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 109 毫秒
1.
The proximity (d12) between a diffusing species and its host crystal necessary for a successful diffusion for diffusion-controlled crystallization of barium tungstate from sodium tungstate melts in platinum crucibles was estimated. These distances increased with increased cooling rates (RT) and crystallization temperatures (T0). Energy (E), enthalpy (ΔHa), entropy (ΔSa) and free-energy (ΔGa) of activation and the pre-exponential factor (k0) were evaluated using an ordinary Arrhenius equation kD1 = k0eE/RT, where kD1 was the diffusion rate-constant. These parameters were virtually unaffected by the changes in T0 and RT.  相似文献   

2.
The separation distance (d12) between a diffusing particle and its host necessary for a successful diffusion was estimated for diffusion-controlled crystal growth of BaWO4 from Na2WO4 melts. Such distances slightly increased with increased cooling rates (RT) and crystallization temperatures (T0). The energy (E), enthalpy (ΔHa), entropy (ΔSa) and free-energy (ΔGa) of activation, and the probability factor (P) were also estimated. These parameters did not change with the changes in RT and T0.  相似文献   

3.
From the estimated diffusion rate-constants (kD1) of strontium tungstate crystal growth from sodium tungstate melts in platinum crucibles, energy (E), enthalpy (ΔHa), entropy (ΔSa) and free-energy (ΔGa) of activation and the pre-exponential factor (k0) for the process were estimated using an ordinary Arrhenius equation kDl = k0eE/RT. These thermodynamic parameters were virtually unaffected by the changes in crystallization temperatures (T0) and cooling rates (R)T. The distance (r12), between a diffusing species and its host necesary for a successful diffusion, increased with T0 and TT but there was no direct correlation.  相似文献   

4.
Kinetics of strontium tungstate crystallization from sodium tungstate melts were studied in platinum crucibles (by DTA) by continuous cooling from initial crystallization temperatures T0 = 800° to 1000° to below the eutectic temperature at cooling rates RT = 0.67° to 3.3° min−1. Heterogeneous nuclei first formed slowly onto metal platinate particles within the solution during induction periods (t ); the main crystal growth then started after the development of some exces solute concentration (ΔC ) at the induction temperature (T ). The actual growth after t was diffusion rate-controlled. The diffusion rate-constants (kDt) for growth after the induction periods along the major axis were estimated; the increased with T0 and RT. These values were higher than those for diffusion-controlled crystal growth of strontium tungstate from sodium tungstate melts in alumina crucibles but much smaller than the real diffusion rate-constants (kDl)real.  相似文献   

5.
Induction periods (t) and mechanism and kinetics of nucleation in barium tungstate crystallization from sodium tungstate melts in platinum crucibles were studied. A theoretical relation has been developed to express the dependence of t on the cooling rates (RT) and the rate (Rc) of development of excess solute concentration. At any crystallization temperature the average rate-constant (kn) for heterogeneous nucleation was related to Rc by (kn/p)1/(p+1) = 1/(tRcγ), where γ is a constant and p is the average number of particles in the critical nuclei. The critical temperature (T), critical supersaturation (S) and γ values were estimated.  相似文献   

6.
Using a power‐law relation between three‐dimensional nucleation rate J and dimensionless supersaturation ratio S, and the theory of regular solutions to describe the temperature dependence of solubility, a novel Nývlt‐like equation of metastable zone width of solution relating maximum supercooling ΔTmax with cooling rate R is proposed in the form: ln(ΔTmax/T0) = Φ + β lnR, with intercept Φ = {(1–m)/m }ln(ΔHs/RGTlim) + (1/m)ln(f/KT0) and slope β = 1/m. Here T0 is the initial saturation temperature of solution in a cooling experiment, ΔHs is the heat of dissolution, RG is the gas constant, Tlim is the temperature of appearance of first nuclei, m is the nucleation order, and K is a new nucleation constant connected with the factor f defined as the number of particles per unit volume. It was found that the value of the term Φ for a system at saturation temperature T0 is essentially determined by the constant m and the factor f. The value of the factor f for a solute–solvent system at initial saturation temperature T0 is determined by solute concentration c0. Analysis of the experiment data for four different solute‐water systems according to the above equation revealed that: (1) the values of Φ and m for a system at a given temperature depend on the method of detection of metstable zone width, and (2) the value of slope β = 1/m for a system is practically a temperature‐independent constant characteristic of the system, but the value of Φ increases with an increase in saturation temperature T0, following an Arrhenius‐type equation with an activation energy Esat. The results showed, among others, that solubility of a solute is an important factor that determines the value of the nucleation order m and the activation energy Esat for diffusion. In general, the lower the solubility of a solute in a given solvent, the higher is the value of m and lower is the value of Esat. (© 2009 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim)  相似文献   

7.
The metastable zone width (MSZW, ΔTm) and induction time (tind) were determined with computer simulation for seeded batch crystallization of potassium sulfate from aqueous solution. The MSZW and induction time determined with simulation showed the same behavior as experimental values reported in the literature; log (ΔTm) increased linearly with an increase in log R (R: cooling rate) and tind decreases in proportion to (ΔT)nT: supercooling, n: nucleation order in the secondary rate expression of B=knT)n). The secondary nucleation parameters (kn and n) were deduced both from the simulated MSZW and induction times by using the previously proposed model [J. Cryst. Growth, 2010, 312, 548–554]. The secondary nucleation rate calculated with the deduced parameters was in agreement with that calculated with the parameters input for simulation.  相似文献   

8.
9.
The splay (k 11), twist (k 22) and bend (k 33) elastic constants determined by the Freedericksz method and the orientational order parameters (s) derived from optical measurements in the nematic phase of six homologues of trans-p-n-alkoxy-α-methyl cyanophenyl cinnamates (n OMCPC) are reported. The data close to the nematic-isotropic transition point (T NI) are compared with T NI and the heats of transition (ΔH). The temperature-variation of elastic constants is discussed in terms of existing theories. The pretransitional increase in the twist and bend constants near the nematic-smectic A transition point (T NA ) of 10 OMCPC has also been analysed.  相似文献   

10.
Based on thermodynamic characteristics of the stable metallic liquid at melting temperature and the supercooled liquid, the present work calculated the mixing enthalpy ΔHmix, the mixing entropy ΔSmix and the Gibbs free energy difference between the supercooled liquid and the resulting crystalline phases ΔG of typical Ti-based amorphous alloys. The results show that for the case of larger ΔSmix, moderate ΔHmix for the stable liquid and smaller ΔG for the supercooled liquid, Ti-based alloys tend to achieve high glass-forming ability (GFA). A new parameter, β, defined as (Tg ? Tk)/(Tl ? Tg), has been introduced to evaluate the GFA of Ti-based bulk amorphous alloys (wherein Tg, Tl, and Tk represent the glass transition temperature, the liquidus temperature, and the Kauzmann temperature, respectively). Experimental data imply that the larger the β, the better the GFA for Ti-based amorphous alloys.  相似文献   

11.
The crystallisation kinetics of strontium tungstate from unstirred saturated solutions in sodium tungstate melts was studied by continuous cooling from initial crystallisation temperatures T0 = 1000° to 800°C to room temperature at cooling rates RT = 0.67° to 3.3°C min−1. The main crystal growth was diffusion rate-controlled; the final crystal growth was rate-controlled by the development rate of excess solute concentration. The estimated diffusion rate constant (kD) values increased with cooling rates and initial crystallisation temperatures. They are higher than the rate constants for diffusion-controlled growth of calcium tungstate from sodium tungstate melts, but very much smaller than those for strontium tungstate from lithium chloride melts.  相似文献   

12.
X-ray small-angle scattering (XSAS) and resistivity (R) measurements were particularly done with an Al-Zn (15 at·,%) alloy in rather wide ranges of both the ageing time ta and ageing temperature Ta, in order to obtain information on the dependence of the growth exponent m of the l = β0tm growth law and the activation energy Eact on ta and Ta. The XSAS-measurements yielded that within the range of the GUINIER radius rG between 1 nm and 2 nm the growth is essentially retarded (m < 0.1) and for rG > 2 nm m depends on Ta ranging from 0.15 to 0.23 with a maximum at 175°C. Reasons for these effects are discussed. The differences between the m-values obtained by means of XSAS-and TEM-measurements are explained by the distinctions of the two methods applied. The Eact taken from XSAS- and R-measurements show a remarkable increase with ta. At the beginning of the decomposition Eact = (0.49 ± 0.05) eV holds well explainable by the migration of quenched-in VZn pairs, but at the end Eact = (1.05 ± 0.07) eV was found. This value was also obtained from TEM-investigations (growth of the length). It fits well the Eact of ZnV pairs in thermal equilibrium at Ta.  相似文献   

13.
G.J. Fan  H. Choo  P.K. Liaw 《Journal of Non》2007,353(1):102-107
Based on theoretical calculations using the fragility concept and the nucleation theory for a model glass-forming system, we propose a dimensionless criterion, ?, expressed by TrgTx/Tg)a, with Trg, the reduced glass-transition temperature, ΔTx, the width of the supercooled liquid region when heating a glass, Tg, the glass transition temperature, and a, the exponent. The application of this simple criterion to various glasses, including network, metallic, and molecular glasses (except pure water), indicates an excellent correlation between the critical cooling rate Rc and ? in a Log Rc-? single master plot with a = 0.143.  相似文献   

14.
The experimental data published in the literature on the metastable zone width, as determined by the maximum supercooling ΔTmax using the conventional polythermal method, of phosphoric acid aqueous solutions containing impurities were analyzed to understand an increase in ΔTmax/T0 with an increase in saturation temperature T0 of solute–solvent system and the effect of impurities on the metastable zone width. For the analysis the following relations were used: ln(ΔTmax/T0)=Φ+βln R (K. Sangwal, Cryst. Res. Technol. 44, 2009, 231−247) and (T0Tmax)2=F(1−Zln R) (K. Sangwal, Cryst. Growth Des. 9, 2009, 942−950; J. Cryst. Growth 311, 2009, 4050−4061), where Φ, β, F and Z are constants. Analysis of the experimental data revealed that: (1) the parameters Φ and F strongly depend on saturation temperature T0 and concentration ci of impurities, but the constants β and Z are independent of T0 and depend on ci, (2) the dependence of the parameters Φ and F on T0 follows an Arrhenius-type equation with activation energy Esat, (3) the activation energy Esat for diffusion of ions/molecules of phosphoric acid containing impurity ions is equal to the differential heat of adsorption Qdiff for these impurities, (4) the effectiveness of an impurity is directly connected with the values of their differential heat of adsorption Qdiff; the lower the values of Qdiff for an impurity, the lower is its effectiveness in promoting nucleation, (5) the activation energy Esat is not related with its heat of dissolution ΔHs and (6) the increase in ΔTmax/T0 with an increase in T0 for phosphoric acid is associated with the activation energy Esat for diffusion of solute molecules in the solution such that Esat<0.  相似文献   

15.
Low-frequency (20 Hz–10 MHz) dielectric response in liquid crystalline (LC) phases of hydrogen-bonded (HB) complex, SA:11OBA, is studied. Synthesis of SA:11OBA with non-mesogenic succinic acid (SA) and mesogenic p-n-undecyloxy benzoic acid (11OBA) and its spectroscopic confirmation are detailed. Phase transition temperatures (TC) involving nematic, smectic-C (SmC), and smectic-G (SmG) phases by capacitance C(T) and loss tanΔ(T) anomaly agree with the microscopy (polarizing optical microscope) and calorimetry (differential scanning calorimetry) results. Low-frequency dispersion infers two types of reorientation processes, viz. higher frequency (~MHz) and lower frequency (kHz) processes. Distinct time scale mechanism are presented. Arrhenius behavior infers influence of HB on activation energy (Ea). Off-centered dispersion, viz. ?′(ω) with ?″(ω), analyzed through the Cole–Cole plots in N, SmC, and SmG phases infers a strong temperature trend for dielectric strength (Δ?) and distribution (α) parameters. Temperature dependence of α-parameter reveals on LC phases increasing fixture of molecular dipole in LC phase structure. Results are discussed in the wake of the body of data reported in LC phases exhibited by other LC compounds.  相似文献   

16.
The electrical properties of n-type titanium dioxide thin films deposited by magnetron-sputtering method have been investigated by temperature-dependent conductivity. We observed that the temperature dependence of the electrical conductivity of titanium dioxide films exhibits a crossover from T?1/4 to T?1/2 dependence in the temperature range between 80 and 110 K. Characteristic parameters describing conductivity, such as the characteristic temperature (T0), hopping distance (Rhop), average hopping energy (Δhop), Coulomb gap (ΔC), localization length (ξ) and density of states (N(EF)), were determined, and their values were discussed within the models describing conductivity in TiO2 thin films.  相似文献   

17.
The formation volume Vv of vacancies is given by Vv = (hv/L) ΔVf with hv = 8L (formation enthalpy hv of vacancies and heat L of fusion given in same units; ΔVf = change of volume due to melting). If there are phase transitions within the solid, L and ΔVf must be replaced by (L + Δ Ht) and by (ΔVf + ΔVt), respectively (Δ Ht and Δ Vt refer to the heat (s) of transition (s) and to the volume change(s) due to transition(s), resp.). The pressure dependence of the melting point is dTm/dp = (TmVv)/hv. Independent of the sign of Vv any increase of the vacancy concentration above the maximum concentration possible within the solid decreases the melting point thus resulting in the observed surface melting. The melting point is fixed by the characteristics of vacancy formation (hv, Sv, Vv) and by the bulk modulus of the solid (Sv = formation entropy of vacancies).  相似文献   

18.
Differential scanning calorimetry (DSC) studies were performed under nonisothermal conditions at various heating rates for glassy Se made by high-energy ball milling. Comparisons were made between the ball-milling technique and the melt-quenching and thin-film techniques. Well-defined endothermic and exothermic peaks were observed at the glass-transition temperature, Tg, and the onset temperature of crystallisation, Tc. The isoconversional method of Vyazovkin was used to determine the variation in the activation energy for crystallisation with temperature, Eα(T). The value of Eα(T) was dependent on the sample preparation method. The thermal stability of the Se glasses was evaluated by calculation of the temperature difference (Tc ? Tg) and the S-parameter. In addition, the glass-forming ability was estimated by the criteria of reduced glass-transition temperature, Trg, and the Hruby number, HR. The structures of the Se samples that resulted from the DSC analysis were identified using an X-ray diffractometer. The glasses formed using the thin film technique were the most stable.  相似文献   

19.
Abstract

Magnetic (ΔnH ) and electric birefringence (ΔnE ) in the isotropic phase of strongly positive (Δ 8)trans-p-n-octyloxy α-methyl-p'-cyanophenyl cinnamate (8 OMCPC) have been measured. It is established that they both exhibit a (T - T*)?1 dependence, T NI - T* being 1.4 K. Also, the induced birefringence is found to be proportional to the square of the applied field, magnetic or electric.  相似文献   

20.
Abstract

The heat capacities of the title compound (C3H11,O—C6H4,- CH=N—C6H4,—C4H9, abbreviation 5O ? 4) with a purity of 99.92 mole percent have been measured with an adiabatic-type calorimeter between 11 and 393 K. The transition temperature and the enthalpy and entropy of phase transition for stable crystal → SG, SG → N and N → isotropic liquid were T c = 299.69 K/ΔH = 22.68 kJ mol?1/ΔS = 75.70 JK?1 mol?1, 325.72/7.11/21.79 and 342.48/1.78/5.22, respectively. The crystal which melts at 285.5 K is a metastable modification. The SA phase hitherto reported in between SG and N does not exist. The glassy So state was realized by rapid cooling of the specimen from the So phase. The molar enthalpy of the glassy SG state at 0 K was by (10.1±0.1) kJ mol?1 higher than that of the stable crystalline state and the residual entropy of the glassy state was (9.40±0.83) JK?1 mol?1. The relaxational heat-capacity anomaly was observed from as low as 100 K and double glass transition phenomenon occurred around 200 K; a quite unusual phenomenon which has never been observed for the glassy states of nematic and cholesteric liquid crystals. The present results give a fair evidence that the unusual glass transition phenomenon previously found for the SG state of 6O?4 (a homologous compound) is not exceptional at all but common to the smectic glasses; at least common to the glassy SG states. Two possible origins responsible for the double glass transitions have been discussed.  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