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 734 毫秒
1.
In this work, the effect of solvent (methanol, CH3OH) molecules on proton transfer (PT) between ammonia and hydrogen halides was studied. We performed MP2 and B3LYP calculations on HX–NH3–(CH3OH) n clusters for three hydrogen halides, HF, HCl, and HBr, with the number of methanol molecules varying from none to three (n = 0–3). The results showed that stepwise association of methanol molecules with the gas-phase complex can eventually facilitate ionization within the complex, producing the $ {\text{NH}}_{4}^{ + } {\text{X}}^{ - } - \left( {{\text{CH}}_{ 3} {\text{OH}}} \right)_{\text{n}} $ cluster. We found that PT occurs on addition of from one (for HBr) to three (for HF) methanol molecules. The interaction energy $ E_{\text{int}} $ and $ \Updelta E_{\text{add}} $ for the complexes were calculated and basis set superposition error (BSSE) correction was also performed. Atoms-in-molecule and natural-bond-orbital analysis were used to study the properties of the hydrogen bonds in the complexes.  相似文献   

2.
The densities, ρ, of binary mixtures of butyl acrylate with 1-butanol, 2-butanol, 2-methyl-1-propanol, and 2-methyl-2-propanol, including those of the pure liquids, were measured over the entire composition range at temperatures of (288.15, 293.15, 298.15, 303.15, 308.15, 313.15, and 318.15) K and atmospheric pressure. From the experimental data, the excess molar volume $ V_{\text{m}}^{\text{E}} $ V m E , partial molar volumes $ \overline{V}_{\text{m,1}} $ V ¯ m,1 and $ \overline{V}_{\text{m,2}} $ V ¯ m,2 , and excess partial molar volumes $ \overline{V}_{\text{m,1}}^{\text{E}} $ V ¯ m,1 E and $ \overline{V}_{\text{m,2}}^{\text{E}} $ V ¯ m,2 E , were calculated over the whole composition range as were the partial molar volumes $ \overline{V}_{\text{m,1}}^{^\circ } $ V ¯ m,1 ° and $ \overline{V}_{\text{m,2}}^{^\circ } $ V ¯ m,2 ° , and excess partial molar volumes $ \overline{V}_{\text{m,1}}^{{^\circ {\text{E}}}} $ V ¯ m,1 ° E and $ \overline{V}_{\text{m,2}}^{{^\circ {\text{E}}}} $ V ¯ m,2 ° E , at infinite dilution,. The $ V_{\text{m}}^{\text{E}} $ V m E values were found to be positive over the whole composition range for all the mixtures and at each temperature studied, indicating the presence of weak (non-specific) interactions between butyl acrylate and alkanol molecules. The deviations in $ V_{\text{m}}^{\text{E}} $ V m E values follow the order: 1-butanol < 2-butanol < 2-methyl-1-propanol < 2-methyl-2-propanol. It is observed that the $ V_{\text{m}}^{\text{E}} $ V m E values depend upon the position of alkyl groups in alkanol molecules and the interactions between butyl acrylate and isomeric butanols decrease with increase in the number of alkyl groups at α-carbon atom in the alkanol molecules.  相似文献   

3.
Densities and refractive indices of mixing of olive oil with the alkanols: methanol, ethanol, 1-propanol, 2-propanol and 1-butanol, have been measured as a function of the composition at T = 298.15 K. Excess molar volumes, $ V_{\text{m}}^{\text{E}} $ , and deviation in refractive index, Δn D, were calculated and correlated by a Redlich–Kister type function, to derive the coefficients and estimate the standard error. For mixtures of olive oil with alkanols, $ V_{\text{m}}^{\text{E}} $ is positive, except with ethanol and methanol where a sigmoidal variation is observed. Δn D is positive over the entire range of mole fraction. The effect of chain length of the alkanols on the excess molar volumes and deviation in refractive index of the mixtures with olive oil are discussed.  相似文献   

4.
Standard transfer Gibbs energies, $ \Updelta_{\text{tr}} G^{^\circ } $ , of a series of homologues α-amino acids have been evaluated by determining the solubility of glycine, alanine, amino butyric acid and norvaline gravimetrically at 298.15 K. Standard entropies of transfer, $ \Updelta_{\text{tr}} S^{^\circ } $ , of the amino acids have also been evaluated by extending the solubility measurement to five equidistant temperatures ranging from 288.15 to 308.15 K. The chemical contributions $ \Updelta_{\text{tr,ch}} G^{^\circ } (i) $ of α-amino acids, as obtained by subtracting theoretically computed contributions to $ \Updelta_{\text{tr}} G^{ \circ } $ due to cavity and dipole–dipole interaction effects from the corresponding experimental $ \Updelta_{\text{tr}} G^{ \circ } $ , are indicative of the superimposed effect of increased basicity and dispersion and decreased hydrophobic hydration (hbh) in DMF–water solvent mixtures as compared to those in water, while, in addition, $ T\Updelta_{\text{tr,ch}} S^{^\circ } (i) $ is guided by structural effects. The computed chemical transfer energies of the –CH2– group, $ \Updelta_{\text{tr,ch}} P^{^\circ } $ (–CH2–) [P = G or S] as obtained by subtracting the value of lower homologue from that of immediately higher homologue, are found to change with composition indicating involvement of several opposing factors in the calculation of the chemical interactions. The $ \Updelta_{\text{tr,ch}} G^{^\circ } $ (–CH2–) values are found to be guided by the decreased hydrophobic effect in DMF–water mixtures, and are indicative of the nature of the three dimensional structure of the aquo-organic solvent system around each solute.  相似文献   

5.
The densities, ρ, and viscosites, η, of mixtures of propanoic acid with equimolar mixtures of N,N-dimethyl formamide + methanol/ethanol/1-propanol, over the entire composition range of propanoic acid and including the pure liquids, have been measured at the temperatures T/K = 303.15, 313.15, and 323.15. From this experimental data, the excess molar volume, $ V_{\text{m}}^{\text{E}} $ , deviation in viscosity, Δη, and excess Gibbs energy of activation of viscous flow, ΔG *E, have been determined at all three temperatures. The influence of temperature on these mixtures has been studied in terms of molecular interactions. The calculated deviation and excess parameters have been fitted to a Redlich–Kister type polynomial and the corresponding standard deviations were also evaluated. Negative values of $ V_{\text{m}}^{\text{E}} $ and positive values of Δη and ΔG *E are observed at all temperatures over the entire composition range in the mixtures studied. The observed negative and positive values of various excess and deviation parameters are attributed to the existence of strong interactions, like dipole–dipole interactions, H-bonding between the carbonyl group of acid molecules and hydroxyl group of alcohol groups, geometrical fitting of smaller molecules into the voids created by larger molecules in the liquid mixtures. The strength of these interactions in the mixtures was found to decrease with the rise in temperature and increase with an increase of chain length of the alcohols. The derived partial molar volumes and excess partial molar volumes also support the $ V_{\text{m}}^{\text{E}} $ results. The experimental viscosity data of all of these liquid mixtures have been correlated with four viscosity models, those of Grunberg and Nissan, Hind et al., Katti and Chaudhri, and Heric and Brewer. The Katti and Chaudhri model was found to be in good agreement with the experimental values.  相似文献   

6.
Measurement of the transport of water with respect to the second solvent component in a binary aqueous mixture gives the Washburn number, $ w_{\text{W}} = (n_{\text{W}} )_{ + } t_{ + } - (n_{\text{W}} )_{ - } t_{ - } $ , in a transport number determination, where the ions move in opposite directions, and give the Erdey–Grúz number, $ \Upsigma n_{\text{W}} = (n_{\text{W}} )_{ + } + (n_{\text{W}} )_{ - } $ , in a diffusion experiment, where the ions move in the same direction. Here n W and t are the number of water molecules and transport number, respectively, of the anion or cation. Combination of the results of these two experiments allows unambiguous determination of values for the solvent transport numbers, $ n_{\text{W}} $ , of the individual ions. While the values of $ n_{\text{W}} $ depend on the cosolvent, at high dilutions of the second component the highest value of $ n_{\text{W}} $ found, $ N_{\text{W}} $ , should approach the number of water molecules transported by the ion in pure water, $ N_{\text{W}}^{0} $ . New data for alkali-metal, alkaline-earth metal, hydrogen and halide ions in dilute mixtures of t-butyl alcohol with water are presented. Values of $ N_{\text{W}} $ rounded to whole numbers thus found are: 12 (Li+), 10 (Na+), 6 (K+), 5 (Rb+), 5 (Cs+), 1 (H+), 13 (Ca2+), 16 (Sr2+) and 15 (Ba2+). Factors influencing preferential solvation are briefly discussed. Detailed recalculations of $ n_{\text{W}} $ in the raffinose–water system from literature data also allows resolution of a problem with the Onsager Relations.  相似文献   

7.
The densities, ρ, and viscosities, η, of binary mixtures of ethylene glycol with formamide, N,N-dimethyl formamide and N,N-dimethyl acetamide, have been measured over the entire composition range at 308.15 K. From this experimental data, excess molar volume, \( V_{\text{m}}^{\text{E}} \) , deviation in viscosity, Δη, and excess Gibbs free energy of activation of viscous flow, \( \Delta G^{{ * {\text{E}}}}, \) have been determined. Negative values of \( V_{\text{m}}^{\text{E}} \) , Δη, and \( \Delta G^{{ * {\text{E}}}} \) are observed over the entire composition range in the mixtures studied. The observed negative values of various excess and deviation parameters are attributed to the existence of strong interactions, like dipole–dipole interactions, H-bonding between the carbonyl group of amide molecules, and hydroxyl group of glycol molecules, geometrical fitting of smaller molecules into the voids created by larger molecules in the liquid mixtures. The excess properties have been fitted to Redlich–Kister-type polynomial, and the corresponding standard deviations have been calculated. The derived partial molar volumes and excess partial molar volumes also support the \( V_{\text{m}}^{\text{E}} \) results. The experimental viscosity data of all of these liquid mixtures have been correlated with four viscosity models.  相似文献   

8.
The effects of bismuth (Bi), antimony (Sb) and strontium (Sr) additions on the characteristic parameters of the evolution of aluminium dendrites in a near eutectic Al–11.3Si–2Cu–0.4Fe alloy during solidification at different cooling rates (0.6–2 °C) were investigated by computer-aided cooling curve thermal analysis (CA-CCTA). Nucleation temperature ( $ T_{\text{N}}^{{\alpha {\text{ - Al}}}} $ ) is defined with a new approach based on second derivative cooling curve. The results showed that $ T_{\text{N}}^{{\alpha {\text{ - Al}}}} $ increased with increasing cooling rate but both the growth temperature ( $ T_{\text{G}}^{{\alpha {\text{ - Al}}}} $ ) and the coherency temperature (T DCP) decreased. Increase in the temperature difference for dendrite coherency ( $ T_{\text{N}}^{{\alpha {\text{ - Al}}}} - T_{\text{DCP}} $ ) with increasing cooling rate indicate a wider range of temperature before the dendrite can impinge on each other and higher fraction solid ( $ f_{\text{S}}^{\text{DCP}} $ ). Additions of Bi, Sb and Sr to the base alloy produced only a minor effect on $ T_{\text{N}}^{{\alpha {\text{ - Al}}}} $ . Additions of Bi and Sb resulted in an increase in fraction solid and an increase of 30 % in the value of $ T_{\text{N}}^{{\alpha {\text{ - Al}}}} \, - \,T_{\text{G}}^{{\alpha {\text{ - Al}}}} $ to almost 13 °C.  相似文献   

9.
pH-spectrophotometric titration data were used to determine the mixed dissociation constants of sildenafil citrate at different ionic strengths I at temperatures of 288.15, 298.15 and 310.15?K, with the use of two different multiwavelength and multivariate treatments of spectral data, SPECFIT32 and SQUAD(84) nonlinear regression analyses, and INDICES factor analysis. The reliability of the dissociation constants of this drug was proven with goodness-of-fit tests of the pH-spectra. The thermodynamic dissociation constants $ {\text{p}}K_{{{\text{a}},i}}^{\text{T}} $ were estimated by a nonlinear regression of (pK a , I) data using the Debye-Hückel equation: $ {\text{p}}K_{{{\text{a}}, 1}}^{\text{T}} $ ?=?2.79 (1), 3.03 (3) and 3.53 (1); $ {\text{p}}K_{{{\text{a}}, 2}}^{\text{T}} $ ?=?4.97 (2), 5.23 (2) and 5.34 (1); $ {\text{p}}K_{{{\text{a}}, 3}}^{\text{T}} $ ?=?8.14 (2), 7.93 (1) and 7.47 (1); $ {\text{p}}K_{{{\text{a}}, 4}}^{\text{T}} $ ?=?9.47 (2), 9.30 (1) and 9.13 (4); and $ {\text{p}}K_{{{\text{a}}, 5}}^{\text{T}} $ ?=?10.73 (5), 10.75 (3) and 10.79 (5) at T?=?288.15, 298.15 and 310.15?K, respectively, where the numbers in parentheses are the standard deviations in the last significant digits. Concurrently, the experimentally determined five thermodynamic dissociation constants are in a good agreement with their computational prediction of the SPARC program based on knowledge of the chemical structures. The factor analysis of spectra in the INDICES program predicts the correct number of light-absorbing components when the instrumental error is known and when the signal-to-error ratio SER is higher than 10. A rough estimation of the dissociation enthalpies ??H 0 (kJ·mol?1) and entropies ??S 0 (J·K?1·mol?1) has been obtained from the temperature variation of the thermodynamic dissociation constants by means of the van??t Hoff equation.  相似文献   

10.
Densities, ??, and viscosities, ??, of binary mixtures of 2-methyl-2-propanol with acetone (AC), ethyl methyl ketone (EMK) and acetophenone (AP), including those of the pure liquids, were measured over the entire composition range at 298.15, 303.15 and 308.15?K. From these experimental data, the excess molar volume $V_{\mathrm{m}}^{\mathrm{E}}$ , deviation in viscosity ????, partial and apparent molar volumes ( $\overline{V}_{\mathrm{m},1}^{\,\circ }$ , $\overline{V}_{\mathrm{m},2}^{\,\circ }$ , $\overline{V}_{\phi ,1}^{\,\circ}$ and $\overline{V}_{\phi,2}^{\,\circ} $ ), and their excess values ( $\overline{V}_{\mathrm{m},1}^{\,\circ \mathrm{E}}$ , $\overline{V}_{\mathrm{m,2}}^{\,\circ \mathrm{ E}}$ , $\overline {V}_{\phi \mathrm{,1}}^{\,\circ \mathrm{ E}}$ and $\overline{V}_{\phi \mathrm{,2}}^{\,\circ \mathrm{ E}}$ ) of the components at infinite dilution were calculated. The interaction between the component molecules follows the order of AP > AC > EMK.  相似文献   

11.
The densities, ρ 123, and speeds of sound, u 123, of ternary o-toluidine (OT, 1) + tetrahydropyran (THP, 2) + pyridine (Py) or benzene or toluene (3) mixtures have been measured as a function of composition at 298.15, 303.15 and 308.15 K. Values of the excess molar volumes, $ V_{123}^{\text{E}} , $ and excess isentropic compressibilities, $ (\kappa_{\text{S}}^{\text{E}} )_{123} , $ of the studied mixtures have been determined by employing the measured experimental data. The observed thermodynamic properties were fitted with the Redlich–Kister equation to determine adjustable ternary parameters and standard deviations. The $ V_{123}^{\text{E}} $ and $ (\kappa_{\text{S}}^{\text{E}} )_{123} $ values were also analyzed in terms of Graph theory. It was observed that Graph theory correctly predicts the sign as well as magnitude of $ V_{123}^{\text{E}} $ and $ (\kappa_{\text{S}}^{\text{E}} )_{123} $ values of the investigated mixtures. Analysis of the data suggests strong interactions and a more close packed arrangement in OT (1) + THP (2) + Py (3) mixtures as compared to those of the OT (1) + THP (2) + benzene (3) or toluene (3) mixtures. This may be due to the presence of a nitrogen atom in Py which results in stronger interactions for the OT:THP molecular entity as compared to those with benzene or toluene.  相似文献   

12.
The dispersive free energy and acid–base forces of epoxidized soybean oil (ESO) were determined by inverse gas chromatography (IGC). Eight non-polar and polar solvents were used as the probes in the temperature range between 303.15 and 343.15 K. The IGC characterization encompassed the adsorption thermodynamic parameters, including the standard enthalpy ( $ \Updelta H_{\text{a}}^{\text{s}} $ ) and the free energy change of adsorption ( $ \Updelta G_{\text{a}}^{\text{s}} $ ), using the retention time of different non-polar and polar probes at the infinite dilution region. Surface characterization showed that ESO has low $ r_{\text{s}}^{\text{d}} $ value, even at 303.15 K, and is a Lewis amphoteric material with predominantly basic character, as confirmed by the Lewis acidity and basicity constants K a and K b, respectively.  相似文献   

13.
Density (ρ), viscosity (η), and speed of sound (U) values for the binary mixture systems of methyl benzoate + 2-propanol and ethyl benzoate + 2-propanol including those of pure liquids were measured over the entire mole fraction range at five different temperatures (303.15, 308.15, 313.15, 318.15, and 323.15) K. From these experimentally determined values, various thermo-acoustic parameters such as excess isentropic compressibility $ \left( {K_{\text{s}}^{\text{E}} } \right) $ , excess molar volume (V E) and excess free length $ \left( {L_{\text{f}}^{\text{E}} } \right) $ , excess Gibb’s free energy (ΔG *E), and excess enthalpy (H E) have been calculated. The excess functions have been fitted to the Redlich–Kister type polynomial equation. The deviations for excess thermo-acoustic parameters have been explained on the basis of the intermolecular interactions present in these binary mixtures. The theoretical values of speed of sound in the mixtures have been evaluated using various theories and have been compared with experimentally determined speed of sound values in order to check the applicability of such theories to the liquid mixture systems under study. Viscosity data have been used to test the applicability of standard viscosity models of Grunberg–Nissan, Hind–Mc Laughlin, Katti–Chaudhary, Heric and Brewer, Frenkel, Tamura and Kurata at various temperatures for the binary liquid systems under study.  相似文献   

14.
The stability constants of 1:1 (M:L) complexes of benzo-15-crown-5 (B15C5) with Li+, Na+, K+ and NH4 + cations, the Gibbs standard free energies ( $ \Updelta {\text{G}}_{\text{c}}^{ \circ } $ ), the standard enthalpy changes ( $ \Updelta {\text{H}}_{\text{c}}^{ \circ } $ ) and standard entropy changes ( $ \Updelta {\text{S}}_{\text{c}}^{ \circ } $ ) for formation of these complexes in acetonitrile–methanol (AN–MeOH) binary mixtures have been determined conductometrically. The conductance data show that the stoichiometry of the complexes formed between the macrocyclic ligand and the studied cations is 1:1 (M:L). In most cases, addition of B15C5 to solutions of these cations, causes a continuous increase in the molar conductivities which indicates that the mobility of complexed cations is more than the uncomplexed ones. The stability constants of the complexes were obtained from fitting of molar conductivity curves using a computer program, GENPLOT. The results show that the selectivity order of B15C5 for the metal cations changes with the nature and composition of the binary mixed solvent. The values of standard enthalpy changes ( $ \Updelta {\text{H}}_{\text{c}}^{ \circ } $ ) for complexation reactions were obtained from the slope of the van’t Hoff plots and the changes in standard entropy ( $ \Updelta {\text{S}}_{\text{c}}^{ \circ } $ ) were calculated from the relationship $ \Updelta {\text{G}}_{{{\text{c}},298.15}}^{ \circ } = \Updelta {\text{H}}_{\text{c}}^{ \circ } - 298.15\Updelta {\text{S}}_{\text{c}}^{ \circ } $ . A non-linear behavior was observed between the stability constants (log Kf) of the complexes and the composition of the acetonitrile–methanol (AN–MeOH) binary solution. The results obtained in this study, show that in most cases, the complexes formed between B15C5 and Li+, Na+, K+ and NH4 + cations are both enthalpy and entropy stabilized and the values of these thermodynamic quantities change with the composition of the binary solution.  相似文献   

15.
β-d-Xylosidase/α-l-arabinofuranosidase from Selenomonas ruminantium is the most active enzyme known for catalyzing hydrolysis of 1,4-β-d-xylooligosaccharides to d-xylose. Catalysis and inhibitor binding by the GH43 β-xylosidase are governed by the protonation states of catalytic base (D14, pK a 5.0) and catalytic acid (E186, pK a 7.2). Biphasic inhibition by triethanolamine of E186A preparations reveals minor contamination by wild-type-like enzyme, the contaminant likely originating from translational misreading. Titration of E186A preparations with triethanolamine allows resolution of binding and kinetic parameters of the E186A mutant from those of the contaminant. The E186A mutation abolishes the pK a assigned to E186; mutant enzyme binds only the neutral aminoalcohol $ \left( {{\text{pH}} - {\text{independent}}\;K_{\text{i}}^{\text{triethanolamine}} = 19\,{\text{mM}}} \right) $ , whereas wild-type enzyme binds only the cationic aminoalcohol $ \left( {{\text{pH}} - {\text{independent}}\;K_{\text{i}}^{\text{triethanolamine}} = 0.065\,{\text{mM}}} \right) $ . At pH 7.0 and 25°C, relative kinetic parameter, $ k_{\text{cat}}^{\text{4NPX}}/k_{\text{cat}}^{\text{4NPA}} $ , for substrates 4-nitrophenyl-β-d-xylopyranoside (4NPX) and 4-nitrophenyl-α-l-arabinofuranoside (4NPA) of E186A is 100-fold that of wild-type enzyme, consistent with the view that, on the enzyme, protonation is of greater importance to the transition state of 4NPA whereas ring deformation dominates the transition state of 4NPX.  相似文献   

16.
The densities, ρ, speeds of sound, u, and heat capacities, (C P)mix, for binary 1-ethyl-3-methylimidazolium tetrafluoroborate (1) + cyclopentanone or cyclohexanone (2) mixtures within temperature range (293.15–308.15 K) and excess molar enthalpies, H E, at 298.15 K have been measured over the entire composition range. The excess molar volumes, V E, excess isentropic compressibilities, \( \kappa_{\text{S}}^{\text{E}}, \) and excess heat capacities, \( C_{\text{P}}^{\text{E}}, \) have been computed from the experimental results. The V E, \( \kappa_{\text{S}}^{\text{E}} \) , H E, and \( C_{\text{P}}^{\text{E}} \) values have been calculated and compared with calculated values from Graph theory. It has been observed that V E, \( \kappa_{\text{S}}^{\text{E}} \) , H E, and \( C_{\text{P}}^{\text{E}} \) values were predicted by Graph theory compare well with their experimental values. The V E, \( \kappa_{\text{S}}^{\text{E}}, \) and H E thermodynamic properties have also been analyzed in terms of Prigogine–Flory–Patterson theory.  相似文献   

17.
N-1-Naphthylethylenediamine dihydrochloride monomethanolate (N-NEDHME) was tested as a corrosion inhibitor for copper in 2 M HNO3 solution using the standard gravimetric technique at 303–343 K. N-NEDHME acts as an inhibitor for copper in an acidic medium. Inhibition efficiency increases with increase in concentration of N-NEDHME but decreases with a rise in temperature. Thermodynamic parameters such as adsorption heat ( $ \Updelta H_{\text{ads}}^\circ $ ), adsorption entropy ( $ \Updelta S_{\text{ads}}^\circ $ ) and adsorption free energy ( $ \Updelta G_{\text{ads}}^\circ $ ) were obtained from experimental data of the temperature studies of the inhibition process at five temperatures ranging from 303 to 343 K. Kinetic parameters activation such as $ E_{a} $ , $ \Updelta H_{\text{a}}^\circ $ , $ \Updelta S_{\text{a}}^\circ $ and pre-exponential factors have been calculated and are discussed. Adsorption of N-NEDHME on the copper surface in 2 M HNO3 follows the Langmuir isotherm model.  相似文献   

18.
In this work, densities and viscosities for the binary mixtures of 1-octyl-3-methylimidazolium tetrafluoroborate, [C8mim][BF4], with butanone, methyl acetate, ethyl acetate, propyl acetate and butyl acetate have been determined at 298.15?K. These data were used to calculate the excess molar volumes $ V_{m}^{\text{E}} $ and the viscosity deviations (ln??)E for the mixtures. It is shown that values of $ V_{m}^{\text{E}} $ are negative, while those of (ln??)E are positive in the whole concentration range. The $ V_{m}^{\text{E}} $ values show their minimum at the composition of x IL ????0.3, and the (ln??)E values exhibit a maximum at the same mole fraction. For the binary systems, the absolute values of $ V_{m}^{\text{E}} $ decrease in the order: butanone?>?methyl acetate?>?ethyl acetate?>?propyl acetate?>?butyl acetate, whereas those of (ln??)E decrease in the order: methyl acetate?>?ethyl acetate?>?propyl acetate?>?butyl acetate?>?butanone. The results are discussed in terms of the ion?Cdipole interactions between cations of the ionic liquid and the organic molecules and hydrogen bonding interactions between anions of the ionic liquid and the organic compounds.  相似文献   

19.
The densities, ρ, refractive indices, n D, and ultrasonic speeds, u, of binary mixtures of acetonitrile (AN) with poly(ethylene glycol) 200 (PEG200), poly(ethylene glycol) 300 (PEG300) and poly(ethylene glycol) 400 (PEG400) were measured over the entire composition range at temperatures (298.15, 303.15, 308.15 and 313.15) K and at atmospheric pressure. From the experimental data, the excess molar volumes, \( V_{\text{m}}^{\text{E}} \) , deviations in refractive indices, \( \Delta n_{\text{D}} \) , excess molar isentropic compressibility, \( K_{{s , {\text{m}}}}^{\text{E}} \) , excess intermolecular free length, \( L_{\text{f}}^{\text{E}} \) , and excess acoustic impedance, Z E, have been evaluated. The partial molar volumes, \( \overline{V}_{\text{m,1}} \) and \( \overline{V}_{\text{m,2}} \) , partial molar isentropic compressibilities, \( \overline{K}_{{s , {\text{m,1}}}} \) and \( \overline{K}_{{s , {\text{m,2}}}} \) , and their excess values over whole composition range and at infinite dilution have also been calculated. The variations of these properties with composition and temperature are discussed in terms of intermolecular interactions in these mixtures. The results indicate the presence of specific interactions among the AN and PEG molecules, which follow the order PEG200 < PEG300 < PEG400.  相似文献   

20.
The standard Gibbs energies $ \left( {\Updelta {}_{\text{t}}G^\circ (i)} \right) $ ( Δ t G ° ( i ) ) and entropies $ \left( {\Updelta {}_{\text{t}}S^\circ } \right) $ ( Δ t S ° ) of transfer in aqueous mixtures of 1,2-dimethoxyethane (DME) containing 0, 20, 40, 60, 80, 100 wt-% DME have been determined from the solubility data of a series of homologous α-amino acids, evaluated by the formol titrimetric method. The observed result of Δt G°(i) and TΔt S°(i) against DME concentration profiles are complicated due to the various interaction effects. The chemical effects on the transfer Gibbs energies ( $ \Updelta_{\text{t}} G_{\text {ch}}^{ \circ } (i) $ Δ t G ch ° ( i ) ) and entropies of transfer $ T\Updelta_{\text{t}} S_{\text {ch}}^{ \circ } (i) $ T Δ t S ch ° ( i ) have been obtained after elimination of the cavity effect, calculated by the scaled particle theory, and dipole–dipole interaction effects, estimated by the use of Keesom-orientation expression for total transfer Gibbs energies Δt G°(i) and entropies Δt S°, respectively. The chemical transfer energetics of the zwitterionic homologous α-amino acids are guided by the composite effects of increased dispersion interaction, basicity and decreased acidity, hydrogen bonding capacity and hydrophobic hydration of the DME mixed solvent as compared to that of reference solvent, water.  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