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 31 毫秒
1.
The solubility of sodium 3-sulfobenzoate in binary (sodium chloride + water), (sodium sulfate + water), and (ethanol + water) solvent mixtures was measured at elevated temperatures from (278.15 to 323.15) K by a steady-state method. The results of these experiments were correlated by a modified Apelblat equation. The dissolution enthalpy and entropy of sodium 3-sulfobenzoate in aqueous solutions of different mole fraction were obtained.  相似文献   

2.
The solubility data of sodium 4-nitrotoluene-2-sulfonate (NTSNa) in aqueous organic solutions (propanol + water) and (ethylene glycol + water) were measured at temperatures ranging from (290 to 351) K using a dynamic method. The mole fraction of water in solvent mixtures ranged from 0 to 0.8. The solubility values are correlated with the electrolyte non-random two-liquid (E-NRTL) model. From the results obtained, the E-NRTL model provides a satisfactory mathematical representation of the experimental results for the (NTSNa + propanol + water) system and an unsatisfactory result for the (NTSNa + ethylene glycol + water) system. Thus, the modified Apelblat model is applied to describe the (NTSNa + ethylene glycol + water) system also. The calculated (solid + liquid) equilibrium temperatures with the modified Apelblat model are in good agreement with the experimental results. The root-mean-square deviations of solubility temperature varied from (0.08 to 0.94) K for two models. The effect of different aqueous organic solutions on the reaction of oxidation 4-nitrotoluene-2-sulfonic acid (NTS) to 4,4′-dinitrostilbene-2,2′-disulfonic acid (DNS) was discussed.  相似文献   

3.
Density ρ, viscosity η, and refractive index nD, values for (tetradecane + benzene, + toluene, + chlorobenzene, + bromobenzene, + anisole) binary mixtures over the entire range of mole fraction have been measured at temperatures (298.15, 303.15, and 308.15) K at atmospheric pressure. The speed of sound u has been measured at T = 298.15 K only. Using these data, excess molar volume VE, deviations in viscosity Δη, Lorentz–Lorenz molar refraction ΔR, speed of sound Δu, and isentropic compressibility Δks have been calculated. These results have been fitted to the Redlich and Kister polynomial equation to estimate the binary interaction parameters and standard deviations. Excess molar volumes have exhibited both positive and negative trends in many mixtures, depending upon the nature of the second component of the mixture. For the (tetradecane + chlorobenzene) binary mixture, an incipient inversion has been observed. Calculated thermodynamic quantities have been discussed in terms of intermolecular interactions between mixing components.  相似文献   

4.
Equilibria of EuO dissolution and dissociation in molten (NaBr + NaI) mixtures of 0.77:0.23 and 0.31:0.69 compositions at T = 973 K were studied by potentiometric titration method using Pt(O2)|ZrO2(Y2O3) indicator electrode. The solubility product indices of EuO are (7.81 ± 0.08) and (8.43 ± 0.16) in the melts of 0.77:0.23 and 0.31:0.69 compositions. The corresponding dissociation constant indices are (4.96 ± 0.04) and (5.54 ± 0.06), respectively (all the parameters are in molality). Non-dissociated EuO is the prevailing form in all the saturated solutions of europium monoxide. The decrease of the iodide ion concentration in the melts results in strengthening of EuO dissociation that is explained by introduction of harder Pearson’s base (Br) in sodium iodide melt. In its turn this increases the fixation degree of Eu2+ in mixed halide complexes. The total solubility of EuO decreases going from NaI melt to the (bromide + iodide) mixtures that is caused by the decrease of ‘physical’ solubility of non-dissociated oxide which occupies hollow spaces of enough large size in the ionic solvents. The quantity of these hollow spaces diminishes at the sequential Br  I substitution.  相似文献   

5.
The mean activity coefficients of NaCl in (sodium chloride  +  sodium bicarbonate  +  water) were determined experimentally in the temperature range 293.15 K to 308.15 K at four NaHCO3molality fractions (0.1, 0.3, 0.5, and 0.7). The measurements were made with an electrochemical cell, using a Na + glass ion-selective electrode and a Cl  solid-state ion-selective electrode. The experimental values reported by Butler and Huston are found to be higher than those calculated from the Pitzer equation using the existing parameters while the experimental results of this work are close to the calculated values, up to an NaHCO3molality fraction of 0.5. At the NaHCO3molality fraction of 0.7, the experimental data are much lower than the calculated values, implying that the interference of HCO3  on the Na + glass ion-selective electrode can only be neglected up to a molality fraction of NaHCO3of 0.5, an observation which is consistent with that of Butler and Huston.  相似文献   

6.
The potential differences E of the cells Pt|H2|H2Ph(m1)  +  KHPh(m2)  +  KCl(m3) in Z|AgCl|Ag and Pt|H2|H2Ph(m1)  +  KHPh(m2)  +  KCl(m3) in Z|Hg2Cl2|Hg have been measured at T =  298.15 K in mixtures Z =  (W + S) of water (W) with cosolvents S =  propylene carbonate (PC) and S =  ethylene carbonate (EC), to determine the first ionization constants K of the o -phthalic acid H2Ph(benzene-1,2-dicarboxylic acid), which are indispensable for the determination of primary pH-metric standards based on the potassium hydrogen phthalate buffer (KHPh) in such solvent mixtures. The value of K is seen to decrease progressively with increasing mass fraction wsof the organic cosolvent, as with all of the other cosolvents studied earlier, but no simple relationship with the cosolvent permittivity is discernible. Since the required values of the standard potential difference Eoof the second cell were hitherto missing, they have now been obtained based on potential difference measurements of the cell Pt|H2|HCl(m) in Z|Hg2Cl2|Hg. The primary medium effect (EWo  EZo, by Owen’s definition) upon HCl in water-rich mixtures Z is seen to increase linearly with increasing ws, as in earlier investigations. In this comparative context, the slope of the primary medium effect against wsplots for the aprotic cosolvents increases regularly with decreasing permittivity, whereas for the protic (alcoholic) cosolvents the slope is ill-defined.  相似文献   

7.
Densities, ρ speeds of sound, u and dynamic viscosities, η of the ternary mixtures {dimethyl carbonate (DMC) + methanol + ethanol} and (dimethyl carbonate + methanol + hexane) were gathered at T = (293.15, 298.15, 308.15, and 313.15) K. From experimental data viscosity deviations, Δη of the ternary mixtures were evaluated. These results have been correlated using the Cibulka equation. The fitting parameters and the standard deviations of the ternary viscosity deviations are given. UNIFAC-VISCO group contribution method was used to predict the dynamic viscosities of the ternary mixtures at several temperatures.  相似文献   

8.
The thermodynamic properties ofZn5(OH)6(CO3)2 , hydrozincite, have been determined by performing solubility and d.s.c. measurements. The solubility constant in aqueous NaClO4media has been measured at temperatures ranging from 288.15 K to 338.15 K at constant ionic strength (I =  1.00 mol · kg  1). Additionally, the dependence of the solubility constant on the ionic strength has been investigated up to I =  3.00 mol · kg  1NaClO4at T =  298.15 K. The standard molar heat capacity Cp, mofunction fromT =  318.15 K to T =  418.15 K, as well as the heat of decomposition of hydrozincite, have been obtained from d.s.c. measurements. All experimental results have been simultaneously evaluated by means of the optimization routine of ChemSage yielding an internally consistent set of thermodynamic data (T =  298.15 K): solubility constant log * Kps 00 =  (9.0  ±  0.1), standard molar Gibbs energy of formationΔfGmo {Zn5(OH)6(CO3)2 }  =  (  3164.6  ±  3.0)kJ · mol  1, standard molar enthalpy of formation ΔfHmo{Zn5(OH)6(CO3)2 }  =  (  3584  ±  15)kJ · mol  1, standard molar entropy Smo{Zn5(OH)6(CO3)2 }  =  (436  ±  50)J · mol  1· K  1and Cp,mo / (J · mol  1· K  1)  =  (119  ±  11)  +  (0.834  ±  0.033)T / K. A three-dimensional predominance diagram is introduced which allows a comprehensive thermodynamic interpretation of phase relations in(Zn2 +  +  H2O  +  CO2) . The axes of this phase diagram correspond to the potential quantities: temperature, partial pressure of carbon dioxide and pH of the aqueous solution. Moreover, it is shown how the stoichiometric composition{n(CO3) / n(Zn)} of the solid compoundsZnCO3 and Zn5(OH)6(CO3)2can be checked by thermodynamically analysing the measured solubility data.  相似文献   

9.
(Solid + liquid) equilibria (SLE) have been measured for naphthalene + o-dichlorobenzene, + m-dichlorobenzene, and + p-dichlorobenzene using differential scanning calorimetry (DSC) over the whole concentration range. It was found that the phase diagram of (naphthalene + m-dichlorobenzene) is of a simple eutectic type with the eutectic point at 244.85 K and 0.058 mole fraction of naphthalene, the phase diagram of (naphthalene + p-dichlorobenzene) is of a simple eutectic type with the eutectic point at 302.85 K and 0.390 mole fraction of naphthalene and in the system of (naphthalene + o-dichlorobenzene), a 1:1 incongruently melting compound is formed and that the phase diagram show a eutectic and a peritectic, the eutectic point is at 232.55 K and 0.130 mole fraction of naphthalene, the peritectic point at 250.15 K and 0.077 mole fraction of naphthalene. Furthermore, the activity coefficients of components in mixtures of (naphthalene + m-dichlorobenzene) and (naphthalene + p-dichlorobenzene) have been correlated by the Scatchard–Hildebrand solubility parameter expression. This approach offers a useful procedure for estimating with good accuracy.  相似文献   

10.
A new set of values for the heat capacity of aqueous mixtures of piperazine (PZ) and n-methyldiethanolamine (MDEA) at different concentrations and temperatures are reported in this paper. The differential scanning calorimetry technique was used to measure the property over the range T = 303.2 K to T = 353.2 K for mixtures containing 0.60 to 0.90 mole fraction water with 15 different concentrations of the system (PZ + MDEA + H2O). Heat capacity for four concentrations of the binary system (PZ + MDEA) was also measured. A Redlich–Kister-type equation was adopted to estimate the excess molar heat capacity, which was used to predict the value of the molar heat capacity at a particular concentration and temperature, which would then be compared against the measured value. A total of 165 data points fit into the model resulted in a low overall average absolute deviation of 4.6% and 0.3% for the excess molar heat capacity and molar heat capacity, respectively. Thus, the results presented here are of acceptable accuracy for use in engineering process design.  相似文献   

11.
Total vapour pressures and excess molar volumes, measured at the temperature 313.15 K, are reported for three binary mixtures (2-pyrrolidone + water), (2-pyrrolidone + methanol) and (2-pyrrolidone + ethanol). The results are compared with previously obtained data for binary mixtures (amide + A), where amide=N-methylformamide, N,N-dimethylformamide and N-methylacetamide, and A= water, methanol, and ethanol.  相似文献   

12.
Vapour pressures of (1-chlorobutane  +  1-butanol, or 2-methyl-2-propanol) at several temperatures between T =  278.15 and T =  323.15 K were measured by a static method. Reduction of the vapour pressures to obtain activity coefficients and excess molar Gibbs energies was carried out by fitting the vapour pressure data to the Redlich–Kister equation according to Barker’s method. For (1-chlorobutane  +  2-methyl-2-propanol) azeotropic mixtures with a minimum boiling temperature were observed over the whole temperature range.  相似文献   

13.
(Liquid + liquid) equilibrium (LLE) data for ternary systems: (heptane + benzene + N-formylmorpholine), (heptane + toluene + N-formylmorpholine), and (heptane + xylene + N-formylmorpholine) have been determined experimentally at temperatures ranging from 298.15 K to 353.15 K. Complete phase diagrams were obtained by determining solubility and tie-line data. Tie-line compositions were correlated by Othmer–Tobias and Bachman methods. The universal quasichemical activity coefficient (UNIQUAC) and the non-random two liquids equation (NRTL) were used to predict the phase equilibrium in the system using the interaction parameters determined from experimental data. It is found that UNIQUAC and NRTL used for LLE could provide a good correlation. Distribution coefficients, separation factors, and selectivity were evaluated for the immiscibility region.  相似文献   

14.
Precise excess volumes of mixing measurements at T = 313.15 K are reported over the whole composition range for binary mixtures: (N,N-dimethylacetamide + water), (N,N-dimethylacetamide + methanol), (N,N-dimethylacetamide + ethanol) and for the ternary mixtures (N,N-dimethylacetamide + methanol + water) and (N,N-dimethylacetamide + ethanol + water). For all the systems, large negative deviations from ideality are observed. The binary results have been fitted using the Redlich–Kister type polynomial. The possibility of predicting the ternary results from the binary ones was examined.  相似文献   

15.
Coexistence curves of ( T, n), ( T, ϕ), and ( T, Ψ), where n, ϕ, and Ψ are the refractive index, volume fraction and effective volume fraction ψ = ϕ / {ϕ +  [(1   ϕ )ϕc / (1   ϕc )]}, respectively, for ternary microemulsion systems of {water  + n -nonane  +  sodium di(2-ethyl-1-hexyl)sulphosuccinate} have been determined at temperatures within 8.7 K above the critical temperature by measurements of refractive index at constant pressure and a constant molar ratio of water to sodium di(2-ethyl-1-hexyl)sulphosuccinate. The critical exponent β deduced from ( T,n ), ( T, ϕ), and ( T, Ψ) coexistence curves was found consistent with nonmonotonic crossover observed in all aqueous ionic solutions. The values of β deduced from the experimental data in the range of 1 K above Tcwere consistent with the universality class of three-dimensional Ising-like systems. The coexistence curves have been interpreted by a combination of the Wegner expansion and the rectilinear diameter. The present results indicate that the molar mass dependence of critical amplitudes, we proposed recently, is valid for microemulsion systems.  相似文献   

16.
The bromide minerals solubility in the mixed system (m1NaBr + m2MgBr2)(aq) have been investigated at T = 323.15 K by the physico-chemical analysis method. The equilibrium crystallization of NaBr·2H2O(cr), NaBr(cr), and MgBr2·6H2O(cr) has been established. The solubility-measurements results obtained have been combined with all other experimental equilibrium solubility data available in literature at T = (273.15 and 298.15) K to construct a chemical model that calculates (solid + liquid) equilibria in the mixed system (m1NaBr + m2MgBr2)(aq). The solubility modeling approach based on fundamental Pitzer specific interaction equations is employed. The model gives a very good agreement with bromide salts equilibrium solubility data. Temperature extrapolation of the mixed system model provides reasonable mineral solubility at high temperature (up to 100 °C). This model expands the previously published temperature variable sodium–potassium–bromide and potassium–magnesium–bromide models by evaluating sodium–magnesium mixing parameters. The resulting model for quaternary system (Na + K + Mg + Br + H2O) is validated by comparing solubility predictions with those given in literature, and not used in the parameterization process. Limitations of the mixed solution models due to data insufficiencies at high temperature are discussed.  相似文献   

17.
Standard thermodynamic functions of transfer of naphthalene and 2-naphthoic acid from water to (water + ethanol) mixtures at T=298.15 K have been determined from solubility measurements at different temperatures. Standard free energies of transfer of both naphthalene and 2-naphthoic acid showed decreasing tendency with the increasing x(EtOH), and the standard entropy and enthalpy of transfer exhibited a change of double peaks with x(EtOH). The ΔtrG0 of 2-naphthoic acid decreased more rapidly than that of naphthalene when x(EtOH) < 0.746 and lower than that of naphthalene when x(EtOH) >0.746 at T=298.15 K. The double peaks in the curves of standard entropy and enthalpy of transfer illustrated that the microstructure of the series of mixed solvents of (water + ethanol) underwent a variable process from ordered to disordered and then from disordered to ordered. The results mean that there is a relatively ordered structure near x(EtOH)=0.13 in the (water + ethanol) solutions besides the existence of a clathrate structure in the water-rich region.  相似文献   

18.
The activity coefficient data were reported for (water  +  potassium chloride  + dl -valine) at T =  298.15 K and (water  +  sodium chloride  + l -valine) at T =  308.15 K. The measurements were performed in an electrochemical cell using ion-selective electrodes. The maximum concentrations of the electrolytes and the amino acids studied were 1.0 molality and 0.4 molality, respectively. The results of the activity coefficients of dl -valine are compared with the activity coefficients of dl -valine in (water  +  sodium chloride  + dl -valine) system obtained from the previous study. The results show that the presence of an electrolyte and the nature of its cation have a significant effect on the activity coefficient of dl -valine in aqueous electrolyte solutions.  相似文献   

19.
《Fluid Phase Equilibria》2006,239(2):146-155
This work reports the measured density, ρ, and viscosity, η, values of liquid mixtures of tetrahydrofuran (1) + 1-chlorobutane (2) + 2-butanol (3) at temperatures of 283.15, 298.15 and 313.15 K over a range of mole fractions and atmospheric pressure. Excess molar volume, VE, viscosity deviations, Δη, and excess free energies of activation of viscous flow, ΔG*E, have been calculated from experimental data and fitted to Cibulka, Singh et al. and Nagata and Sakura equations. The results were analyzed in terms of the molecular interaction between the components of the mixtures. Excess molar volumes and viscosity deviations were predicted from binary contributions using geometrical solution models, Tsao and Smith; Jacob and Fitzner; Kholer; Rastogi et al.; Radojkovic et al. Finally, experimental results are compared with those obtained by applying group-contribution method proposed by Wu.  相似文献   

20.
In this study, a new set of data for the equilibrium solubility of carbon dioxide in the amine solvent system that consists of triethanolamine (TEA), piperazine (PZ), and water is presented. Equilibrium solubility values were obtained at T = (313.2, 333.2, and 353.2) K and pressures up to 153 kPa using the vapour-recirculation equilibrium cell. The TEA concentrations in the considered ternary (solvent) mixture were (2 and 3) kmol · m?3 and those of PZ’s were (0.5, 1.0, and 1.5) kmol · m?3. The solubility data (CO2 loading in the amine solution) obtained were correlated as a function of CO2 partial pressure, system temperature, and amine composition via the modified Kent–Eisenberg model. Results showed that the model applied is generally satisfactory in representing the CO2 absorption into mixed aqueous solutions of TEA and PZ.  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