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.
The solubility of Nd at the Ba sites and the superconductivity of YBa2?x Nd x Cu3O y were investigated by X-ray powder diffraction and measurements of the electrical resistance and ac susceptibility. The single Re123 phase was obtained for x≤0.30. The onset transition temperature $T_{\text{c}}^{{\text{on}}}$ is insensitive to the Nd content x in the region of x≤0.40. All are higher than 95 K. The zero resistance transition temperatures $T_{\text{c}}^{{\text{zero}}}$ , however, exhibits two-step variation with the increase of x. For x≤0.25, $T_{\text{c}}^{{\text{zero}}}$ are all above 92 K. The highest $T_{\text{c}}^{{\text{zero}}}$ of 94 K was obtained for x=0.25. For x≥0.3 $T_{\text{c}}^{{\text{zero}}}$ drops sharply to about 84 K. Finally $T_{\text{c}}^{{\text{zero}}}$ falls to 30 K and $T_{\text{c}}^{{\text{zero}}}$ is below 10 K for x=0.5. The two-step variation of T c might be an indication of the existence of two trap levels for holes.  相似文献   

2.
This paper defines an effective microwave surface resistance $R_{\text{s}}^{{\text{eff}}}$ for the nonuniform distribution of microwave surface resistance R s in the strip of a microstrip. It is proved that $R_{\text{s}}^{{\text{eff}}}$ is equivalent to the expression of R s used in experiments, and that the $R_{\text{s}}^{{\text{eff}}}$ is dominated by the edge part, i.e., the area of width λ2/2t from the strip edge, where λ is the magnetic penetration depth and t is the film thickness. Under the assumption that $R_s \sim \left( {H_{{\text{rf}}}^y } \right)^n$ where $H_{{\text{rf}}}^y$ is the component of rf magnetic field along the film thickness and n is an integer, the ratio of the contributions of the edge part and the rest of the strip to $R_s^{{\text{eff}}}$ is calculated by using an approximate analytical expression of the surface current density distribution J s in the strip and $H_{{\text{rf}}}^y$ calculated by the London equation. The effect of film's edge on R s was studied using a microstrip resonator. It is found that the perfectness of the edge could affect the magnitude of the power dependence of R s significantly, which agreed with our analysis.  相似文献   

3.
Electrical conductivity (σ) of “pure” and ZnO doped SnO2 has been measured at different temperatures and oxygen partial pressures ( \(p_{{\text{O}}_{\text{2}} } \) )- From the variation of electrical conductivity of these materials three partial pressure ranges have been identifieD. In the high partial pressure rangeσ increases with decreasing \(p_{{\text{O}}_{\text{2}} } \) followed by a \(p_{{\text{O}}_{\text{2}} } \) independent region at lower \(p_{{\text{O}}_{\text{2}} } \) ´s and finally increases once again with a further decrease of \(p_{{\text{O}}_{\text{2}} } \) . These variations have been explained on the basis of an anti-Frenkel type defect structure and an interstitial solid solution of ZnO in SnO2. The activation energy for the conduction process has been estimated and the values are found to differ in two different temperature ranges. In the low temperature range the conductivity is attributed mainly to the chemisorption of oxygen on the surface of the specimen.  相似文献   

4.
We present a novel approach to the analysis of the normal state in-plane $\sigma _{ab} $ and out-of-plane σc conductivities of anisotropic layered crystals such as oxygen deficient YBa 2 Cu 3 O x . It can be shown that the resistive anisotropy is determined by the ratio of the phase coherence lengths in the respective directions; i.e., $\sigma _{ab} /\sigma _c = \ell _{ab}^2 /\ell _c^2 $ . From the idea that at all doping levels and temperatures T the out-of-plane transport in these crystals is incoherent, follows that $\ell _c $ is T-independent, equal to the spacing $\ell _0 $ between the neighboring bilayers. Thus, the T-dependence of $\ell _{ab} $ is given by the measured anisotropy, and $\sigma _{ab} (\ell _{ab} )$ dependence is obtained by plotting $\sigma _{ab} {\text{ }}vs{\text{ }}\ell = {\text{ (}}\sigma _{ab} /\sigma _c )^{1/2} \ell _0 $ .The analysis of several single crystals of YBa 2 Cu 3 O x (6.35 < x < 6.93) shows that for all of them $\sigma _{ab} (\ell ) $ is described by a universal dependence $\sigma _{ab} /\overline \sigma = f(\ell /\overline \ell ) $ with doping dependent parameters $\overline \sigma {\text{ }}and{\text{ }}\overline \ell $ .  相似文献   

5.
The effects of bulk porosity on the thermoelectric properties of porous n-type ${\text{Fe}}_{{\text{0}}{\text{.94}}} {\text{Co}}_{{\text{0}}{\text{.06}}} {\text{Si}}_{\text{2}} $ compounds prepared by pressureless sintering were examined. A small amount of metallic phase ?-FeSi remained after annealing at 800 °C for 100 h. As the sintering temperature increased from 1150 to 1175 °C, the phase transition to β-FeSi2 during annealing occurred more rapidly. The porous specimen, sintered at 1150 °C for 2 h with coarse powders (<45 μm) and then annealed at 800 °C for 100 h, showed the highest Seebeck coefficient of $ - 363_{\mu } {VK}^{ - {1}} $ at 400 °C and the highest power factor of ${\text{1}}{\text{.57}} \times 10^{ - 3} {\text{Wm}}^{ - 1} {\text{K}}^{ - 2} $ at 400 °C.  相似文献   

6.
The effects of fluorine co-doping by means of a post-thermal annealing process of iron-doped BST thin films in a fluorine-containing atmosphere have been investigated. XPS and ToF-SIMS sputter depth profiling verified a homogeneous fluorine distribution in the thin films. By employing EPR, it was shown that singly charged ( $ {\text{Fe}}_{\text{Ti}}^{\prime } $ $ {\text{V}}_{\text{O}}^{ \cdot \cdot } $ )· defect complexes, as well as ‘isolated’ $ {\text{Fe}}_{\text{Ti}}^{\prime } $ centres with a distribution of $ {\text{F}}_{\text{O}}^{ \cdot } $ sites in remote coordination spheres exist in the fluorinated films. Tunability enhancement due to fluorine co-doping as well as a Q-factor enhancement due to iron doping is demonstrated.  相似文献   

7.
Inverse gas chromatography at infinite dilution was applied to determine the thermodynamic interactions of poly(l-lactide) (PLLA) and the composite of biphasic calcium phosphate and PLLA (BCP/PLLA). The specific retention volumes, $ V_{\text{g}}^{0} $ , of 11 organic compounds of different chemical nature and polarity (non-polar, donor or acceptor) were determined in the temperature range of 308–378 K for PLLA and 308–398 K for BCP/PLLA. The weight fraction activity coefficients of test sorbates, $ \Omega_{1}^{\infty } $ , and the Flory–Huggins interaction parameters, $ \chi_{12}^{\infty } $ , were estimated and discussed in terms of interactions of the sorbates with PLLA and BCP/PLLA. Also, the partial molar free energy, $ \Delta G_{1}^{\infty } $ , the partial molar heat of mixing, $ \Delta H_{1}^{\infty } $ , the sorption molar free energy, $ \Delta G_{1}^{\text{S}} $ , the sorption enthalpy, $ \Delta H_{1}^{\text{S}} $ , and the sorption entropy, $ \Delta S_{1}^{\text{S}} $ , were analyzed. A different chromatographic behavior of the two investigated samples, PLLA and BCP/PLLA, was observed. The values of $ \Omega_{1}^{\infty } $ indicated n-alkanes, diethyl ether, tetrahydrofurane (THF), cyclohexane, benzene, dioxane (except for 338 K), and ethyl acetate (EtAc) (except for 338 K) as non-solvents, and chloroform (CHCl3) as good solvent (except for 378 K) for PLLA. For BCP/PLLA, CHCl3, EtAc (for 378 K), dioxane (except for 378 K), and THF were indicated as good solvents.  相似文献   

8.
We report the impact of hydrostatic pressure on the superconductivity and normal-state resistivity of FeTe0.5Se0.5 superconductor. At the ambient pressure, the FeTe0.5Se0.5 compound shows the superconducting transition temperature \(T_{\mathrm {c}}^{\text {onset}} \) at above 13 K and \(T_{\mathrm {c}}^{\rho =0} \) at 11.5 K. We measure pressure-dependent resistivity from 250 to 5 K, which shows that the normal-state resistivity increases initially for the applied pressures of up to 0.55 GPa, and then the same is decreased monotonically with increasing pressure of up to 1.97 GPa. On the other hand, the superconducting transition temperatures ( \(T_{\mathrm {c}}^{\text {onset}} \) and \(T_{\mathrm {c}}^{\rho =0} )\) increase monotonically with increasing pressure. Namely the \(T_{\mathrm {c}}^{\text {onset}} \) increases from 13 to 25 K and \(T_{\mathrm {c}}^{\rho =0} \) from 11.5 to 20 K for the pressure range of 0–1.97 GPa. Our results suggest that superconductivity in this class of Fe-based compounds is very sensitive to pressure as the estimated pressure coefficient d T c(onset)/dP is ~5.8 K/GPa. It may be suggested that the FeTe0.5Se0.5 superconductor is a strong electron-correlated system. The enhancement of T c with applying pressure is mainly attributed to an increase of charge carriers at the Fermi surface.  相似文献   

9.
10.
A model based on the effect of a modest applied dc electric field on grain growth is proposed for the contribution of the space charge to the grain boundary (GB) energy in 3 mol % yttria-stabilized zirconia (3 YSZ). The model considers that the total GB energy \( \gamma_{\text{b}}^0 \) (the capillary driving force for grain growth) consists of three major components: (a) \( \gamma_{\text{b}}^\varSigma \) due to the misorientation between neighboring grains, (b) \( \gamma_{\text{b}}^{\text{s}} \) due to the size misfit between the segregated solute and the solvent cations, and (c) \( \gamma_{\text{b}}^{\text{e}} \) the electrostatic (space charge) component, which results from the segregation of the aliovalent yttrium ions to the grain boundaries. The former two components combined comprise 40 % of the total GB energy in 3 YSZ and the electrostatic component 60 %. Based on the model, the calculated magnitudes of the three components were in qualitative accord with theoretical considerations and with values reported in the literature. A reduction in \( \gamma_{\text{b}}^{\text{e}} , \) and in turn in \( \gamma_{\text{b}}^0 , \) results from the bias exerted by the applied field on the space charge potential that occurs with the segregation of the yttrium ions to the grain boundaries. The observed reduction of grain growth in 3 YSZ by an applied electric field is attributed mainly to the reduction in \( \gamma_{\text{b}}^{\text{e}} \) by the field.  相似文献   

11.
First-principles cluster calculations are reported of the local electronic structure of the three compounds: La2CuO4, Sr2CuO2Cl2, and Sr2CuO2F2. The copper ${\text{3d}}_{x^2 - y^2 } $ and the planar oxygen 2p σ atomic orbitals exhibit a similar degree of covalency. The out-of-plane orbitals, however, are quite different with the ${\text{3}}d_{3z^2 - r^2 } $ atomic orbital lowered significantly in energy for chlorine and fluorine apical positions.  相似文献   

12.
An ion interacting with quasiparticles in liquid3He is treated theoretically by summing most divergent terms in perturbation series of the self-energy of the ion and the vertex part. The ion Green's function is renormalized by a factorZ(T), and the vertex part byZ(T) ?1, where \({\text{Z(T)}} = {\text{(T/T}}_F {\text{)}}^{{\text{2V}}_{{\text{0}}^{\rho ^{\text{2}} } }^{\text{2}} } \) , forT 0?T?T F. Here,T 0=(m/M)T F, withm the3He mass andM the ion mass, and \({\text{V}}_{0^\rho } \) is the strength of the interaction. The factor explains the weak temperature dependence of the mobility around the minimum atT 0; we also discuss its effect on the behavior of the mobility in3He-B nearT c.  相似文献   

13.
Polycrystalline samples La2 ? x Sr x CuO4 ? δ (0.06 ≤ x ≤ 0.5) and La1.85 ? x Sr0.15 + x Cu1 ? x Co x O4 ? δ (0 ≤ x ≤ 1) were synthesized by a conventional solid state reaction method. The structure, phonon vibration, and conduction were investigated by means of XRD, infrared (IR) spectra, and resistance. It is found that the increase of itinerant hole carriers could mask the in-plane ${\text{Cu}}{\kern 1pt} - {\kern 1pt} {\text{O(1)}}$ stretching vibration mode (689 cm?1). The softening of the ${\text{Cu}}{\kern 1pt} - {\kern 1pt} {\text{O(1)}}$ phonon vibration mode gives an index of the weakening of ${\text{Cu}}{\kern 1pt} - {\kern 1pt} {\text{O}}$ hybridization. The distortion of CuO6 octahedron and the microstructural inhomogeneity induced by Co doping lead to the widening of the IR absorption peaks.  相似文献   

14.
A moisture content effect on fracture characterisation of Pinus pinaster under mode I is addressed. The double cantilever beam test is selected for mode I loading, based on specimens scaled down to the growth ring level. Specimens are stabilised using aqueous solutions at several equilibrium moisture contents ( \( M_{\text{e}} \) ) ranging from 0 % to about 13 %. The strain energy release rate ( \( G_{\text{I}} \) ) is evaluated by applying the compliance-based beam method, from which the Resistance-curve is determined directly from load–displacement (Pδ) data without crack length measurement. The crack tip opening displacement in mode I ( \( w_{\text{I}} \) ) is determined by post-processing the displacements at the initial crack tip determined from digital image correlation. \( G_{\text{I}} \) and \( w_{\text{I}} \) are then combined for the direct identification of the cohesive law, \( \sigma_{\text{I}} = {\text{f}}(w_{\text{I}} ) \) , which is assumed in cohesive zone modelling. Experimentally, crack propagation consistently occurs in the earlywood layer. The increase of \( M_{\text{e}} \) does not show any influence on the initiation stage of the fracture process zone. However, a statistical correlation exists for the critical ( \( G_{\text{I,max}} \) ) and at maximum load ( \( G_{{{\text{I,}}P_{ \hbox{max} } }} \) ) values of \( G_{\text{I}} \) with regard to \( M_{\text{e}} \) . Consistently, the area under the identified cohesive curve increases with \( M_{\text{e}} \) , although high scatter and low correlation between maximum cohesive strength ( \( \sigma_{\text{Iu}} \) ) and \( M_{\text{e}} \) are observed. The methodology is also validated using finite element analysis including cohesive elements and taking into account the growth rings heterogeneity. The numerical results show that the identification of the cohesive law is insensitive to the variability of the growth ring structure observed experimentally.  相似文献   

15.
The study of the propagation of laser-generated Lamb waves in viscoelastic plates, after taking into account the viscoelastic and thermophysical properties, is reported. Based on the thermoviscoelastic theory, a finite-element model for the laser-generated Lamb waves is developed in the frequency domain, and the temporal temperature and displacement waveforms are obtained by using inverse fast Fourier transforms. The influences of each elastic and viscous modulus $(C_{11}^{\prime },\; C_{22}^{\prime },\; C_{12}^{\prime },\; C_{66}^{\prime },\; C_{11}^{{\prime }{\prime }},\; C_{22}^{{\prime }{\prime }},\; C_{12}^{{\prime }{\prime }},$ and $C_{66}^{{\prime }{\prime }})$ on the propagation of the laser-generated Lamb waves (mainly for the $S_{0}$ and $A_{0}$ modes) are investigated in detail. The numerical results show that the attenuation of the $S_{0}$ mode is mainly determined by $C_{22}^{\prime },\; C_{12}^{\prime },\; C_{11}^{{\prime }{\prime }}$ , and $C_{12}^{{\prime }{\prime }}$ , and is independent of $C_{66}^{\prime }$ and $C_{66}^{{\prime }{\prime }}$ . In addition, the attenuation of the $A_{0}$ mode is more influenced by $C_{11}^{\prime },\; C_{22}^{\prime },\; C_{11}^{{\prime }{\prime }},$ and $C_{12}^{{\prime }{\prime }}$ . The finite-element method in the frequency domain provides a useful tool to investigate the mechanical properties of the viscoelastic plates.  相似文献   

16.
The generalized fracture mechanics approach is applied to two ductile steels, namely mild steel and 18/8 stainless steel in plane stress. The theory defines a fracture parameter \(\mathcal{T}\) , which is a truly plastic analogue of theJ contour integral and, for an edge crack specimen, is given by $$\mathcal{T} = k_1 ( \in _0 )cW_{0_c } $$ wherek 1 is an explicit function,c is the crack length andε 0, W0c are respectively the strain and input energy density at fracture, remote from the crack. The functionk 1(ε o) is derived experimentally and the constancy of \(\mathcal{T}\) with respect to crack length and applied load is demonstrated. The variation of \(\mathcal{T}\) with crack extension during slow growth is investigated, as is the rate dependence of \(\mathcal{T}\) in mild steel.  相似文献   

17.
For high-hardness materials, particularly for ceramics, the relationship of hardness to load is revealed very strongly. An equation is proposed for conversion of Vickers hardness from one load to another: $$HV = HV_1 \left( {\frac{P}{{P_1 }}} \right)^{1 - 2/n}$$ where HV and HV1 are the hardness with loads on the indentor of P and P1 respectively. The parameter n is determined from the equation P = const dn, where d is the indentation diagonal. The parameter n may also be determined on the basis of a normalized curve of the value of HV/E (E is Young's modulus). The physical nature of the relationship of hardness to load is discussed and the hardness \(HV_{d_f }\) is introduced with a fixed indentation diagonal df (and not with a fixed load) calculated using the equation $$HV_{d_f } = HV\left( {\frac{d}{{d_f }}} \right)^{2 - n}$$ . The introduction of \(HV_{d_f }\) makes it possible to unify measurement of microhardness for different materials at different temperatures. Curves are given simplifying conversion of hardness from one load to another and determination of the hardness \(HV_{d_f }\) .  相似文献   

18.
This paper investigates the anisotropy of the adiabatic relaxation time $T_{2}^{\prime}$ (T 2(ω), ω→∞, where ω is the Larmor frequency) of adsorbed monolayer solid 3He. We calculate $T_{2}^{\prime}$ based on the phenomenological Heisenberg Hamiltonian with the nearest-neighbor and the next nearest-neighbor exchange interactions. Furthermore, the four particle exchange in the triangular plane is also incorporated. To calculate  $T_{2}^{\prime}$ , it is necessary to compute the second and the fourth moments of the resonance line $M_{2}^{0}$ and $M_{4}^{0}$ . To obtain them, we can use the formulas published in previous studies. From the results, we can observe no significant difference between the nearest-neighbor Heisenberg model and the four particle spin interaction models. A comparison of the present results and the experimental data is briefly discussed.  相似文献   

19.
In this paper, the effect of silica fume (SF) on the chloride diffusion coefficient of concrete subjected to repeated loading was examined. Portland cement was replaced by 5 and 10 % SF. Five cycles repeated loadings were applied to concrete specimens, the maximum loadings were 40 and 80 % of the axial cylinder compressive strength ( $ f_{\text{c}}^{\prime } $ ), respectively. The diffusion coefficients were calculated from the steady state in the chloride migration test using the Nernst-Planck equation. The service life of concrete in chloride environment was predicted by Life-365 model. The results indicate that the diffusion coefficients of concrete containing 5 and 10 % SF replacements are lower than that of the control concrete at the age of 28 days. This trend increases with the increase of SF replacement. Five cycles repeated loading at 40 % $ f_{\text{c}}^{\prime } $ or 80 % $ f_{\text{c}}^{\prime } $ increase the diffusion coefficients (D 28) for all mixes investigated in this study. However, the effect of 80 % $ f_{\text{c}}^{\prime } $ on D 28 of concrete with 10 % SF is significantly lower than that of the control concrete without SF. Compared with the control concrete without SF, 10 % SF replacements increase the service life of concrete by more than 10 times.  相似文献   

20.
Inhomogeneities are known to develop within thermoelements exposed to elevated temperatures, resulting in temperature measurement errors. While the Seebeck coefficient drift in base-metal thermocouples due to aging at temperatures over \(200\,^{\circ }\mathrm{C}\) has been extensively investigated, there have been very few investigations into possible Seebeck changes at lower temperatures. Despite warnings about possible effects, most practitioners assume changes in homogeneity are either not significant or not able to develop at temperatures less than \(200\,^{\circ }\mathrm{C}\) . This study reports on measurements of inhomogeneities in base-metal thermocouples arising from heat treatment at temperatures in the region of \(200\,^{\circ }\mathrm{C}\) . Thermoelectric scans of thermocouples were carried out following exposure of a range of mineral-insulated metal-sheathed base-metal thermocouples, from two large manufacturers, of Types E, J, K, N, and T, to either a linear-gradient furnace within the range of \(100\,^{\circ }\mathrm{C}\) to \(320\,^{\circ }\mathrm{C}\) or uniform temperature zones of \(100\,^{\circ }\mathrm{C}\) , \(150\,^{\circ }\mathrm{C}\) , and \(200\,^{\circ }\mathrm{C}\) . The experiments reveal noticeable drift in all base-metal types for temperatures as low as \(100\,^{\circ }\mathrm{C}\) and exposure times as short as 1 h. The most sensitive thermoelement alloys appear to be Constantan, Alumel, and Nicrosil. It is concluded that the common working assumption that base-metal thermocouples suffer no thermally induced changes in the Seebeck coefficient below \(200\,^{\circ }\mathrm{C}\) is false. This observation has significant implications for many high-stability monitoring and control systems reliant on base-metal thermocouples that operate in the range of \(100\,^{\circ }\mathrm{C}\) to \(200\,^{\circ }\mathrm{C}\) . Additionally, thermoelectric scanning of base-metal thermocouples should be carried out at temperatures well below \(150\,^{\circ }\mathrm{C}\) to avoid erasure of strain effects or imprinting of new thermal signatures.  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