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.
Mauro C.C. Ribeiro 《Journal of Non》2009,355(31-33):1659-1662
Molecular dynamics (MD) simulations of LiCl·6H2O showed that the diffusion coefficient D, and also the structural relaxation time <τ>, follow a power law at high temperatures, D?1  (T ? To)?μ, with the same experimental parameters for viscosity (To = 207 K, μ = 2.08). Decoupling between D and <τ> occurs at Tx  1.1To. High frequency acoustic excitations for the LiCl·6H2O model were obtained by the calculation of time correlation functions of mass current fluctuations. The temperature dependence of the instantaneous shear modulus, G(T), was considered in the shoving model for supercooled liquids [J.C. Dyre, T. Christensen, N.B. Olsen, J. Non-Cryst. Solids 352 (2006) 4635] resulting in a linear relationship log (D?1) vs. G/T.  相似文献   

2.
Raman scattering spectra in glass forming toluene were studied in the temperature range 50–323 K with the goal of extracting information about homogeneous, inhomogeneous and orientational broadening. It was found that the temperature dependence of inhomogeneous line width allows one to depict two peculiar temperatures: TA and Tg, where Tg is the glass transition temperature and TA is the temperature of transition from an Arrhenius-like to a non-Arrhenius behavior for the α-relaxation time dependence on temperature, τα(T). Temperature dependence of the orientational phase loss time τOPL was found to correspond well to τα at T > TA and continues approximately Arrhenius behavior for lower temperature in contrast to τα(T). Also, a comparative analysis of homogeneous broadening of polarized and depolarized lines was done, which provided an estimation of the orientational broadening γNL(T). The found γNL(T) decreases linearly as the temperature decreases and goes to zero at T ~ Tc, where Tc is the critical temperature in framework of the mode-coupling theory (note that Tc is close to other peculiar temperatures TB and Tβ, but we did not intent to distinguish among them in the present work). Thus, it was shown that the Raman line shape analysis in molecular glass forming materials allows one to extract peculiar temperatures: TA, Tg, and, probably, Tc.The test of the possibility to use a probe molecule for the Raman line shape analysis has revealed that the extracted data for probe molecule lines do not characterize the host matrices, at least in the low-viscous state (T > TA).  相似文献   

3.
《Journal of Non》2007,353(41-43):3853-3861
The molecular dynamics of glass-forming poly(methyl phenyl siloxane) (PMPS) is studied by thermal (10−3–5 × 102 Hz), dielectric (10−3–109 Hz) and neutron (5 × 108–1012 Hz) spectroscopy. Because of the broad frequency range of 15 orders of magnitude the study provides a precise determination of glassy dynamics in a wide temperature range using different probes. The relaxation rates extracted from the different methods agree quantitatively in both their absolute values and in their temperature dependencies. A detailed analysis of the temperature dependence of the relaxation rate fp by a derivative technique shows that the α-relaxation of PMPS has to be characterized by a high and a low temperature branch separated by a crossover temperature TB = 250 K. In both temperature ranges the temperature dependence of fp has to be described by Vogel/Fulcher/Tammann laws with different Vogel temperatures. Also the analysis of the dielectric strength in its temperature dependence gives a crossover behavior from a low to a high temperature region with a similar value of TB. TB can be interpreted as onset of cooperative fluctuations and the formation of dynamical heterogeneities. The dependence of the relaxation rate on the scattering vector Q extracted from neutron scattering obeys a power law τ  Q−Slope, where the power Slope varies between Slope = 2 and Slope = 3.5 with increasing temperature. This anomalous dependence of the relaxation time on the momentum transfer is discussed in terms of dynamic heterogeneities in the underlying motional processes even at temperatures above TB. Besides the segmental dynamics the fast Methyl group rotation is considered as well. The relaxation rates of this process have an activated temperature dependence with an activation energy of 8.3 kJ/mol. The data were discussed in the framework of the threefold jump model were the incoherent elastic scattering from ‘fixed’ atoms which are frozen on the time scale of the Methyl group rotation was taken into account.  相似文献   

4.
The relaxation of the transverse magnetization components caused by both dipolar interactions between the spins of different polymer chains and the dipolar coupling between CH-protons on an isolated Kuhn segment along a single polymer chain have been calculated. Explicit expressions for the transverse relaxation function are given in terms of the absolute mean squared displacement of the Kuhn segment during melt gr(t), the tangent vector dynamical correlation function 〈bn(t)b(0)〉, the segmental relaxation time τs, the Kuhn segment length b, the bond length a0, the internuclear distance d, and the spin number density ρs. It is shown that the functional dependence of the intramolecular relaxation function on 〈bn(t)b(0)〉 is fairly weak. The time-dependence of the intramolecular contribution to the transverse relaxation function is dominated by the probability density distribution function of the end-to-end vector of the Kuhn segment. The long-time decay of the intramolecular contribution to the transverse relaxation function is found to scale as t? 3/2 for τs < < t < < τmax, where τmaxis the maximum relaxation time of polymer chains in melts. For times much less than the spin–spin relaxation time, T2  10? 3 ? 10? 2s, we show that the intermolecular contribution to the relaxation function is given by the following expression: exp(? λ1(b, τs, ρs)t2/gr3/2(t)). Both the numerical coefficient and the functional dependence of λ1on b, τs and ρs reproduce the expression obtained from the frequently used second cumulant approximation. For longer times (T2  t < < τmax), the intermolecular contribution is determined by the following relation: exp(? λ2(b, τs, ρs, t)gr(t)). We show that λ2 increases logarithmically with t. The molecular mass independence of λ1and λ2 shows that, in polymer melts with molecular masses Mw far above the critical value Mc, the relevant experimental window for the decay of the intermolecular relaxation function is connected with the anomalous diffusion regime. Comparison with the experimental data suggests that the intermolecular contribution plays a significant role in the NMR relaxation process in polymer systems close to the melting point.  相似文献   

5.
《Journal of Non》2006,352(30-31):3310-3314
The structure of superionic glasses in ternary systems x(0.4Li2S–0.6P2S5)–(1  x)LiI and x(0.5Li2S–0.5P2S5)–(1  x)LiI (x = 0.9, 0.75) has been studied by molecular dynamics (MD) simulations. The configurations obtained by MD were analyzed by graph theory. Phosphorus is surrounded by sulfur and iodine atoms. Li is surrounded by sulfurs alone and LiI clusters are not present as speculated by earlier spectroscopic reports. The equilibrium configuration is made up of (Li, S) and (P, S, I) rich regions which creates wide channels for Li+ movement. Reported variations of glass transition temperature (Tg) and ionic conductivity (σ) with LiI addition are explained based on the simulation results.  相似文献   

6.
《Journal of Non》2006,352(42-49):4946-4955
Dimensional (D) and enthalpy relaxation (ΔH) of oriented polymer glasses (PS and PC) have been studied as function of temperature, between Tg and Tg−20 °C, and aging time t, ranging to several weeks. The dimensional relaxation (shrinkage) and enthalpy relaxation curves verify the logarithm law D(t)  H(t)  log t, between an incubation τi and a final relaxation time τf. The time τf to reach the equilibrium (D and ΔH) follows the Vogel–Tamann–Fulcher (VFT) law. Enthalpy relaxation and shrinkage exhibit important differences. Enthalpy relaxation of oriented and isotropic polymers follows the same logarithm law, independent of the draw ratio λ and the mode of deformation, the relaxation time τf coincides with the relaxation time of the α segmental motions. Shrinkage depends on λ and the mode of deformation, the relaxation time τf is attributed to the normal mode, the relaxation time of the whole chain. Finally the shrinkages of PS and PC show some differences. PC at short aging times presents another type of dimensional relaxation which would be due to the β motions. This would be in close connection with the ductile (PC) and fragile (PS) behavior of these two polymers far below Tg.  相似文献   

7.
《Journal of Non》2007,353(11-12):1065-1069
In the present work the dependence of electrical properties of a-SiC:H thin films on annealing temperature, Ta, has been extensively studied. From the measurements of dark dc electrical conductivity, σD, in the high temperature range (from 283 up to 493 K), was found that the conductivity activation energy, Ea, is invariant for Ta  673 K and equal to 0.64 eV, whereas for Ta from 673 up to 873 K, Ea increases at about 0.2 eV reaching to a maximum value 0.85 eV at Ta = 873 K, suggesting the optimum material quality. This behavior of Ea as a function of Ta is mainly attributed to relaxation of the strain in the amorphous network, which is possibly combined with weak hydrogen emission for temperatures up to 873 K. For further increase of Ta (>873 K) the phenomenon of hydrogen emission, causes rapid decrease of Ea down to 0.24 eV at Ta = 998 K, deteriorating the material quality. These results are also supported by the measurements of dark dc electrical conductivity in the low temperature range (from 133 up to 283 K), where the dependence of the density of gap states at the Fermi level, N(EF), on annealing temperature presents the minimum value at Ta = 873 K. The Meyer–Nelder rule was found to hold for the a-SiC:H thin films for annealing temperatures up to 873 K. Finally, the dependence of dark dc electrical conductivity at room temperature, σDRT, on Ta showed to reflect directly the dependence of Ea on Ta.  相似文献   

8.
We measured and collected literature data for the crystal growth rate, u(T), of μ-cordierite (2MgO · 2Al2O3 · 5SiO2) and diopside (CaO · MgO · 2SiO2) in their isochemical glass forming melts. The data cover exceptionally wide temperature ranges, i.e. 800–1350 °C for cordierite and 750–1378 °C for diopside. The maximum of u(T) occurs at about 1250 °C for both systems. A smooth shoulder is observed around 970 °C for μ-cordierite. Based on measured and collected viscosity data, we fitted u(T) using standard crystal growth models. For diopside, the experimental u(T) fits well to the 2D surface nucleation model and also to the screw dislocation growth mechanism. However, the screw dislocation model yields parameters of more significant physical meaning. For cordierite, these two models also describe the experimental growth rates. However, the best fittings of u(T) including the observed shoulder, were attained for a combined mechanism, assuming that the melt/crystal interface growing from screw dislocations is additionally roughened by superimposed 2D surface nucleation at large undercoolings, starting at a temperature around the shoulder. The good fittings indicate that viscosity can be used to assess the transport mechanism that determines crystal growth in these two systems, from the melting point Tm down to about Tg, with no sign of a breakdown of the Stokes–Einstein/Eyring equation.  相似文献   

9.
The transport properties in La0.7?xYxPb0.3MnO3 (0.0 ? x ? 0.2) is investigated. The substitution of La3+ ions by smaller nonmagnetic Y3+ leads to greater spin disorder and induces variations in the magnetotransport behavior. From resistivity versus temperature curves a metal–insulator transition phenomenon is observed at the transition temperature, TP, decreases as the Y content increases. The resistivity is well fitted using the equation ρ(T) = ρnexp[(T1/T)n] with n = 1/4 and n = 1/2 at high and intermediate temperatures, respectively. The characteristic temperature T1 varies with Y content in a manner consistent with the localization model of variable range hopping. Below TP, resistivity varies as a function of power law contributions, ρ = ρ0 + ρ2T2 + ρ5/2T5/2, corresponding to the electron scattering processes in the ferromagnetic phase.  相似文献   

10.
《Journal of Non》2006,352(42-49):4735-4741
We compare the susceptibility spectra (10−6 Hz–1012 Hz) of glass forming liquids and plastically crystalline (PC) phases. In both the cases, a similar spectral change is observed while cooling. Whereas at high-temperatures the frequency-temperature superposition (FTS) principle holds for the α-process with a stretching parameter significantly below 1 it fails below a certain temperature Tx. Below Tx, in the case of supercooled liquids, in addition to a broadening of the α-relaxation peak the excess wing appears, and the corresponding power-law exponents β (α-peak) and γ (excess wing) of the distribution of correlation times show a similar dependence on the time constant τα, explicitly 1/β and 1/γ both are linear in lg τα. In the PC systems studied, an excess wing is missing and the failure of the FTS principle for the α-relaxation peak directly shows up below Tx. Again, the parameter of the Cole–Davidson susceptibility 1/βCD is linear in lg τα below Tx, and constant above, allowing to identify Tx. In PC phases the crossover temperature Tx may be found much closer to the glass transition temperature Tg as compared to supercooled liquids, and thus can be well studied by standard dielectric spectroscopy.  相似文献   

11.
Thermal diffusivity (D) at high temperature (T) was measured from 15 samples of commercial SiO2 glasses (types I, II, and III with varying hydroxyl contents) using laser-flash analysis (LFA) to isolate vibrational transport, in order to determine effects of impurities, annealing, and melting. As T increases, Dglass decreases, approaching a constant (~ 0.69 mm2s? 1) above ~ 700 K. From ~ 1000 K to the glass transition, the slope of D is small but variable. Increases of D with T of up to 6% correlate with either low water and/or low fictive temperature and are attributed to removal of strain and defects during annealing. Upon crossing the glass transition, D substantially decreases to 0.46 mm2s? 1 for the anhydrous melt. Hydration decreases Dglass, makes the glass transition occur over wider temperature intervals and at lower T, and promotes nucleation of cristobalite from supercooled melt. Due to the importance of thermal history, a spread in D of about 5% occurs for any given chemical type. Combining prior steady-state, cryogenic data with our average results on type I glass provides thermal conductivity (klat = ρCPD) for type I: klat increases from ~ 0 K, becoming nearly constant above 1500 K, and drops by ~ 30% at Tg. We find that D? 1(T) correlates with thermal expansivity times temperature from ~ 0 K to melting due to both properties arising from anharmonicity.  相似文献   

12.
The scope of this work is to determine the crystalline phases of devitrified barium magnesium phosphate glasses and the glass composition which presents the best resistance to crystallization. Barium magnesium phosphate glasses with composition xMgO · (1 ? x)(60P2O5 · 40BaO) mol% (x = 0, 0.15, 0.3, 0.4, 0.5, and 0.6) were analyzed by differential thermal analysis (DTA) to evaluate the thermal stability against crystallization, and X-ray diffraction (XRD) to identify the crystalline phases formed after devitrification. The glass transition temperature (Tg) increases as the MgO content increases. The maximum temperature attributed to the crystallization peak in the DTA curve (Tc) increases when x increases in the range 0 ? x ? 0.3, and it decreases for x > 0.3. The most thermally stable glass composition against crystallization is for x = 0.3. After the devitrification, the number of coexisting crystalline phases increases as the MgO content increases. For x = 0.3 there is the coexistence of γBa(PO3)2 and Ba2MgP4O13 phases for devitrified glasses. The trend of the Tc is explained based on the assumptions of changes in the Mg2+ coordination number and the amphoterical features of MgO.  相似文献   

13.
《Journal of Non》2006,352(21-22):2129-2136
Conductivity vs. temperature (σT) at zero and 8-T field, magneto-resistance (Δρ/ρ), magnetization vs. temperature (M–T) and magnetization vs. field (M–H) of Al70.5Pd22Mn7.5 quasicrystal have been studied in the temperature range of 1.4 K to 300 K. The σT variations in both the field conditions show a σT minima. In addition to this the σT variation at 8 T shows a maxima also at ∼6 K. Comparative analysis shows that the observed σT minima arises due to competing inelastic scattering events in the presence of weak-localization effect. These events are e–ph scattering in the dirty-metallic limit (τi  T−2), and the Kondo-type spin-flip scattering (τsf). The maxima observed for the σT variation at 8 T, has been attributed to suppression of the spin-flip scattering in the presence of field. The magneto-resistance is found to be large and positive. It was properly accounted only when the Stoner-enhancement, found in the case of spin fluctuating systems, was taken into consideration.  相似文献   

14.
Yuanzheng Yue 《Journal of Non》2009,355(10-12):737-744
This paper describes how the fragility of a liquid is linked to the ratio between the energy barrier (Eeq) for the equilibrium viscous behavior and that (Eiso) for the non-equilibrium iso-structural viscous behavior. Using the concept of iso-structural viscosity, two functions describing the variation of the configurational entropy (Sc) with temperature (T) are obtained from the Avramov-Milchev (AM) and the Vogel-Fulcher-Tammann (VFT) viscosity equations, respectively. The two Sc(T) functions exhibit different relations to the liquid fragility. The AM Sc(T) function is a power function with the exponent of F ? 1, where F is the AM fragility index. In this case, Sc vanishes at T = 0 K. For the VFT function, Sc vanishes as T is lowered to a finite temperature T0, whereas it reaches the maximum value Sc,max at infinitively high T. Sc,max is proportional to the VFT fragility index. Thus, the VFT equation is not only a dynamical, but also a thermodynamic model. It is proved that for oxide liquids, the VFT equation describes viscosity data better than the AM equation, provided the pre-exponential factor η0 is fixed to a generally accepted value, e.g., 10?3.5 Pa s.  相似文献   

15.
In this work, the thermal lens spectrometry was applied to measure the thermo-optical properties of Nd2O3-doped low silica calcium aluminosilicate glasses as a function of temperature, between 4.3 and 300 K. The thermal relaxation calorimetry was used to determine the specific heat, cp. The results showed a decrease of the thermal diffusivity of about one order of magnitude from 4.3 K up to 300 K, with a T?1 dependence in the interval between 20 and 70 K and a T?0.35 between 4.3 and 20 K. The fluorescence quantum efficiencies of the doped samples were calculated down to 50 K, showing a variation of the order of 12% and 25% for the samples with 0.6 and 1.04 mol% of Nd2O3, respectively. In addition, the temperature corresponding to the maximum in cp/T3, the so-called boson peak, was observed at about 17 K for the undoped sample and at lower temperatures for the doped glasses. In conclusion, our results showed the ability of the time resolved thermal lens to determine the thermo-optical properties of glasses at temperatures lower than 300 K, bringing new possibilities for experiments in a wide range of optical materials.  相似文献   

16.
The elastic properties of alkali germanate glasses, xR2O?(100 ? x)GeO2 (R = Li, Na, K, Rb, Cs ; x = 14, 28), have been studied by Brillouin scattering in the wide temperature range up to 1200 °C. The remarkable aging effect of Brillouin shift ΔνL has been observed below a glass transition temperature Tg  500 °C. The temperature dependence of longitudinal sound velocity VL of well annealed glasses shows the gradual decrease below Tg, while on further heating the remarkable decrease is observed above Tg. The scaled temperature dependence of VL is nearly independent on alkali metals below the melting temperature Tm. While on further heating above Tm, the drastic decrease of VL and increase of αL show the remarkable alkali dependence. It may be attributed to the appearance of dynamic process related to ionic hopping of alkali metals released from glass network above Tm.  相似文献   

17.
《Journal of Non》2007,353(32-40):3113-3121
The collective SC(Q, ω), single-particle SS(Q, ω) dynamic structure factors as well as the spectra for transverse-current correlations CT(Q, ω) for molten nickel close to melting have been calculated from a massive molecular dynamics simulation using an Embedded Atom Model potential. A number of features are found in the structure factors that are absent in the spectra of simpler liquids such as molten alkali metals. Estimates for the shear component of the generalized, wavevector-dependent viscosity are obtained from analysis of CT(Q, ω). Such data together with those concerning the strength of the thermal conduction channels for the excitation decay enable us to subtract such contributions to form the linewidths of Brillouin peaks and thus to evaluate the Q-dependent bulk viscosity coefficient. The result shows that contrary to inferences based upon geophysical data, both bulk and shear-viscosity coefficients are of the same order of magnitude.  相似文献   

18.
《Journal of Non》2007,353(8-10):859-861
Metastable Fe25Cu75 solid solution obtained by high-energy ball milling has been studied by means of neutron thermo-diffraction and magnetization measurements. This material, which crystallizes in a face centered cubic (FCC) crystal structure, is ferromagnetic, with a value for the Curie temperature, around 190 K. The linear thermal expansion coefficient, αT, has been estimated from the temperature dependence of the lattice parameter in the temperature range 5–300 K. An increase in the value of αT following the normal trend observed in other metallic compounds is observed below 150 K, however, above this temperature αT remains almost constant indicating a slight magneto-volume effect in this material.  相似文献   

19.
S. Striepe  J. Deubener 《Journal of Non》2012,358(12-13):1480-1485
Kinetic fragility indices m and F1/2 as well as glass transition temperature Tg of alkaline earth zinc phosphate melts of molar composition 20 MO–30 ZnO–50 P2O5 (with M = Ba, Sr, Ca, Mg) were determined using viscometry and differential scanning calorimetry (DSC). Beam bending and concentric cylinder experiments were performed to measure the flow resistance in temperature ranges above glass transition and close to liquidus, respectively. Different upscan rates of DSC runs through the glass transition were used to correlate changes of the fictive temperature with kinetic fragility. Both methods revealed that glass transition temperature correlates negatively and kinetic fragility positively with the size of M. Metal cation mixing (M + Zn) led to a negative deviation from linearity for Tg, while exchanging M resulted in a linear dependence of Tg, if scaled with averaged charge-to-distance ratio. The fictive temperature method overestimated the compositional dependence of m by a ratio up to 1.9.  相似文献   

20.
《Journal of Non》2006,352(28-29):3103-3108
The thermal behavior of (Pt0.4Pd0.3Ni0.3)100−xPx (x = 16–25 at.%) glassy alloys has been investigated. It is found that the crystallization behavior of the (Pt0.4Pd0.3Ni0.3)100−xPx glassy alloys changes from a single-stage exothermic reaction to a two-stage exothermic reaction depending on phosphorous content. When the phosphorous content is 23 at.%, the glassy alloy exhibits the largest supercooled liquid region (ΔTx) and a sharp single exothermic peak. Fixing the phosphorous content at 23 at.%, the Pt77−xyPdxNiyP23 (x = 7.7–61.6 at.%, y = 7.7–61.6 at.%) glassy alloys have a wide composition range in which the glassy alloys exhibit a large supercooled liquid region (ΔTx beyond 60 K). In this range, the Pt30.8Pd23.1Ni23.1P23 glass has the largest ΔTx (77 K) and a high reduced glass transition temperature (Trg) of 0.60. This alloy can be cast into fully glassy rods with a diameter of 3 mm. Under uni-axial compression, bulk Pt30.8Pd23.1Ni23.1P23 glassy alloy has an elastic strain of ∼2%, an ultimate strain (to fracture) of ∼6.4%, a Young’s modulus of ∼106 GPa and a failure strength of ∼1390 MPa.  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