a value 23% aboveL e (0). It is for the first time that this Kondo anomaly is established in its full temperature dependence.  相似文献   

10.
We consider the thermal conductivity of steeply repulsive inverse power fluids (SRP) in which the particles interact with a pair potential, φ(r) = ε(σ/r)n. The time correlation function for the heat flux, Cλ(t), and the time average, Cλ(0) are calculated numerically by molecular dynamics simulations, and accurate expressions for these are also derived for the SRP fluid. We show, by molecular dynamics simulations, that close to the hard-sphere limit this time correlation function has the same analytic form as for the shear and pressure correlation functions for the shear and bulk viscosity, i.e. Cλ(t)/Cλ(0) = 1 ?T* (nt*)2 + 0((nt*)4), where T* = k B T/ε, is the reduced temperature, k B is Boltzmann's constant and t* = (ε/σ2)1/2 t is the reduced time. The thermal conductivity for the limiting case of hard spheres is numerically very close to that given by the traditional Enskog relation. At low densities the normalized relaxation times are typically largest for the thermal conductivity, followed by shear and then bulk viscosity. Close to the maximum fluid density, the latter two increase rapidly with density (especially for the shear) but continue a monotonic decline for the thermal conductivity. This reflects the relative insensitivity of the thermal conductivity to the approach to the fluid-solid phase boundary.  相似文献   

11.
The thermal conductivity of a trapped dipolar Bose condensed gas is calculated as a function of temperature in the framework of linear response theory. The contributions of the interactions between condensed and noncondensed atoms and between noncondensed atoms in the presence of both contact and dipole-dipole interactions are taken into account to the thermal relaxation time, by evaluating the self-energies of the system in the Beliaev approximation. We will show that above the Bose-Einstein condensation temperature (T?>?T BEC ) in the absence of dipole-dipole interaction, the temperature dependence of the thermal conductivity reduces to that of an ideal Bose gas. In a trapped Bose-condensed gas for temperature interval k B T?<<?n 0 g B E p ?<<?k B T (n 0 is the condensed density and g B is the strength of the contact interaction), the relaxation rates due to dipolar and contact interactions between condensed and noncondensed atoms change as \( {\tau}_{dd12}^{-1}\propto {e}^{-E/{k}_BT} \) and τ c12?∝?T ?5, respectively, and the contact interaction plays the dominant role in the temperature dependence of the thermal conductivity, which leads to the T ?3 behavior of the thermal conductivity. In the low-temperature limit, k B T?<<?n 0 g B , E p ?>>?k B T, since the relaxation rate \( {\tau}_{c12}^{-1} \) is independent of temperature and the relaxation rate due to dipolar interaction goes to zero exponentially, the T 2 temperature behavior for the thermal conductivity comes from the thermal mean velocity of the particles. We will also show that in the high-temperature limit (k B T?>?n 0 g B ) and low momenta, the relaxation rates \( {\tau}_{c12}^{-1} \) and \( {\tau}_{dd12}^{-1} \) change linearly with temperature for both dipolar and contact interactions and the thermal conductivity scales linearly with temperature.  相似文献   

12.
Measurements of the Hall coefficient R H (T, B) of Si:P with P concentration N between 3.54 and 7.0·1018 cm?3 are reported for the temperature range 0.04 K ≤ T ≤ 4K and in magnetic fields up to 7 T. Even far above the metal-insulator transition (MIT), a sign change of the temperature coefficient similar to the behavior of the conductivity σ(T) in moderate fields is not observed in R H (T). Field and temperature dependence of R H both increase as the MIT (at the critical concentration N c = 3.52 · 1018 cm?3) is approached. A careful extrapolation to T → 0 and B → 0 indicates that R H ?1 scales to zero as R H ?1 ~| N ? N c μH with μ H = (0.44 ± 0.04) in agreement with previous results.  相似文献   

13.
The anomalous reduced sound velocity of multiferroic TbMn2O5 (TMO) has been studied using Green's function technique. To achieve this aim, the anharmonic phonon-phonon interaction and the spin-phonon interaction were used. It was shown that the reduced velocity of sound of TMO exhibits a kink at the ferroelectric phase transition temperature TC. This can be explained as an effect of vanishing ferroelectric ordering above TC. It was found that the reduced sound velocity increases with increasing V(3) (the third-order atomic force constants of the anharmonic phonons) in the interval T?<?TC, whereas the reduced sound velocity remains unchanged in the interval TC?<?T?<?TN. It was also found that the reduced sound velocity increases with the increase of V(4) (the fourth-order atomic force constants of the anharmonic phonons) in the interval T?<?TN. In addition, the ferroelectric phase transition temperature TC decreases when V(4) increases in the interval T?<?TN. Those theoretical results are in agreement with the experimental data.  相似文献   

14.
The temperature dependence of the non-linear susceptibility ≈2(T) of random anisotropy magnets in the Ising limit (speromagnets) is calculated for temperatures above the freezing temperature Tf within the framework of the correlated molecular field theory. For the effective susceptibility exponent λs(T) = (T?Tf)≈2d-1≈2/dT a non-monotonic temperature dependence is found as for the case of spin glasses. This must be taken into account in order to obtain reliable values for the critical susceptibility exponent from experimental data.  相似文献   

15.
A generalization of the Vollhardt-Wölfle localization theory is proposed to make it possible to study the spatial-temporal dispersion of the kinetic coefficients of a d-dimensional disordered system in the low-frequency, long-wavelength range (ω?F and q?k F ). It is shown that the critical behavior of the generalized diffusion coefficient D(q,ω) near the Anderson transition agrees with the general Berezinskii-Gor’kov localization criterion. More precisely, on the metallic side of the transition the static diffusion coefficient D(q,0) vanishes at a mobility threshold λ c common for all q: D(q, 0)∝t=(λ c ?λ)/λ c →0, where λ=1/(2π?F τ) is a dimensionless coupling constant. On the insulator side, q≠0 D(q,ω)∝? as ω→0 for all finite q. Within these limits, the scale of the spatial dispersion of D(q,ω) decreases in proportion to t in the metallic phase and in proportion to ωξ 2, where ξ is the localization length, in the insulator phase until it reaches its lower limit ~λ F. The suppression of the spatial dispersion of D(q,ω) near the Anderson transition up to the atomic scale confirms the asymptotic validity of the Vollhardt-Wölfle approximation: D(q,ω)?D(ω) as |t|→0 and ω→0. By contrast, the scale of the spatial dispersion of the electrical conductivity in the insulator phase is of order of the localization length and diverges in proportion to |t|?v as |t|→0.  相似文献   

16.
The three-dimensional ordering temperatures of the quasi-one-dimensional antiferromagnet (CH3)4NMnCl3 containing 0, 0.13, 0.25 and 0.48 at% Cu2+ ions were studied as a function of an external magnetic field up to 60 kOe by means of proton NMR. A strong increase of TN(H) with H was found even for the impure system. The TN versus H curve for the pure system cam be explained by introducing a simple form for the effect of the field on the correlation length, ξ(T, H) = ξ(T) × [1+(λJ/kT)H2], while that for the impure system is unlike the one for the pure system.  相似文献   

17.
The thermal conductivity k and resistivity ρ of biocarbon matrices, prepared by carbonizing medium-density fiberboard at T carb = 850 and 1500°C in the presence of a Ni-based catalyst (samples MDF-C( Ni)) and without a catalyst (samples MDF-C), have been measured for the first time in the temperature range of 5–300 K. X-ray diffraction analysis has revealed that the bulk graphite phase arises only at T carb = 1500°C. It has been shown that the temperature dependences of the thermal conductivity of samples MDFC- 850 and MDF-C-850(Ni) in the range of 80–300 K are to each other and follow the law of k(T) ~ T 1.65, but the use of the Ni-catalyst leads to an increase in the thermal conductivity by a factor of approximately 1.5, due to the formation of a greater fraction of the nanocrystalline phase in the presence of the Ni-catalyst at T carb = 850°C. In biocarbon MDF-C-1500 prepared without a catalyst, the dependence is k(T) ~ T 1.65, and it is controlled by the nanocrystalline phase. In MDF-C-1500(Ni), the bulk graphite phase formed increases the thermal conductivity by a factor of 1.5–2 compared to the thermal conductivity of MDF-C-1500 in the entire temperature range of 5–300 K; k(T = 300 K) reaches the values of ~10 W m–1 K–1, characteristic of biocarbon obtained without a catalyst only at high temperatures of T carb = 2400°C. It has been shown that MDF-C-1500(Ni) in the temperature range of 40?300 K is characterized by the dependence, k(T) ~ T 1.3, which can be described in terms of the model of partially graphitized biocarbon as a composite of an amorphous matrix with spherical inclusions of the graphite phase.  相似文献   

18.
We discuss the finite-temperature generalization of time-dependent density functional theory (TDDFT). The theory is directly analogous to that at temperature T = 0. For example, the finite-T TDDFT exchange-correlation kernel fxc(T, n) in the local density approximation can again be expressed as a density derivative of the exchange correlation potential fxc(T, n) = [?vxc(T, n)∕?n]δ(r ? r), where n = NV is the electron number density. An approximation for the kernel fxc(T, n) is obtained from the finite-T generalization of the retarded cumulant expansion applied to the homogeneous electron gas. Results for fxc and the loss function are presented for a wide range of temperatures and densities including the warm dense matter regime, where TTF, the electron degeneracy temperature. The theory also permits a physical interpretation of the exchange and correlation contributions to the theory.  相似文献   

19.
The dependence of the thermal conductivity of indium antimonide on temperature (in the range 300–450 K) and hydrostatic pressure (up to 0.4 GPa) has been investigated. It is shown that the phonon thermal conductivity λph obeys the law T ?n (n ≥ 1). Hydrostatic pressure affects the magnitude and temperature dependence of the thermal conductivity of InSb: with an increase in pressure, the thermal conductivity increases, while the parameter n in the dependence λphT ?n decreases.  相似文献   

20.
J. Kestin  S.T. Ro  W.A. Wakeham 《Physica A》1983,119(3):615-638
The paper presents new, absolute measurements of the viscosity and thermal conductivity of mixtures of hydrogen with carbon monoxide, carbon dioxide and methane near room temperature. The viscosity measurements have been performed at a pressure of 0.1 MPa and have an estimated accuracy of ±0.1%. The data have been employed to evaluate quantities characteristic of the unlike interaction, including the binary diffusion coefficient. These results allow the thermal conductivity to be calculated with the aid of the appropriate kinetic-theory expressions. The calculations have been performed for both the zero-density limit and over the range of pressures encompassed by the measurements (0.8–12 MPa). It has not proved possible to represent the experimental thermal conductivity data within their uncertainty of ±0.3% by means of the available kinetic-theory equations. This finding is attributed to the fact that the equations employed are not as accurate as the experimental data for systems in which the mass ratio of the species differs substantially from unity.  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