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 265 毫秒
1.
Vapor-liquid equilibria (VLE) have been measured for five 1-hexene/n-hexane/ionic liquid systems and 1-hexene/n-hexane/NMP (N-methyl-2-pyrrolidone) system with a headspace-gas chromatography (HSGC) apparatus at 333.15 K. The ionic liquids investigated were 1,3-dimethylimidazolium tetrafluoroborate [C2MIM]+[BF4], 1-butyl-3-methylimidazolium tetrafluoroborate [C4MIM]+[BF4], 1-methyl-3-octylimidazolium tetrafluoroborate [C8MIM]+[BF4], 1,3-dimethylimidazolium dicyanamide [C2MIM]+[N(CN)2] and 1-octylquinolinium bis(trifluoromethylsulfonyl)amide [C8Chin]+[BTA]. It was found that at low feeding concentration of 1-hexene and n-hexane, the separation ability of ionic liquids is in the order of [C2MIM]+[BF4] > [C4MIM]+[BF4] ≈ [C2MIM]+[N(CN)2] > [C8MIM]+[BF4] > [C8Chin]+[BTA], which is consistent with the priori prediction of the COSMO-RS (conductor-like screening model for real solvents) model. But at high feeding concentration, the separation ability of ionic liquids is in the order of [C2MIM]+[BF4] < [C4MIM]+[BF4] ≈ [C2MIM]+[N(CN)2] < [C8MIM]+[BF4] < [C8Chin]+[BTA]. The liquid demixing effect should be taken into account. The activity coefficients of 1-hexene and n-hexane at infinite dilution calculated with the COSMO-RS model were correlated using the NRTL, Wilson and UNIQUAC model. In this work the predictive results from the COSMO-RS model and UNIFAC model for the 1-hexene/n-hexane and 1-hexene/n-hexane/NMP systems were compared. The UNIFAC model is one of the most important academic contributions by Prof. Jürgen Gmehling.  相似文献   

2.
Density functional theory is employed to study the interaction energies between dibenzothiophene (DBT) and 1-alkyl-3-methylimidazolium tetrafluoroborate ([C n mim]+[BF4]?). The structures of DBT, 1-ethyl-3-methylimidazolium tetrafluoroborate ([C2mim]+[BF4]?), 1-butyl-3-methylimidazolium tetrafluoroborate ([C4mim]+[BF4]?), 1-hexyl-3-methylimidazolium tetrafluoroborate ([C6mim]+[BF4]?), 1-octyl-3-methylimidazolium tetrafluoroborate ([C8mim]+[BF4]?), [C2mim]+[BF4]?–DBT, [C4mim]+[BF4]?–DBT, [C6mim]+[BF4]?–DBT and [C8mim]+[BF4]?–DBT systems are optimized systematically at the B3LYP/6-31G(d,p) level, and the most stable geometries are obtained by NBO and AIM analyses. The results indicate that DBT and imidazolium rings of ionic liquids are parallel to each other. It is found that the [BF4]? anion prefers to be located close to a C1–H9 proton ring in the vicinity of the imidazolium ring and the most stable gas-phase structure of [C n mim]+[BF4]? has four hydrogen bonds between [C n mim]+ and [BF4]?. There are hydrogen bonding interactions, π–π and C–H–π interactions between [C8mim]+[BF4]? and DBT, which is confirmed by NBO and AIM analyses. The calculated interaction energies for the studied ionic liquids can be used to interpret a better extracting ability of [C8mim]+[BF4]? to remove DBT, due to stronger interactions between [C8mim]+[BF4]? and DBT, in agreement with the experimental results of dibenzothiophene extraction by [C n mim]+[BF4]?.  相似文献   

3.
In this study, 1-n-tetradecyl-3-methylimidazolium bromide ([C14mim]Br), one kind of imidazolium ionic liquid (imi-IL), was incorporated into polypropylene (PP) via melt blending. The measurement of surface resistance (Rs) and volume resistance (Rv) of neat PP and PP/[C14mim]Br blends indicated that [C14mim]Br had excellent antistatic properties. The PP/[C14mim]Br blend had the best antistatic ability, when the weight ratio of [C14mim]Br to PP reached 3/100. The surface resistance of PP/[C14mim]Br decreased from the 7.67 × 1013 to 1.40 × 107 Ω whereas the volume resistance of PP decreased from 2.67 × 1014 to 2.60 × 107 Ω. Semicrystalline morphology and crystal structure were investigated by polarized optical microscopy (POM) and X-ray diffraction (XRD). The spherulites of PP became smaller with the addition of [C14mim]Br, implying that [C14mim]Br had a nucleating effect in the PP matrix. The XRD study indicates the crystallization process of PP was affected by [C14mim]Br, and the crystallinity of PP was thus decreased. Scanning electron microscopy (SEM) studies reveal that [C14mim]Br had good dispersion in PP; thermogravimetric analysis (TGA) shows the addition of [C14mim]Br remarkably increased the thermal stability of PP even though it is a small molecule.  相似文献   

4.
In this work, experimental data of isobaric vapour–liquid equilibria for the ternary system ethanol + water + 1-hexyl-3-methylimidazolium chloride ([C6mim][Cl]) and for the corresponding binary systems containing the ionic liquid (ethanol + [C6mim][Cl], water + [C6mim][Cl]) were carried out at 101.300 kPa. VLE experimental data of binary and ternary systems were correlated using the NRTL equation. In a previous work [N. Calvar, B. González, E. Gómez, A. Domínguez, J. Chem. Eng. Data 51 (2006) 2178–2181], the VLE of the ternary system ethanol + water + [C4mim][Cl] was determined and correlated, so we can study the influence of different ionic liquids in the behaviour of the azeotropic mixture ethanol–water.  相似文献   

5.
Alkylimidazolium salts are a very important class of compounds. So far, calorimetry has hardly been used to characterize their solution behaviour. The enthalpies obtained from indirect methods have an intrinsic large uncertainty, and nowadays it is clear that calorimetry is the most sensitive technique for directly measuring the thermodynamic properties of aggregation.In this work, isothermal titration calorimetry (ITC) was used along with conductivity to determine the thermodynamics of aggregation of 1-alkyl-3-methylimidazolium chlorides ([Cnmim]Cl, n = 8, 10, 12, and 14) in aqueous solution. The critical micelle concentrations, cmc, were obtained from conductivity and calorimetry, and the enthalpies of micelle formation, ΔHmic, were derived from the calorimetric titrations. From conductivity, we could also derive the values for the degree of ionisation of the micelles (α), the molar conductivity (ΛM) of the [Cnmim]Cl micellar species and the molar conductivity at infinite dilution (Λ) for the [Cnmim]+ cations.Values are therefore reported for the enthalpy (ΔHmic), the Gibbs free energy (ΔGmic) and entropy (ΔSmic) changes for micelle formation. Further, the aggregate sizes and aggregation numbers were obtained by light scattering (LS) measurements.The observed variation of the thermodynamic properties with the alkyl chain length is discussed in detail and compared with the traditional cationic surfactants 1-alkyl-trimethylammonium chlorides, [CnTA]Cl. The difference in the values of the thermodynamic parameters for both types of surfactants is here related to the structure of their head groups.  相似文献   

6.
1-Octadecyl-3-methylimidazolium chloride ([C18 mim]Cl) is a kind of imidazolium ionic liquid with high thermal stability. [C18 mim]Cl was used to modify pristine Na-montmorillonite and a series of organo-montmorillonite (OMMT) with different loading levels of 1-octadecyl-3-methylimidazolium cation ([C18 mim]+) were obtained. X-ray diffraction (XRD) and thermogravimetric analysis (TGA) results show that there are different loading levels and aggregative state of [C18 mim]+ in the interlayer of OMMT. The effects of OMMT interlayer micro-circumstances on the PP melting intercalation were studied by XRD and transmission electron microscopy (TEM). Results indicate that the melting intercalation of PP into the interlayer of OMMT is not only related with d-spaces but also has something to do with the interlayer micro-circumstance of OMMT. Based on these facts, three types of interlayer absorption models of [C18 mim]+ in the interlayer of OMMT were conceived. In addition, the aggregative state of [C18 min]+ in the interlayer of OMMT, interlayer polarity and d-spaces of OMMT were discussed. According to these models, we try to illustrate the effect of interlayer micro-circumstance of OMMT on the PP melting intercalation.  相似文献   

7.
Deep blue tetrachloronickelate(II) salts of 1-alkyl-3-methylimidazolium cations [Cnmim]+ with intermediate chain lengths (n = 5, 7, 8 and an n = 4/6 mixture) have been prepared from imidazolium chlorides and NiCl2. They are liquid at ambient temperatures (ca. 20 °C) and above, and display viscosities and thermal stability comparable to related [BF4] compounds. The similarities may reflect the compensating effects of the dinegative charge of the [NiCl4]2− ion and its larger size.  相似文献   

8.
The solubility of gallic acid in (water + ethanol) binary solvents was determined from (293.15 to 318.15) K at atmospheric pressure using a thermostatted reactor and UV/vis spectrophotometer analysis. The effects of binary solvents composition and temperature on the solubility were discussed. It was found that gallic acid solubility in (water + ethanol) mixed solvents presents a maximum-solubility effect. Two empirical equations were proposed to correlate the solubility data. The calculated solubilities show good agreement with the experimental data within the studied temperature range. Using the experimentally measured solubilities, the thermodynamic properties of dissolution of the gallic acid such as Gibbs energy (ΔsolG°), molar enthalpy of dissolution (ΔsolH°), and molar entropy of dissolution (ΔsolS°) were calculated.  相似文献   

9.
The solubility of hydrogen sulfide in a series of 1-(2-hydroxyethyl)-3-methylimidazolium ([HOemim]+)-based ionic liquids (ILs) containing different anions, viz. hexafluorophosphate ([PF6]), trifluoromethanesulfonate ([OTf]), and bis-(trifluoromethyl)sulfonylimide ([Tf2N]) at temperatures ranging from 303.15 to 353.15 K and pressures of up to about 1.8 MPa was measured by a volumetric based static apparatus. The solubility data were correlated using two models: (1) the Krichevsky–Kasarnovsky equation and (2) the extended Henry's law combined with the Pitzer's virial expansion for the excess Gibbs energy. Henry's law constants (at zero pressure) in mole-fraction and molality scales were obtained at different temperatures by means of these two models. Using the solubility data, the partial molar thermodynamic functions of solution, i.e. Gibbs energy, enthalpy, and entropy were calculated. Comparison showed that the solubility of H2S is greater than that of CO2 in the corresponding ILs studied in this work and that the solubility of both gases increases as the number of trifluoromethyl (–CF3) groups in the anion increases, i.e. the solubility behavior of both gases follows the order [HOemim][Tf2N] ≥ [HOemim][OTf] > [HOemim][PF6] > [HOemim][BF4].  相似文献   

10.
Considering the ionic nature of ionic liquids (ILs), ionic association is expected to be essential in solutions of ILs and to have an important influence on their applications. Although numerous studies have been reported for the ionic association behavior of ILs in solution, quantitative results are quite scarce. Herein, the conductivities of the ILs [Cnmim]Br (n=4, 6, 8, 10, 12), [C4mim][BF4], and [C4mim][PF6] in various molecular solvents (water, methanol, 1‐propanol, 1‐pentanol, acetonitrile, and acetone) are determined at 298.15 K as a function of IL concentration. The conductance data are analyzed by the Lee–Wheaton conductivity equation in terms of the ionic association constant (KA) and the limiting molar conductance (Λm0). Combined with the values for the Br? anion reported in the literature, the limiting molar conductivities and the transference numbers of the cations and [BF4]? and [PF6]? anions are calculated in the molecular solvents. It is shown that the alkyl chain length of the cations and type of anion affect the ionic association constants and limiting molar conductivities of the ILs. For a given anion (Br?), the Λm0 values decrease with increasing alkyl chain length of the cations in all the molecular solvents, whereas the KA values of the ILs decrease in organic solvents but increase in water as the alkyl chain length of the cations increases. For the [C4mim]+ cation, the limiting molar conductivities of the ILs decrease in the order Br?>[BF4]?>[PF6]?, and their ionic association constants follow the order [BF4]?>[PF6]?>Br? in water, acetone, and acetonitrile. Furthermore, and similar to the classical electrolytes, a linear relationship is observed between ln KA of the ILs and the reciprocal of the dielectric constants of the molecular solvents. The ILs are solvated to a different extent by the molecular solvents, and ionic association is affected significantly by ionic solvation. This information is expected to be useful for the modulation of the IL conductance by the alkyl chain length of the cations, type of anion, and physical properties of the molecular solvents.  相似文献   

11.
Oxidative addition of methyl iodide to Vaska’s complex in the ionic liquids 1-butyl-3-methylimidazolium triflate [C4mim][OTf], [C4mim] bis(trifluormethylsulfonyl)imide [Tf2N], and N-hexylpyridinium [C6pyr][Tf2N] occurred cleanly to give the expected Ir(III) oxidative addition product. Pseudo-first order rate constants were determined for the oxidative addition reaction in each solvent ([Vaska’s] = 0.25 mM, [CH3I] = 37.5 mM). The observed rate constants under these conditions were 5-10 times slower than the rate seen in DMF. At high methyl iodide concentrations (>23 mM), the expected first order dependence on methyl iodide was not observed. In each ionic liquid, there was no change in the reaction rates within experimental error over the methyl iodide concentration range of 23-75 mM. At lower methyl iodide concentration, a decrease in rate was observed in [C4mim][Tf2N] with decreasing concentration of methyl iodide.  相似文献   

12.
Several imidazolium-based ionic liquids (ILs) with varying cation alkyl chain length (C4–C10) and anion type (tetrafluoroborate ([BF4]), hexafluorophosphate ([PF6]) and bis(trifluoromethylsulfonyl)imide ([Tf2N])) were used as reaction media in the microwave polymerization of methacrylate-based stationary phases. Scanning electron micrographs and backpressures of poly(butyl methacrylate-ethylene dimethacrylate) (poly(BMA-EDMA)) monoliths synthesized in the presence of these ionic liquids demonstrated that porosity and permeability decreased when cation alkyl chain length and anion hydrophobicity were increased. Performance of these monoliths was assessed for their ability to separate parabens by capillary electrochromatography (CEC). Intra-batch precision (n = 3 columns) for retention time and peak area ranged was 0.80–1.13% and 3.71–4.58%, respectively. In addition, a good repeatability of RSDRetention time = <0.30% and ∼1.0%, RSDPeak area = <1.30% and <4.3%, and RSDEfficiency = <0.6% and <11.5% for intra-day and inter-day, respectively exemplify monolith performance reliability for poly(BMA-EDMA) fabricated using 1-hexyl-3-methylimidazolium tetrafluoroborate ([C6mim][BF4]) porogen. This monolith was also tested for its potential in nanoLC to separate protein digests in gradient mode. ILs as porogens also fabricated different alkyl methacrylate (AMA) (C4–C18) monoliths. Furthermore, employing binary IL porogen mixture such as 1-butyl-3-methylimidazolium tetrafluoroborate ([C4mim][BF4]) and 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([C4mim][Tf2N]) successfully decreased the denseness of the monolith, than when using [C4mim][Tf2N] IL alone, enabling a chromatographic run to be performed with 1:1 ratio produced baseline separation for the analytes. The combination of ILs and microwave irradiation made polymer synthesis very fast (∼10 min), entirely green (organic solvent-free) and energy saving process.  相似文献   

13.
14.
The syntheses and spectroscopic (NMR, MS) investigations of the antimonates [Ph4P]+[Me2SbCl4] (1), [Me4Sb]+[Me2SbCl4] (2), [Et4N]+[Ph2SbCl4] (3), [Bu4N]+[Ph2SbCl4] (4), [Me4Sb]+[Ph2SbCl4] (5), [Et3MeSb]+[Ph2SbCl4] (6), [Et4N]+[Ph2SbF4] (7) and [Et4N]+[Ph2SbBr4] (8) are reported. Halogen scrambling reactions of Et4NBr or Ph4EBr (E = P, Sb) with R2SbCl3 (R = Me, Ph) produce mixtures of compounds from which crystals of [Et4N]+[Ph2SbBr1.24Cl2.76] (9), [Et4N]+[Ph2SbBr2.92Cl1.08] (10) or [Ph4Sb]+[Me2SbCl4] (11) were isolated. The crystal and molecular structures of 1 and 3-11 are reported.  相似文献   

15.
The vapour pressure of binary mixtures of hydrogen sulphide with ethane, propane, and n-butane was measured at T = 182.33 K covering most of the composition range. The excess Gibbs free energy of these mixtures has been derived from the measurements made. For the equimolar mixtures for (H2S + C2H6), (820.1 ± 2.4) J · mol−1 for (H2S + C3H8), and (818.6 ± 0.9) J · mol−1 for (H2S + n-C4H10). The binary mixtures of H2S with ethane and with propane exhibit azeotropes, but that with n-butane does not.  相似文献   

16.
The compounds [MBr2(an)2] (where M is Mn(II), Fe(II), Co(II), Ni(II), Cu(II) or Zn(II); an = aniline) were synthesized and characterized by melting points, elemental analysis, thermal studies, and electronic and IR spectroscopy. The enthalpies of dissolution of the adducts, metal(II) bromides and aniline in methanol, aqueous 1.2 M HCl or 25% (v/v) aqueous 1.2 M HCl in methanol were measured. The following thermochemical parameters for the adducts have been determined by thermochemical cycles: the standard enthalpies for the Lewis acid/base reactions (ΔrH°), the standard enthalpies of formation (ΔfH°), the standard enthalpies of decomposition (ΔDH°), the lattice standard enthalpies (ΔMH°) and the standard enthalpies of the Lewis acid/base reactions in the gaseous phase (ΔrH°(g)). The mean bond dissociation enthalpies of the M(II)-nitrogen bonds () and the enthalpies of formation of the adducts from the ions in the gaseous phase: M2+(g) + Br(g) + an(g) → [MBr2(an)2](g), (ΔfiH°) have been estimated.  相似文献   

17.
Using dynamic method and the laser monitoring observation technique, the solubility of cefodizime disodium in water + (ethanol, 1-propanol, and 2-propanol) was measured as a function of temperature from 278.15 K to 318.15 K under atmospheric pressure. The experimental data were correlated with a simple model of molecular thermodynamics for solubility of solid in liquid. The model parameters were fitted, and the solution enthalpies ΔsolH and solution entropies ΔsolS were estimated. ΔsolH and ΔsolS are all positive. The endothermic effect of solution process may be due to the fact that the newly bond energy between cefodizime disodium and solvent molecules is not powerful enough to compensate the energy needed to break the original association bond in various solvents, and the system needs to absorb heat from surroundings and manifests as ΔsolH > 0. The reason for the entropy increase during the dissolution process is that the solutes disrupt the alignment of solvent molecules and therefore reduced the degree of order of the system while they were dissolved in various solvents. The positive ΔsolH and ΔsolS revealed that the dissolution process of cefodizime disodium was an entropy-driven process.  相似文献   

18.
In this work, we have studied influence of ionic liquids (ILs) on the azeotrope composition for the system {diisopropyl ether (DIPE) + ethanol} using trihexyltetradecylphosphonium chloride ([P666,14][Cl]) and trihexyltetradecylphosphonium bis(2,2,4-trimethylpentyl) phosphinate ([P666,14][TMPP]). Isothermal vapor-liquid equilibrium data at 333.15 K are reported for the ternary systems {DIPE + ethanol + [P666,14][Cl]} and {DIPE + ethanol + [P666,14][TMPP]} with varying the mole fraction of ILs from 0.05 to 0.10. The experimental ternary VLE data were correlated using the Wilson equation. In addition, excess molar volumes (VE) and deviations in molar refractivity (ΔR) data at 298.15 K are reported for the binary systems {DIPE + [P666,14][Cl]} and {ethanol + [P666,14][Cl]} by a digital vibrating tube densimeter and a precision digital refractometer. The VE and ΔR were correlated by the Redlich-Kister equation.  相似文献   

19.
The molar enthalpies of solution of an alanine-based ionic liquid (IL) [C4mim][Ala], 1-butyl-3-methylimidazolium alanine, containing various amount of water and various molalities Δsol H m(wc), were measured with a solution-reaction isoperibol calorimeter at (298.15±0.01) K, where wc denotes water content. According to Archer’s method, the standard molar enthalpies of solution of [C4mim][Ala] containing known amounts of water, DsolHmo(wc)\Delta_{\mathrm{sol}}H_{\mathrm{m}}^{\mathrm{o}}(\mathrm{wc}) , were obtained. In order to eliminate the effect of the small amount of residual water in the source [C4mim][Ala], a linear fitting of DsolHmo(wc)\Delta_{\mathrm{sol}}H_{\mathrm{m}}^{\mathrm{o}}(\mathrm{wc}) against water content was carried out, yielding a good straight line where the intercept is the standard molar enthalpy of solution of anhydrous [C4mim][Ala], DsolHmo(pure IL)=-(61.42±0.08)\Delta_{\mathrm{sol}}H_{\mathrm{m}}^{\mathrm{o}}(\mathrm{pure}\ \mathrm{IL})=-(61.42\pm 0.08) kJ⋅mol−1. The hydration enthalpy of the alanine anion [Ala] was estimated using Glasser’s lattice energy theory.  相似文献   

20.
Liquid–liquid equilibrium diagrams were determined for (IL + water) systems using the family of ILs 1-alkyl-3-methylimidazolium tetrafluoroborates, where the alkyl groups are hexyl and octyl ([Cxmim][BF4] with x = 6 and 8). The gravimetric method was used to determine the equilibrium compositions at temperatures ranging from 278.15 to 340.15 K. Both systems present an upper critical solution temperature (UCST), which increases from [C6mim][BF4] to [C8mim][BF4]. The experimental data were correlated using the NRTL and eNRTL models. The binary interaction parameters were calculated for each system and model, and good agreement between experimental and calculated equilibrium compositions was obtained. Finally, the apparent Gibbs energy, enthalpy and entropy of water solution in the ILs were calculated using a modified van’t Hoff equation. The three thermodynamic functions were found to be positive for both ILs.  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