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.
Taylor dispersion is widely used to measure binary mutual diffusion. Studies of three- and four-component solutions show that the dispersion method is also well suited for multicomponent diffusion measurements, including cross-coefficients for coupled diffusion. Numerical procedures are reported here to calculate mutual diffusion coefficients from dispersion profiles measured for solutions of any number of components. The proposed analysis is used to measure the sixteen quinary mutual diffusion coefficients of five-component aqueous mannitol + glycine + urea + KCl solutions and aqueous NBu4Cl + LiCl + KCl + HCl solutions. Mannitol, glycine, urea and KCl interact weakly at the low solute concentrations used (0.010 mol·dm?3). The diffusion coefficients of this system are compared with pseudo-binary predictions. Strong coupling of the NBu4Cl, LiCl, KCl and HCl fluxes is interpreted by using ionic conductivities and Nernst equations to calculate limiting quinary diffusion coefficients for mixed electrolytes that interact by the electric field generated by ion concentration gradients.  相似文献   

2.
3.
Excess molar volumes, $ V_{123}^{\text{E}} $ V 123 E , of 1, 3-dioxolane (1) + toluene (2) + o- or p-xylene (3) ternary mixtures have been determined dilatometrically over the entire composition range at 298.15 K. For thermodynamic consistency the experimental values were fitted to Redlich–Kister Equation. The $ V_{123}^{\text{E}} $ V 123 E values of 1, 3-dioxolane (1) + toluene (2) + o- or p-xylene (3) ternary mixtures have been found to be negative over the whole composition range. It has been observed that $ V_{123}^{\text{E}} $ V 123 E values calculated by graph theory are of the same sign and magnitude with respect to their experimental values.  相似文献   

4.
The excess viscosities, ηE, and excess energy of activation (ΔΕη)Ε of dynamic viscosity have been investigated by using dynamic viscosity measurements for N,N-dimethylformamide + water (DMFW) mixtures over the entire range of mole fractions at five different temperatures. The results were also fitted with the Redlich–Kister equation. This system exhibited very large positive values of ηE and (ΔΕη)Ε due to the increased dipole–dipole interactions and correlation length between unlike molecules. The activation parameters ΔΗσ and ΔSσ have been also calculated, and they show that the critical region has an important effect on the dynamic viscosity properties. The results obtained are discussed from the viewpoint of the existence of interactions between the components.  相似文献   

5.
Influence of inorganic salts on the system of liquid phase equilibrium of water + furfuryl alcohol + cyclopentanone at 298.2 K was studied. Different salt concentrations (0, 1 and 2 wt%) and the type of salt (LiCl, NaCl, KCl, and RbCl) were investigated. The results showed that the two-phase region of the ternary system enlarged by addition of salt. NRTL model was applied, and good correlation between the experimental data and the model was achieved as confirmed by the low rmsd values.  相似文献   

6.
Liquid–liquid equilibria of the methanol + ethylbenzene + methylcyclohexane ternary system are reported at 278.15, 283.15, and 293.15 K. The effect of the temperature on the liquid–liquid equilibrium is discussed. All chemical concentrations were quantified by gas chromatography using a thermal conductivity detector. Experimental data for the ternary system are compared with values calculated by the NRTL and UNIQUAC equations. It was found that both equations gave comparable quality representations of the experimental data for this ternary system. Distribution curves were also analyzed. Data for the ternary system is available from the literature at 303.15 K.  相似文献   

7.
A phenomenological model of \(\mathrm {CO}\) oxidation with \(\mathrm {NO}\) reaction proceeding over composite (supported) catalysts is proposed and solved numerically using the finite difference method. The model is based on the coupled system of PDEs subject to nonclassical conjugate conditions at the catalyst-support interface and includes: the bulk diffusion of reactants from a bounded vessel towards the catalyst surface and the bulk one of the reaction products from the surface into the same vessel, adsorption and desorption of particles of reactants, and surface diffusion of adsorbed molecules. The readsorption of the reaction product N\(_2\)O is also taken into account. The influence of the rate constants of the adsorbed particle jumping via the catalyst-support interface and reaction rate constants on the surface reactivity is investigated. It is shown that the turnover rates of the CO and NO into products N\(_2\)O, CO\(_2\), and N\(_2\) are nonmonotonic time functions and depending on values of the kinetic parameters may possess one or two maxima. The N\(_2\)O readsorption in case of the existence of two maxima essentially increases the turnover rates and extends the duration of their high values. The mechanism and conditions for arising of the second maximum is discussed. It is also shown that the variation of the particle jumping rate constants influences differently the size of the jump discontinuity of concentrations of different adsorbates at the catalyst-support interface.  相似文献   

8.
9.
The preferential solvation parameters (δx1,3) of genistin in ethanol/acetone (1) + water (2) and daidzein in ethanol (1) + water (2) co-solvent mixtures at elevated temperatures were derived from available solubility data using the inverse Kirkwood–Buff integral method. The values of δx1,3 varied non-linearly with the co-solvent (1) proportion in all the aqueous mixtures. For the three co-solvent mixtures, the values of δx1,3 were negative in water-rich mixtures, which indicated that daidzein or genistin was preferentially solvated by water and can act as Lewis bases to establish hydrogen bonds with the proton-donor functional groups of water (1). The same behavior was also observed for daidzein in ethanol (1) + water (2) and acetone (1) + water (2) mixtures with co-solvent-rich composition. For daidzein in ethanol (1) + water (2) mixtures with composition 0.24 < x1 < 1, and genistin in ethanol (1) + water (2) and acetone (1) + water (2) mixtures with intermediate compositions, the local mole fractions of ethanol or acetone were higher than those of the mixtures and therefore the δx1,3 values were positive, which indicated that genistin and daidzein were preferentially solvated by the co-solvent. In these regions, daidzein and genistin were acting as a Lewis acid with ethanol or acetone molecules.  相似文献   

10.
Abstract  The viscosity and density of ternary mixtures of 1-butanol + 2-butanol + 1,3-butanediol and the binary systems 1-butanol + 2-butanol, 1-butanol + 1,3-butanediol, and 2-butanol + 1,3-butanediol were measured at 303.15 K and atmospheric pressure over the entire range of compositions. Excess molar volumes V E and viscosity deviations Δη were obtained from the experimental results for the binary and ternary systems and fitted to Redlich–Kister’s and Cibulka’s equations in terms of mole fractions. The results obtained for the viscosity of liquid mixtures were used to test the semi-empirical relations of Grunberg–Nissan, Hind, and the two-parameter McAllister, Kendall, and Frenkel equations. The experimental data for the ternary system and the constituting binaries are analyzed to discuss the nature and strength of intermolecular interactions in these mixtures. Graphical abstract     相似文献   

11.
The complex triplet potential energy surface for the reaction of HCNO with NH is investigated at the G3B3 level using the B3LYP/6-311++G(d,p), and QCISD/6-311++G(d,p) geometries. Various possible isomerization and dissociation pathways are probed. The initial association between HCNO and NH is found to be carbon to nitrogen attack leading to HNCHNO 2a, which can convert to 2b, 2c, and 2d. Subsequently, 1,4-H-shift of 2a to form NCHNOH 3a followed by dissociation to P 2 (1HCN + 3HON) is the most feasible pathway. Much less competitively, 2d undergoes successive 1,3-H-shift and C-N cleavage to form HNCNOH 8b, and then to product P 3 (1HNC + 3HON), the second feasible pathway. 8b can alternatively isomerize to 8c followed by N–O bond rupture to generate P 6 (2OH + 2HNCN), the lesser followed feasible pathway. In addition, 2b takes continuously 1,3- and 1,2-H-shift to form NC(H)NHO 6a, then to ONHCNH 7a which can convert to 7b. Eventually, 7b may take C-N bond fission to produce P 5 (1HNC + 3HNO), the least feasible pathway. The present paper may be helpful for future experimental identification of the product distributions for the title reaction, and may be helpful to deeply understand the mechanism of the title reaction.  相似文献   

12.
Liquid–liquid equilibrium (LLE) data were measured at 298.15 K and atmospheric pressure for the quaternary [water + methanol + ethyl methyl carbonate (EMC) + octane] and (water + ethanol + EMC + octane) systems, and relevant ternary (water + methanol + EMC), (water + ethanol + EMC), (water + ethanol + octane) and (water + EMC + octane) systems. The modified and extended UNIQUAC activity coefficient models were employed to correlate the experimental LLE results, which were also correlated by the Othmer–Tobias equation. The separation factors were calculated from the LLE data. The extracting selectivity of methanol and ethanol from aqueous solutions using pure EMC and mixed solvent (EMC + octane) was studied.  相似文献   

13.
Thermal rate constants of the CH4 + O2 = CH3 + HO2 reaction were calculated from first principles using both the conventional transition state theory (TST) and canonical variational TST methods with correction from the explicit hindered rotation treatment. The CCSD(T)/aug-cc-pVTZ//BH&HLYP/aug-cc-pVDZ method was used to characterize the necessary potential energy surface along the minimum energy path. We found that the correction for hindered rotation treatment, as well as the re-crossing effects noticeably affect the rate constants of the title process. The calculated rate constants for both forward and reverse directions are expressed in the modified Arrhenius form as \(k_{\text{forward}}^{\text{CVT/HR}} = 2.157 \times 10^{ - 18} \times T^{2.412} \times \,\exp \,( - \frac{25812}{T})\) and \(k_{\text{reverse}}^{\text{CVT/HR}} = 1.375 \times 10^{ - 19} \times T^{2.183} \times \,{\kern 1pt} \exp \,\,(\frac{2032}{T})\) (cm3 molecule?1 s?1) for the temperature range of 300–2,500 K. Being in good agreement with literature data, the results provide solid basis information for the investigation of the entire alkane + O2 = alkyl radical + HO2 reaction class.  相似文献   

14.
The solid–liquid phase equilibria for the ternary system 2-methyl-4-nitroaniline + 2-methyl-6-nitroaniline + ethyl acetate was determined experimentally by the method of isothermal solution saturation at temperatures of (293.15, 303.15 and 313.15) K under the pressure of 101.2 kPa. Based on the obtained solubility data, the isothermal phase diagrams of the system were constructed. At each temperature, there are two pure solid phases formed, which correspond to pure 2-methyl-4-nitroaniline and pure 2-methyl-6-nitroaniline, which was confirmed by Schreinemakers’ wet residue method and X-ray powder diffraction. The crystallization regions of pure 2-methyl-4-nitroaniline and pure 2-methyl-6-nitroaniline increased with decreasing temperature. The crystalline region of 2-methyl-4-nitroaniline was larger than that of 2-methyl-6-nitroaniline at a fixed temperature. The solubility data were correlated with the NRTL and Wilson models. The values of the root-mean-square deviations are 5.01 × 10?3 for the NRTL model, and 6.43 × 10?3 for the Wilson model. The solid–liquid equilibria, phase diagrams and the thermodynamic models for the ternary system can provide the foundation for separating 2-methyl-6-nitroaniline or 2-methyl-4-nitroaniline from its mixtures.  相似文献   

15.
Liquid–liquid equilibria (LLE) and tie-line data of systems containing 1-butanol, water and NaCl, Na2SO4, NH4Cl or (NH4)2SO4 were investigated at 303.15, 313.15 and 323.15 K and atmospheric pressure. The salt decreases mutual solubilities of these two solvents leading to a higher degree of phase separation at equilibrium. The effect is more pronounced at high salt concentration. Temperature in the studied range had a minor effect on LLE behavior of these mixtures. Experimental data were correlated using a modified extended UNIQUAC model. Satisfactory agreement between the calculated and measured mass fractions of the components was achieved.  相似文献   

16.
Full-dimensional time-dependent wave packet calculations were made to study the \(\hbox{OH}+\hbox{CO} \rightarrow \hbox{H}+\hbox{CO}_2\) reaction on the Lakin–Troya–Schatz–Harding potential energy surface. Because of the presence of deep wells supporting long-lived collision complex, one needs to propagate the wave packet up to 450,000 a.u. of time to fully converge the total reaction probabilities. Our calculation revealed that the CO bond was substantially excited vibrationally in the complex wells, making it necessary to include sufficient CO vibration basis functions to yield quantitatively accurate results for the reaction. We calculated the total reaction probabilities from the ground initial state and two vibrationally excited states for the total angular momentum J = 0. The total reaction probability for the ground initial state is quite small in magnitude with many narrow and overlapping resonances due to the small complex-formation reaction probability and small probability for complex decaying into product channel. Initial OH vibrational excitation considerably enhances the reactivity because it enhances the probability for complex decaying into product channel, while initial CO excitation has little effects on the reactivity. We also calculated the reaction probabilities for a number of J > 0 states by using the centrifugal sudden approximation. By doing some calculations with multiple K-blocks included, we found that the centrifugal sudden approximation can be employed to calculate the rate constant for the reaction rather accurately. The calculated rate constants only agree with experimental measurements qualitatively, suggesting more theoretical studies be carried out for this prototypical complex-formation four-atom reaction.  相似文献   

17.
The solubility of nimodipine (NMD) in several PEG 400 + ethanol mixtures was determined at five temperatures from 293.15 to 313.15 K. The thermodynamic functions of Gibbs energy, enthalpy and entropy of solution and of mixing were obtained by using the van't Hoff and Gibbs equations from these solubility data and drug properties of fusion. The highest solubility value was obtained in PEG 400 and the lowest in ethanol at all temperatures. A non-linear enthalpy–entropy relationship was observed from a plotof enthalpy vs. Gibbs energy of solution. Accordingly, the driving mechanism for NMD solubility in ethanol-rich mixtures is the enthalpy, whereas in PEG 400-rich mixtures thedriving mechanism is the entropy, although the molecular events involved are unclear.  相似文献   

18.
19.
The mechanisms of CH2SH with NO2 reaction were investigated on the singlet and triplet potential energy surfaces (PES) at the BMC-CCSD//B3LYP/6-311 + G(d,p) level. The result shows that the title reaction is more favourable on the singlet PES thermodynamically, and it is less competitive on the triplet PES. On the singlet PES, the initial addition of CH2SH with NO2 leads to HSCH2NO2 (IM2) without any transition state, followed by a concerted step involving C–N fission and shift of H atom from S to O giving out CH2S + trans-HONO, which is the major products of the title reaction. With higher barrier height, the minor products are CH2S + HNO2, formed by a similar concerted step from the initial adduct HSCH2ONO (IM1). The direct abstraction route of H atom in SH group abstracted by O atom might be of some importance. It starts from the addition of the reactants to form a weak interaction molecular complex (MC3), subsequently, surmounts a low barrier height leading to another complex (MC2), which gives out CH2S + trans-HONO finally. Other direct hydrogen abstraction channels could be negligible with higher barrier heights and less stable products.  相似文献   

20.
The molar excess enthalpies of eight systems of butylamines + propanols were determined at 298.15 K using a twin-microcalorimeter. All excess enthalpies were exothermic and large. An equilibrium constant K 1 expressed in terms of mole fractions and standard thermodynamic properties of formation (Δf H, Δf G, Δf S) of 1:1 complex were evaluated by ideal mixtures of monomeric molecules and their associated complexes. Concentration dependence of the FT-Raman spectrum showed systematic changes of bands. Spectroscopic considerations based on this and ab initio calculations on molecules were performed at the Mp2/6-311G(d,p) level of theory. Interaction energies between butylamine and propanol were calculated by the supermolecular and NBO methods. The results were discussed with previous results to clarify the steric and positional effect of the amino and hydroxyl group.  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