Comparison of Ionic Liquids to Conventional ... - ACS Publications

Apr 15, 2016 - Roberto I. Canales. † and Joan F. Brennecke*. Department of Chemical and Biomolecular Engineering, University of Notre Dame, Notre Da...
2 downloads 0 Views 3MB Size
Review pubs.acs.org/jced

Comparison of Ionic Liquids to Conventional Organic Solvents for Extraction of Aromatics from Aliphatics Roberto I. Canales† and Joan F. Brennecke* Department of Chemical and Biomolecular Engineering, University of Notre Dame, Notre Dame, Indiana 46556, United States ABSTRACT: Commonly used solvents for separation of aromatics from aliphatics using liquid−liquid extraction are reviewed in terms of their physical properties and separation performance. The experimental liquid− liquid measurements for conventional solvents are contrasted with the phase behavior when using an ionic liquid as the extracting component. For comparison purposes, ternary systems of heptane + toluene + extraction solvent are used as representative mixtures, where sulfolane is the main organic extraction solvent discussed. Since ionic liquid properties are drastically changed by changing the anion, comparisons are performed for 1ethyl-3-methylimidazolium based ionic liquids using different anions. Ionic liquids containing cyano-substituted and bis(trifluoromethylsulfonyl)imide anions provide good selectivities and distribution ratios. On the basis of the phase behavior and capacities, as well as their relatively low viscosities, they appear to be competitive alternatives to organic solvents.



INTRODUCTION The main source of aromatics in industry, for example, benzene, toluene, xylenes, and ethylbenzene, are catalytic reforming and steam cracking processes. During catalytic reforming, low-octane heavy naphtha from paraffinic or naphthenic oil is processed at 723−823 K and 1.7−7.0 MPa to obtain a high-octane reformate gasoline by isomerizing aliphatics, ring-formation from these alkanes, and finally dehydrogenating the cyclic aliphatics to form aromatics. A typical reformate gasoline stream can have an aromatic content ranging from 20 to 65 wt %,1,2 in which the main fraction of aromatics in this product is composed of xylenes and a smaller amount of toluene.3 In the steam cracking process any of a diverse range of oil streams, from light hydrocarbons to heavy naphtha, are fed to the reactor with the objective of producing olefins, such as ethylene, propylene, and other higher olefins. This process is highly endothermic, so near ambient pressure and temperatures over 950 K are required, depending on the feed.4 The product from the steam cracking is called pyrolysis gasoline, which is rich in benzene, and contains between 50 and over 90 wt % of aromatics.1,3 Aromatics from reformate and pyrolysis gasolines are generally not separated by simple distillation because the boiling points of the components in the mixture span a very narrow range and a variety of azeotropes exist. Different processes are used to perform the separation of the aromatics, depending on their concentration in the effluent. When the concentration of aromatics is high, i.e. over 90 wt %, it is economically viable for the aromatics to be separated through azeotropic distillation by adding a strong polar compound. Streams with a medium aromatic concentration, that is, from 65 to 90 wt %, can be processed by extractive distillation using, for example, N-methylpyrrolidone, phenols, glycols, etc. Streams © XXXX American Chemical Society

with lower concentrations of aromatics, that is, from 20 to 65 wt %, are separated by liquid−liquid extraction, which is reviewed in this work.1 For aromatic concentrations below 20 wt %, other techniques have to be pursued to economically remove the aromatics.5 The recovery of the aromatics from the extraction solvents used in the various processes is done by direct distillation or stripping with a secondary solvent, followed by further distillation.1,3 The main solvents reported in the literature for aliphatic− aromatic liquid−liquid extraction, which have also been used in industrial separations,1 are sulfolane6−41 and glycols,35,37,42−59 but many compounds have been tested for application in aliphatic−aromatic separations, including dimethyl sulfoxide,30,39 ethylene and propylene carbonate,30,57,60−64 Nmethylpyrrolidone,65−68 N-formylmorpholine,27,56,69−74 and dimethylformamide,64 among others. The ternary aliphatic + aromatic + solvent systems, with compositional analysis, distribution factors, and selectivities, are detailed in each study, which is useful for performance comparisons among different compound combinations. Moreover, multicomponent systems with more than three constituents have been published. Those publications focus on the addition of a modifier to the extraction solvent or the use of two typical solvents29,30,38,56,64,68,75 for the extraction of mixed aromatics from a single or multiple aliphatics.8,12,13,76−81 Ionic liquids have been introduced as an alternative to conventional solvents for liquid−liquid separations due to their very low vapor pressure, wide liquid range, and high density compared with common aliphatics and aromatics, such as Received: January 26, 2016 Accepted: April 4, 2016

A

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Table 1. Properties of Common Solvents Used in Aliphatic−Aromatic Liquid−Liquid Extraction at 298.15 K and 0.1013 MPa solvents ethylene glycol dimethylformamide dimethyl sulfoxide ethylene carbonate N-methylpyrrolidone propylene carbonate diethylene glycol N-formylmorpholine sulfolane toluene heptane a

MW (g·mol−1) 62.07 73.09 78.13 88.06 99.13 102.09 106.12 115.13 120.17 92.14 100.20

Tb (K) 470.65 426.25 462.25 519.15 475.05 514.75 518.65 513.15 560.45 383.75 371.53

Tm (K)

ρ (g·cm−3)

260.15 212.75 291.75 309.48 248.75 224.35 262.85 296.15 301.55 178.15 182.60

1.10 0.94 1.10 1.32b 1.03 1.20 1.11 1.15 1.26a 0.86 0.68

σ (mN·m−1)

ref

13 370 60 1.3 50 6 17c

48.60 35.83 41.70 50.60 40.25 34.60 27.93

9.1a 3787c 6066c

47.95a 27.76 19.63

188−190 172, 191 172, 192 188, 193−195 172, 191 188, 194−196 188, 197 198, 199 172, 200 188, 191 188,191

μ (mPa·s)

Pvap (Pa)

16.06 0.90 2.00 1.85b 1.67 2.47 30.2 7.67 10.35a 0.56 0.39

c

At 303.15 K. bAt 313.15 K. cEstimated with a correlation.201

heptane and toluene. One of the challenges is trying to find an ionic liquid with low viscosity, high selectivity, high capacity, and low toxicity, in order to improve the existing methods practiced in industry for aliphatic/aromatic separation. Liquid−liquid phase equilibrium, capacity, and selectivity have been reported for several combinations of aliphatics and aromatics using ionic liquids as the extraction solvent. Among the ionic liquids studied, the cations used are mostly imidazolium41,82−128 and pyridinium.41,93,116,128−149 A few studies use pyrrolidinium,150 ammonium,95,131,151 morpholinium,152 and other cations,86,153 while the most investigated ionic liquid anion is bis(trifluoromethylsulfonyl)imide ([Tf2N]−). In addition, aromatic extractions have been performed with ionic liquid mixtures in an attempt to improve both selectivity and capacity.154−159 In some cases, infinite dilution activity coefficients of aromatic and aliphatic compounds in ionic liquids have been used as a first pass evaluation of the suitability of a particular ionic liquid for aliphatic−aromatic separation.160,161 Here we focus on liquid− liquid equilibrium studies of ternary mixtures of heptane + toluene + imidazolium-based ionic liquids, paying particular attention to those containing the 1-ethy-3-methylimidazolium ([emim]+) cation. The method usually proposed for recovery of the aromatic from the ionic liquid extraction solvent is simple distillation or stripping, taking advantage of the thermal stability and low volatility of the ionic liquid.162 Jongmans et al.163 simulated the ethylbenzene/styrene extractive distillation process using 4methyl-N-butylpyridinium tetrafluoroborate ([4-mebupy][BF4]), proposing several regeneration techniques: simple evaporation, two stage evaporation with the second unit at very low pressure, nitrogen stripping, distillation, and supercritical CO2 extraction. The most promising recovery technique according to this author is the low pressure regeneration. Scurto et al.,164,165 Aki et al.,166 Mellein and Brennecke,167 Canales and Brennecke168 and Canales et al.169 have studied the salting-out effect of sub- and supercritical CO2 on mixtures of ionic liquids with different organic solvents and water. The addition of CO2 results in a liquid−liquid phase split, where the organic-rich or aqueous-rich phase can be quite pure. This technique is another promising method for recovering the aromatic and regenerating the ionic liquid for reuse if one can find an ionic liquid that releases the aromatic at moderate pressures and low CO2 compositions in the liquid phase. Lower pressures and use of less CO2 makes the liquid−liquid phase split process more attractive than supercritical CO2 extraction.

The objective of this brief review is to show the performance of ionic liquids in liquid−liquid separation of aromatic from aliphatics compared with common organic solvents. We will focus on physical and transport properties (e.g., density and viscosity) and the performance of the solvents in terms of selectivity and distribution ratio.



LIQUID−LIQUID SEPARATION OF AROMATICS FROM ALIPHATICS USING COMMON SOLVENTS Liquid−liquid separation is the most popular process in industry to extract aromatics from aliphatics. Selection of a suitable solvent is the crucial first step in performing successful liquid−liquid extractions. The best solvents have good selectivity and capacity, high thermal stability, good availability, low cost, high surface tension and low to moderate viscosity.170 More environmentally friendly solvents, which have a smaller impact on human health and the environment, is another feature to consider. These compounds, frequently called “green” solvents, must have a very low toxicity, persistence, and volatility.171 Table 1 presents some properties, such as boiling temperature at 0.1013 MPa, melting point, density, viscosity, vapor pressure and surface tension at 298.15 K of the typical solvents used for liquid−liquid extraction of aromatics from aliphatics. These properties are compared with the representative aliphatic and aromatic used in this work: heptane and toluene. The main characteristics observed for the conventional extraction solvents are their relatively low vapor pressure, low viscosity, and high surface tension. The low vapor pressure is an important factor for avoiding significant loss of the solvent, which would require replacement and add cost. Low viscosity is essential to decrease mass transfer problems in the mixing and separation of the liquid phases. A high surface tension is necessary for avoiding the presence of emulsions.170 All of the solvents have higher density than either heptane or toluene. This is virtually a requirement for selective extraction of the aromatic, since the aromatics are more dense than the aliphatics. If the solvent is to have good capacity and selectivity for the aromatic compound, the extract will inevitably be the more dense of the two liquid phases, facilitating their gravimetric separation. The aliphatics, with a small portion of solvent and aromatics, will be the raffinate, which is the low density upper phase in the liquid−liquid equilibrium. All the solvents show a liquid range of approximately 200 K and some of them melt close to the room temperature, such as sulfolane, dimethyl sulfoxide, ethylene carbonate, and N-formylmorpholine. The boiling point of the solvent is also important because B

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

it has to be significantly higher than that of the aromatic if the aromatic is to be recovered by simple distillation. It is also noted that the molecular weight of the extraction solvents are comparatively close to the values for heptane and toluene. The molecular structures of the different solvents shown in Table 1 are presented in Figure 1.

Figure 1. Molecular structure of some typical solvents used in aromatic/aliphatic liquid−liquid extraction. Figure 2. Comparison of ternary liquid−liquid equilibrium data for the heptane + toluene + sulfolane system at 298.15 K13,14,26,34,36 in (a) mole fraction and (b) mass fraction.

Sulfolane is the most studied separation solvent in the literature and is widely used in industry in extractive distillation and liquid−liquid aromatic extraction.1 In Figure 2, a comparison of the literature data for liquid−liquid equilibrium of the heptane + toluene + sulfolane system at 298.15 K is presented in mole and mass fraction bases, showing high consistency in the tie lines and coexistence curves among results from different researchers. Therefore, sulfolane is a good choice as a “base-case” solvent with which to compare experimental measurements of aliphatic + aromatic + solvent liquid−liquid equilibrium. However, note that sulfolane is more highly toxic than some other extraction solvents.172 It is not surprising that the mole and weight fraction diagrams are similar since the molecular weights of the three components are similar. There is usually some temperature effect on liquid−liquid equilibrium, with larger miscibility gaps generally observed at lower temperatures. However, the pressure effect on the phase behavior is negligible close to atmospheric conditions.170 Figure 3 presents a comparison of the tie lines in the system heptane + toluene + sulfolane at three different temperatures: 298.15 K, 313.15 K, and 323.15 K. In this mixture, the immiscible area is slightly larger at the lower temperature, as expected, but the tie lines follow the same slope and the compositions in both phases are not considerably different. Thus, there is not a strong temperature effect for this mixture over the relatively small temperature range shown in Figure 3. Figure 4 shows a comparison of the immiscibility gaps when different solvents are used for separating toluene from heptane at 313.15 K, where the largest immiscible region is with ethylene carbonate and the smallest is with N-methylpyrroli-

done. All the solvents appear to have a plait point, so high toluene concentration mixtures cannot be separated. Comparison of the ternary liquid−liquid phase equilibrium diagrams is useful in determining the compositions of the initial heptane + toluene mixtures that can be separated with a particular solvent. However, the best analysis is provided by calculating the capacity and selectivity of the solvent at different compositions of the two phases. The capacity of the solvent to dissolve heptane or toluene is defined as the distribution ratio (or distribution factor) Di. It can be defined in mole fraction (Dx,i in eq 1) or mass fraction (Dw,i in eq 2) basis where xαi or β is the mole fraction of compound i in the α or β phase and wαi or β is the mass fraction of compound i in the α or β phase, calling α as the lower extract phase and β as the upper raffinate phase, Dx , i = Dw , i =

xiα xiβ

(1)

wiα wiβ

(2)

A solvent with good capacity for separating the aromatic has to show a lower distribution ratio for the aliphatic than the aromatic. A high distribution ratio of the aromatic is required to decrease the solvent usage in the liquid−liquid extraction. The selectivity (S in eq 3) is calculated as the distribution ratio of C

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Figure 3. Temperature effect on the liquid−liquid equilibrium for heptane + toluene + sulfolane mixtures36 in (a) mole fraction and (b) mass fraction: (black square) 298.15 K, (red circle) 313.15 K, and (blue triangle) 323.15 K.

Figure 4. Liquid−liquid equilibrium of the heptane + toluene + solvent system at 313.15 K in (a) mole fraction and (b) mass fraction: (black square) sulfolane,36,41 (red circle) ethylene carbonate,30 (blue triangle) N-formylmorpholine,56,69 (green diamond) N-methylpyrrolidone,68 (cross) ethylene glycol59

the aromatic divided by the distribution ratio of the aliphatic in mole or mass fraction. S=

β α β Daromatic xα /xaromatic waromatic /waromatic = aromatic = α β α β Daliphatic xaliphatic /xaliphatic waliphatic /waliphatic

As shown in Figure 2, there is good agreement among literature data for the heptane + toluene + sulfolane system at 298.15 K. In Figure 6, the selectivity and distribution ratio are shown as a function of xβtoluene (and wβtoluene for mass distribution ratio) for the same data shown in Figure 2 at 298.15 K. Despite the apparent agreement in the liquid−liquid equilibrium data, there are significant differences in the selectivities at low toluene concentrations. This is because the selectivity is very sensitive to small variations in the mole fractions of the two liquid phases at low toluene concentrations. For example, the data of Tripathi et al.36 gives a selectivity of 44.9 at xβtoluene = 0.0898 but for Chen et al.13 S = 28.6 at xβtoluene = 0.083. These large variations in selectivity are decreased when xαtoluene > 0.2, where the compositions of the two liquid phases are closer to the plait point. All of the reports show increasing distribution ratio with increasing toluene composition, but as seen with the selectivity, there is more disparity at low toluene compositions. To evaluate data from different research groups, experimental compositions and selectivities should be analyzed for thermodynamic consistency, as is done with the Gibbs− Duhem equation for vapor−liquid equilibrium. Consistency of liquid−liquid equilibrium data is commonly checked by applying the correlations proposed by Hand173 and Othmer and Tobias,174 but it has been demonstrated that these “tests” are not sensitive enough to weight fraction and selectivity

(3)

As a consequence, a good liquid−liquid separation solvent or a mixture of extracting solvents has to have a high distribution ratio for the aromatic, and a low distribution ratio for the aliphatic that results in a high selectivity for the target compound, in order to obtain low amounts of aromatic in the aliphatic-rich phase. Selectivity is highly dependent on composition, so it also depends on the temperature. Figure 5 shows the selectivity and distribution ratio for the heptane + toluene + sulfolane system at different temperatures obtained from the work of Tripathi et al.,36 where selectivity decreases and the distribution ratio tends to increase by enhancing the concentration of toluene in the upper phase (xβtoluene). The most significant temperature effect on selectivity is observed at the lowest xβtoluene, where the selectivity is about 45 at 298.15 K but closer to 30 at 323.15 K. There is no clear temperature effect on the distribution ratio in the heptane + toluene + sulfolane system. Distribution ratio values decrease slightly when based on mass fraction because the molecular weight of sulfolane is bit higher than the molecular weights of toluene and heptane. D

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Figure 5. (a) Selectivity, (b) distribution ratio in mole fraction basis and (c) distribution ratio in mass fraction basis versus composition of toluene in the upper raffinate phase for the heptane + toluene + sulfolane36 system at (black square) 298.15 K, (red circle) 313.15 K, and (blue triangle) 323.15 K.

Figure 6. (a) Selectivity, (b) distribution factor in mole fraction basis and (c) distribution factor in mass fraction basis versus composition of toluene in the upper raffinate phase for the heptane + toluene + sulfolane system at 298.15 K. (black square) Tripathi et al.;36 (red circle) Lin et al.; (blue triangle) Chen et al.;13 (green diamond) De Fré and Verhoeye;14 (cross) Rawat and Gulati.34

variations and do not have a strong theoretical basis.175 Unfortunately, to our knowledge, there are no procedures to check thermodynamic consistency of liquid−liquid equilibrium data based on a robust theoretical technique.

to evaluate the economic feasibility of using ionic liquids for aliphatic−aromatic separations. Typical cations and anions used for aliphatic/aromatic liquid−liquid separation are shown in Tables 2 and 3, respectively. Ionic liquids using the 1-ethyl-3-methylimidazolium ([emim]+) cation are commonly studied since the short alkyl chains result in good selectivity and distribution ratios (as discussed below). Table 4 presents densities, viscosities, and surface tensions of several [emim]+ based ionic liquids with different anions, in comparison to sulfolane, heptane, and toluene. The density and surface tension of ionic liquids are very similar to common liquid−liquid extraction solvents. The



LIQUID−LIQUID SEPARATION OF AROMATIC FROM ALIPHATICS USING IONIC LIQUIDS Ionic liquids are usually immiscible with aliphatics, with very small mutual solubilities. On the other hand, the solubility of aromatics in many ionic liquids is quite high, but generally they are not completely miscible. This behavior allows the ionic liquid to separate aromatic/aliphatic mixture at any feed composition but a selectivity and capacity analysis is necessary E

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Table 2. Ionic Liquid Cations Used for Liquid−Liquid Extraction of Aromatics from Aliphatics

Table 3. Ionic Liquid Anions Used for Liquid−Liquid Extraction of Aromatics from Aliphatics

a A, B, C, D, n = number of carbons in the alkyl chain; X, Y = initial of alkyl chain.

main difference is the higher viscosities of the ionic liquids, ranging from 15 to 320 mPa·s. On the basis of viscosity, the most competitive ionic liquid alternatives are ones with the [emim]+ cation and anions containing cyano groups, such as [DCA]−, [TCM]−, and [SCN]−, as well as [Tf2N]−, all of which have viscosities similar to sulfolane. The vapor pressures of the ionic liquids are generally very small compared to organic solvents and this represents an advantage for the reduction of the gas emissions to the atmosphere and the reduction of the regeneration cost of the solvent. The molecular weights of ionic liquids are usually much higher than those of common extraction solvents. A typical ternary diagram for heptane + toluene +1-ethyl-3methyl imidazolium bis(trifluoromethylsulfonyl)imide ([emim][Tf2N]) is shown in Figure 7 at two different temperatures in mole and mass fractions. There is very little difference in the data at 298.15 K and 313.15 K, from the groups of Arce et al.84 and Garcia et al.,176 respectively, and tie lines have similar slope. When compared with the results of Corderi ́ et al. at 298.15 K, there is a small variation in the compositions of the lower phase and the slopes of their tie lines are different than the other groups when the toluene composition is over 0.2 mole fraction. These differences could lead to significant variation in the selectivities. When compositions are shown in mass fractions, the difference between the different experimental values is more difficult to distinguish because the values are dominated by the high molecular weight of [emim][Tf2N]. In Figure 8, the effect of the length of the alkyl chains of the imidazolium cation (1-alkyl-3-methylimidazolium) on the selectivity and the distribution ratio of toluene at two different temperatures (298.15 and 313.15 K) are explored. The effect of temperature on the selectivity and distribution ratio is very small, as seen in the heptane + toluene + [emim][Tf2N] and

[hmim][Tf2N] systems, from the work of Arce et al.84 at 298.15 K and Garcia et al.104 at 313.15 K. As noted previously for the data in the ternary diagram of Figure 7, the selectivities calculated from Corderi ́ et al.93 will underestimate the values for [emim][Tf2N] compared to the other researchers, but show similar distribution ratios at low toluene compositions. When the distribution ratios are recalculated in terms of the mass fractions, the values are decreased dramatically, as expected, due to the high molecular weights of the ionic liquids. The selectivities for the different [Cxmim][Tf2N] ionic liquids is [hmim] < [bmim] < [emim] < [mmim], which means that a longer alkyl chain decreases the selectivity. The F

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Table 4. Properties of [emim]+ Ionic Liquids Used in Liquid−liquid Extraction of Aromatics from Aliphatics at 298.15 K solvents sulfolane [SCN]− [OAc]− [TCM]− [DCA]− [CH3SO3]− [ESO4]− [DEP]− [Tf2N]− toluene heptane a

MW (g·mol−1) 120.17 169.25 170.21 201.23 177.21 206.26 236.29 264.26 391.31 92.14 100.20

ρ (g·cm−3)

μ (mPa·s)

a

a

1.26 1.11684 1.09778 1.08162 1.10198 1.2398 1.24018 1.146 1.51900 0.86 0.68

10.35 22.15 132.91 15.02 19.90 155 101.42 320.89 32.99 0.56 0.39

σ (mN·m−1)

ref

47.95a 57.76 42.9 50.94 40.3 45.1 46.9 34.46 35.2 27.76 19.63

172, 200 202 203 118, 204 203 205, 206 203 207 208, 209 188,191 188,191

At 303.15 K.

0.799), Dx,heptane = 0.294 (Dw,heptane = 0.144) and Dx,toluene = 0.823 (Dw,toluene = 0.404), obtaining a S = 2.80. At these conditions, the change achieved by increasing the alkyl chain length from butyl to hexyl is ΔDx,heptane = 0.114 (ΔDw,heptane = 0.059), ΔDx,toluene = 0.037 (ΔDw,toluene = 0.034) and ΔS = −1.56, from which one can conclude that the effect on the heptane mole distribution ratio is three times greater (twice in the case of mass fraction) than the effect on the toluene distribution ratio. Figure 9 shows the selectivity and toluene distribution ratio for separating toluene from heptane for ionic liquids with the [emim]− cation and a variety of different anions. The results are compared with sulfolane. The sulfolane data selected for comparison was the data that showed the higher selectivity and distribution ratio of toluene compared with other values from literature.36 In terms of selectivity, all the ionic liquids presented in Figure 9 are competitive with sulfolane, since the lowest value of selectivity for ionic liquids is for [emim][Tf2N], which is very close to the sulfolane curve. The highest selectivity is when [emim][SCN] is used, with selectivities over 90 at the lowest compositions. High selectivities are also observed with [emim][CH3SO3] and [emim][DCA]. In the case of the toluene mole distribution ratio comparison, only three ionic liquids are competitive with sulfolane: [emim][DEP], [emim][TCM], and [emim][Tf2N]. Ionic liquids with the highest selectivities fall far below the sulfolane curve in terms of distribution ratio, which means that a large amount of solvent would be required for separating the aliphatic/aromatic mixture. The problem with [emim][Tf2N] is the higher concentration of aliphatic dissolved in the ionic liquid. In general, the ionic liquids presented in Figure 9 show better or comparable selectivities to sulfolane, but just [emim][TCM], [emim][DEP], and [emim][Tf2N] have a higher or similar mole distribution ratio of toluene. The main disadvantage of using [emim][DEP] is its high viscosity. However, [emim][TCM] and [emim][Tf2N] may be viable alternatives to replace sulfolane. For mass fraction distribution ratios, the values for all the ionic liquids fall below the sulfolane line. Once again, this is due to the high molecular weight of the ionic liquids. Consequently, a larger mass of ionic liquid than sulfolane would be necessary to separate heptane from toluene, which could increase operating and capital costs. To overcome the general trend for ionic liquids of high selectivity but low distribution ratio, Larriba et al.158,159 proposed using a mixture of ionic liquids: one with high selectivity, such as [emim][DCA], and another with high toluene distribution ratio, such as [4empy][Tf2N]. There is still

Figure 7. Liquid−liquid equilibrium of the ternary heptane + toluene + [emim][Tf2N] system in (a) molar compositions and (b) mass compositions at 298.15 K (black square) Arce et al.,84 (red circle) Corderi ́ et al.; and at 313.15 K (blue triangle) Garcia et al.104

distribution ratio of toluene follows the opposite trend: a longer alkyl chain increases the distribution ratio. This difference can be explained because the ionic liquids with longer alkyl chains increase the aromatic solubility, but also increases the aliphatic solubility. This directly affects the distribution ratios, causing a larger effect for the aliphatic than the aromatic. For example, Garcia et al.104 report Dx,heptane = 0.180 (Dw,heptane = 0.085) and Dx,toluene = 0.786 (Dw,toluene = 0.370), producing S = 4.36, for xβtoluene = 0.816 (wβtoluene = 0.803) when using [bmim][Tf2N]. However, with [hmim][Tf2N] at xβtoluene = 0.812 (wβtoluene = G

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Figure 8. Effect of the cation alkyl chain length on the (a) selectivity, (b) distribution factor in mole fraction basis, and (c) distribution factor in mass fraction basis versus composition of toluene in the upper raffinate phase for the heptane + toluene + [Cxmim][Tf2N] system. Solid symbols, 298.15 K; open symbols, 313.15 K; (open green diamond) [mmim][Tf2N];104 (solid black square) [emim][Tf2N];84 (solid gray square) [emim][Tf2N];93 (open black square) [emim][Tf2N];104 (open red triangle) [bmim][Tf2N];176 (solid blue circle) [hmim][Tf2N];91 (open blue circle) [hmim][Tf2N].104

Figure 9. Effect of the anion on the (a) selectivity and (b) distribution factor in mole fraction basis and (c) distribution factor in mass fraction basis versus composition of toluene in the upper raffinate phase for the heptane + toluene + [emim]+ cation-based ionic liquids systems. Solid symbols, 298.15 K; open symbols, 313.15 K. (Solid black square) [Tf2N]−;84 (open black square) [Tf2N]−;104 (solid blue circle) [DEP]−;87 (solid green diamond) [OAc]−;92 (open orange triangle) [CH 3 SO 3 ] − ; 107 (open purple star) [CF 3 SO 3 ] − ; 107 (cross) [CHF2CF2SO3]−;107 (solid red inverted triangle) [ESO4]−;113 (open red inverted triangle) [ESO4]−;41 (×) [SCN]−;117 (asterisk) [DCA]−;118 (dash) [TCM]−;118 and (black square with line) sulfolane.36

a need to expand the knowledge of ionic liquids for use in aromatic/aliphatic separation processes in order to improve the mass distribution ratios using low cost ionic liquids. A broader overview on the use of different ionic liquids in several types of aliphatic/aromatic mixtures and the effect of diverse modifications in the molecules on the distribution ratio and selectivity is analyzed in the work of Ferreira et al.177

the extraction solvent and the aromatic has to be separated to obtain a pure aromatic and to recover the solvent for reuse in the process. This separation, as mentioned above, is usually done by direct distillation or solvent stripping followed by distillation. For ionic liquids solvents, the separation by distillation should be easier due to the high thermal stability and low volatility of the ionic liquids. Despite the large number



AROMATIC + IONIC LIQUID BINARY SYSTEMS After the liquid−liquid separation is performed for the aliphatic/aromatic mixture, the extract composed mainly of H

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

calculating the selectivity of a system using data from different authors, even when no significant variations are detected in the ternary phase equilibrium diagrams. Selectivity remains the key parameter for choosing a better extraction solvent. As suggested by the results shown here, the mole fraction basis is the best way to present the ternary liquid−liquid equilibrium diagrams because it is more sensitive to changes when the extraction solvent has a high molecular weight, but the mass fraction basis is recommended for presentation of distribution ratios, due to its importance in defining the mass of solvent required to perform the separation. Further studies are needed to evaluate if the ionic liquid solvent can be recovered and reused in the process. Finally, the economic feasibility of the liquid−liquid extraction using an ionic liquid with the recovery of the ionic liquid solvent using different techniques needs to be explored in more depth in order to determine the real applicability of ionic liquids for aliphatic−aromatic separations.

of studies of liquid−liquid equilibrium of ionic liquids and aromatics, heats of mixing and excess properties for these binary systems are extremely scarce.178−180 Aromatics have a large solubility in ionic liquids but a miscibility gap is commonly observed. The high solubility of aromatics in, for example, imidazolium ionic liquids with alkyl chains is attributed to interactions of the strong electrostatic fields of the aromatic rings in both the ionic liquid and the toluene, where the van der Waals interactions of the toluene with nonpolar chains on the cation of the ionic liquid are also an important component of the wide mutual miscibility.181,182 According to Holbrey et al.183 this behavior supports the formation of clathrates, because a small interaction produces miscibility or immiscibility of the binary system and strong interactions produce the crystallization of the salt, but in this case there is a partial miscibility. Further understanding in this topic is needed because, for example, some ionic liquids, such as [P66614][Tf2N], are completely miscibility with benzene,86 so the clathrate theory is not applicable for this case. The solubility of the ionic liquid in the aromatic-rich phase is very low or even not detectable, depending on the nature of the ionic liquid and the aromatic in the binary mixture. Several works suggest the appearance of an upper critical solution temperature (UCST) in different combinations of [Tf2N]− ionic liquids with aromatics.184−187 Lachwa et al.186 reported liquid−liquid equilibrium for binary systems, including [Cnmim][Tf2N] + benzene, toluene, and α-methylstyrene where the UCST is also observed and a detectable concentration of ionic liquid is detected over 300 K for benzene and over 390 K for α-methylstyrene in [hmim][Tf2N]. An increase in the alkyl chain length for the imidazolium cation produces a decrease in the immiscibility area due to the increasing aromatic solubility in the ionic liquid.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +1 (574) 631-5847. Fax: +1 (574) 631-8366. Present Address †

Laboratory of Thermodynamics, Departament of Biochemical and Chemical Engineering, Technische Universität Dortmund, Emil-Figge-Str. 70, D-44227 Dortmund, Germany Funding

We acknowledge support from the University of Notre Dame Incropera-Remick Fund for Excellence. Notes



The authors declare no competing financial interest.



SUMMARY Ionic liquids are potentially competitive alternatives to currently used organic solvents for successfully separating aromatics from aliphatics. Many ionic liquids have densities and surface tensions that would be acceptable. For the [emim]+ cation, there are several choices of anions (e.g., [SCN]−, [DCA]−, [TCM]− and [Tf2N]−), which yield ionic liquids with sufficiently low viscosities (only 2 or 3 times that of sulfolane). Some anion−cation combinations show comparable or better selectivity than sulfolane, but the challenge of decreasing the amount of solvent used due to their low mass distribution ratio is still an issue to consider. The [emim]+ based ionic liquids with better selectivities than sulfolane involve [SCN]−, [CH3SO3]−, and [DCA]− anions. Selectivities of [emim][SCN] and [emim][DCA] are over four times that of sulfolane but their mass distribution ratios are only about 30% of the value for sulfolane. The [emim]+ ionic liquids with the best mass distribution ratios have the following anions: [TCM]−, [Tf2N]−, and [DEP]−. [emim][TCM] and [emim][Tf2N] have double the selectivity of sulfolane, but their mass distribution ratios are still a bit lower (0.64 and 0.52 times the value of the sulfolane, respectively). Nonetheless, [emim][Tf2N] and [emim][TCM] appear to be the better choices for heptane/toluene separation from among the ionic liquids considered in this work, due to their high selectivity and reasonably high mass distribution ratios. Selectivity is very sensitive to small variations in the experimental liquid−liquid equilibrium compositions, so large differences can be observed under certain conditions when

REFERENCES

(1) Weissermel, K.; Arpe, H.-J. AromaticsProduction and Conversion. In Industrial Organic Chemistry; Wiley-VCH Verlag GmbH: 2008; pp 313−336. (2) Robinson, P. Petroleum Processing Overview. In Practical Advances in Petroleum Processing SE-1; Hsu, C., Robinson, P., Eds.; Springer: New York, 2006; pp 1−78. (3) Hombourger, T.; Gouzien, L.; Mikitenko, P.; Bonfils, P. Solvent Extraction in the Oil Industry. In Petroleum Refining, Vol. 2−Separation Processes; Wauquier, J.-P., Ed.; Editions Technip: 2000; pp 359−546. (4) Matar, S.; Hatch, L. F. Crude Oil Processing and Production of Hydrocarbon Intermediates. In Chemistry of Petrochemical Processes; 2nd ed.; Gulf Professional Publishing: Woburn, 2001; pp 49−110. (5) Meindersma, G. W. Extraction of Aromatics from Naphtha with Ionic Liquids. Thesis, University of Twente, 2005. (6) Ashcroft, S. J.; Clayton, A. D.; Shearn, R. B. Liquid-Liquid Equilibriums for Three Ternary and Six Quaternary Systems Containing Sulfolane, N-Heptane, Toluene, 2-Propanol, and Water at 303.15 K. J. Chem. Eng. Data 1982, 27, 148−151. (7) Ashour, I.; Abu-Eishah, S. I. Liquid−Liquid Equilibria for Cyclohexane + Ethylbenzene + Sulfolane at (303.15, 313.15, and 323.15) K. J. Chem. Eng. Data 2006, 51, 859−863. (8) Ashour, I.; Abu-Eishah, S. I. Liquid−Liquid Equilibria of Ternary and Six-Component Systems Including Cyclohexane, Benzene, Toluene, Ethylbenzene, Cumene, and Sulfolane at 303.15 K. J. Chem. Eng. Data 2006, 51, 1717−1722. (9) Cassell, G. W.; Hassan, M. M.; Hines, A. L. Correlation of the Phase Equilibrium Data for the Heptane-Toluene-Sulfolane and Heptane-Xylene-Sulfolane Systems. J. Chem. Eng. Data 1989, 34, 434−438. I

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Xylene, Propanol, Sulfolane, and Water at T = 303.15 K. J. Chem. Thermodyn. 2007, 39, 1085−1089. (30) Mohsen-Nia, M.; Modarress, H.; Doulabi, F.; Bagheri, H. Liquid + Liquid Equilibria for Ternary Mixtures of (solvent + Aromatic Hydrocarbon + Alkane). J. Chem. Thermodyn. 2005, 37, 1111−1118. (31) Mondragón-Garduño, M.; Romero-Martínez, A.; Trejo, A. Liquidliquid Equilibria for Ternary Systems. I. C6-Isomers + Sulfolane + Toluene at 298.15 K. Fluid Phase Equilib. 1991, 64, 291− 303. (32) Naithani, J.; Khanna, M. K.; Nanoti, S. M.; Rawat, B. S. Quaternary Liquid-Liquid Equilibrium Studies on HydrocarbonSolvent Systems. J. Chem. Eng. Data 1992, 37, 104−106. (33) Rappel, R.; de Góis, L. M. N.; Mattedi, S. Liquid−liquid Equilibria Data for Systems Containing Aromatic + Nonaromatic + Sulfolane at 308.15 and 323.15 K. Fluid Phase Equilib. 2002, 202, 263− 276. (34) Rawat, B. S.; Gulati, I. B. Liquid-Liquid Equilibrium Studies for Separation of Aromatics. J. Appl. Chem. Biotechnol. 1976, 26, 425−435. (35) Rawat, B. S.; Prasad, G. Liquid-Liquid Equilibria for Benzene-NHeptane Systems with Triethylene Glycol, Tetraethylene Glycol, and Sulfolane Containing Water at Elevated Temperatures. J. Chem. Eng. Data 1980, 25, 227−230. (36) Tripathi, R. P.; Ram, A. R.; Rao, P. B. Liquid-Liquid Equilibriums in Ternary System Toluene-N-Heptane-Sulfolane. J. Chem. Eng. Data 1975, 20, 261−264. (37) Yorulmaz, Y.; Karpuzcu, F. Sulpholane versus Diethylene Glycol in Recovery of Aromatics. Chem. Eng. Res. Des. 1985, 63, 184−190. (38) Rawat, B. S.; Gulati, I. B. Studies on the Extraction of Aromatics with Sulpholane and Its Combination with Thiodiglycol. J. Chem. Technol. Biotechnol. 1981, 31, 25−32. (39) Sharipov, A. K. Oxides of Organic Sulfides for Refining and Petrochemistry. Chem. Technol. Fuels Oils 2001, 37, 62−72. (40) Mohammad Doulabi, F. S.; Mohsen-Nia, M. Ternary Liquid− Liquid Equilibria for Systems of (Sulfolane + Toluene or Chloronaphthalene + Octane). J. Chem. Eng. Data 2006, 51, 1431− 1435. (41) Meindersma, G. W.; Podt, A. J. G.; de Haan, A. B. Ternary Liquid−liquid Equilibria for Mixtures of Toluene+n-Heptane+an Ionic Liquid. Fluid Phase Equilib. 2006, 247, 158−168. (42) Al-Sahhaf, T. A.; Kapetanovic, E. Measurement and Prediction of Phase Equilibria in the Extraction of Aromatics from Naphtha Reformate by Tetraethylene Glycol. Fluid Phase Equilib. 1996, 118, 271−285. (43) Darwish, N. A. Liquid−liquid Equilibrium for the System NHeptane+o-Xylene+triethylene Glycol from 293.1 to 323.1 K. Chem. Eng. Process. 2002, 41, 737−741. (44) Darwish, N. A.; Abdelkarim, M. A.; Hilal, N.; Ashour, I. Analysis and Evaluation of the Liquid−Liquid Equilibrium Data of the Extraction of Aromatics from Hydrocarbons by Tetraethylene Glycol. J. Chem. Eng. Data 2003, 48, 1614−1619. (45) Mohammad Reza Seyedein Ghannad, S.; Lotfollahi, M. N.; Asl, A. H. (Liquid + Liquid) Equilibria for Mixtures of (ethylene Glycol + Benzene + Cyclohexane) at Temperatures (298.15, 308.15, and 318.15) K. J. Chem. Thermodyn. 2011, 43, 329−333. (46) Graham, H. L. Extraction of Benzene with Diand Tri-Ethylene Glycols. J. Chem. Eng. Data 1962, 7, 214−217. (47) Hughes, M. A.; Haoran, Y. Liquid-Liquid Equilibria for Separation of Toluene from Heptane by Benzyl Alcohol Tri(ethylene Glycol) Mixtures. J. Chem. Eng. Data 1990, 35, 467−471. (48) Ashour, I.; Abdulkarim, M. A. Liquid−Liquid Equilibrium for the System Heptane + O-Xylene + Diethylene Glycol over the Temperature Range of 288.15 to 318.15 K. J. Chem. Eng. Data 2005, 50, 924−927. (49) Al Qattan, M. A.; Al-Sahhaf, T. A.; Fahim, M. A. Liquid-Liquid Equilibria in Some Binary and Ternary Mixtures with Tetraethylene Glycol. J. Chem. Eng. Data 1994, 39, 111−113. (50) Wang, W.; Gou, Z.; Zhu, S. Liquid−Liquid Equilibria for Aromatics Extraction Systems with Tetraethylene Glycol. J. Chem. Eng. Data 1998, 43, 81−83.

(10) Cassell, G. W.; Dural, N.; Hines, A. L. Liquid-Liquid Equilibrium of Sulfolane-Benzene-Pentane and Sulfolane-ToluenePentane. Ind. Eng. Chem. Res. 1989, 28, 1369−1374. (11) Cassell, G. W.; Hassan, M. M.; Junes, A. L. Phase Equilibria of the Cylohexane-Toluene-Sulfolane and Hexane-Toluene-Sulfolane Ternary Systems. Chem. Eng. Commun. 1989, 85, 233−243. (12) Chen, J.; Duan, L.-P.; Mi, J.-G.; Fei, W.-Y.; Li, Z.-C. Liquid− liquid Equilibria of Multi-Component Systems Including N-Hexane, N-Octane, Benzene, Toluene, Xylene and Sulfolane at 298.15 K and Atmospheric Pressure. Fluid Phase Equilib. 2000, 173, 109−119. (13) Chen, J.; Li, Z.; Duan, L. Liquid−Liquid Equilibria of Ternary and Quaternary Systems Including Cyclohexane, 1-Heptene, Benzene, Toluene, and Sulfolane at 298.15 K. J. Chem. Eng. Data 2000, 45, 689− 692. (14) De Fré, R. M.; Verhoeye, L. A. Phase Equilibria in Systems Composed of an Aliphatic and an Aromatic Hydrocarbon and Sulpholane. J. Appl. Chem. Biotechnol. 1976, 26, 469−487. (15) Hassan, M. S.; Fahim, M. A.; Mumford, C. J. Correlation of Phase Equilibria of Naphtha Reformate with Sulfolane. J. Chem. Eng. Data 1988, 33, 162−165. (16) Hauschild, T.; Knapp, H. Liquid-Liquid-Equilibria and Densities of Multicomponent Mixtures Containing Heptane-Ethylbenzene in Sulfolane. J. Solution Chem. 1991, 20, 125−138. (17) Kao, C.-F.; Lin, W.-C. Liquid−liquid Equilibria of Alkane (C10/ C12/C14)+1,4-Diisopropylbenzene+sulfolane at 323.15, 348.15 and 373.15 K. Fluid Phase Equilib. 1999, 163, 9−20. (18) Kao, C.-F.; Lin, W.-C. Liquid−liquid Equilibria of Alkane (C10−C14)+octylbenzene+sulfolane. Fluid Phase Equilib. 1999, 165, 67−77. (19) Kao, C.-F.; Lin, W.-C. Liquid−Liquid Equilibria of the Ternary Systems Dodecane + Butylbenzene + Sulfolane, Dodecane + 1,4Diisopropylbenzene + Sulfolane, and Dodecane + Octylbenzene + Sulfolane. J. Chem. Eng. Data 1999, 44, 338−342. (20) Kao, C.-F.; Lin, W.-C. Liquid−Liquid Equilibria of Alkane (C10−C14) + Butylbenzene + Sulfolane. J. Chem. Eng. Data 1999, 44, 803−808. (21) Lee, S.; Kim, H. Liquid-Liquid Equilibria for the Ternary Systems Sulfolane + Octane + Benzene, Sulfolane + Octane + Toluene and Sulfolane + Octane + P-Xylene. J. Chem. Eng. Data 1995, 40, 499−503. (22) Lee, S.; Kim, H. Liquid−Liquid Equilibria for the Ternary Systems Sulfolane + Octane + Benzene, Sulfolane + Octane + Toluene, and Sulfolane + Octane + P-Xylene at Elevated Temperatures. J. Chem. Eng. Data 1998, 43, 358−361. (23) Letcher, T. M.; Redhi, G. G.; Radloff, S. E.; Domańska, U. Liquid−Liquid Equilibria of the Ternary Mixtures with Sulfolane at 303.15 K. J. Chem. Eng. Data 1996, 41, 634−638. (24) Lin, W.-C.; Kao, N.-H. Liquid−Liquid Equilibria of Octane + (Benzene or Toluene or M-Xylene) + Sulfolane at 323.15, 348.15, and 373.15 K. J. Chem. Eng. Data 2002, 47, 1007−1011. (25) Lin, W.-C.; Yang, C.-H.; Tsai, T.-H. Influence of the Temperature on the Liquid−Liquid Equilibria of Tridecane + Butylbenzene + Sulfolane and Tridecane + 1,4-Diisopropylbenzene + Sulfolane. J. Chem. Eng. Data 2002, 47, 245−249. (26) Lin, W.-C.; Tsai, T.-H.; Lin, T.-Y.; Yang, C.-H. Influence of the Temperature on the Liquid−Liquid Equilibria of Heptane + Toluene + Sulfolane and Heptane + M-Xylene + Sulfolane. J. Chem. Eng. Data 2008, 53, 760−764. (27) Mahmoudi, J.; Lotfollahi, M. N. Liquid + Liquid) Equilibria of (sulfolane + Benzene + N-Hexane), (N-Formylmorpholine + Benzene + N-Hexane), and (sulfolane + N-Formylmorpholine + Benzene + NHexane) at Temperatures Ranging from (298.15 to 318.15) K: Experimental Results and Correlation. J. Chem. Thermodyn. 2010, 42, 466−471. (28) Masohan, A.; Nanoti, S. M.; Sharma, K. G.; Puri, S. N.; Gupta, P.; Rawat, B. S. Liquidliquid Equilibrium Studies on Hydrocarbon (C10-C20)sulfolane Systems. Fluid Phase Equilib. 1990, 61, 89−98. (29) Mohsen-Nia, M.; Paikar, I. Liquid + Liquid) Equilibria of Ternary and Quaternary Systems Containing N-Hexane, Toluene, MJ

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

(51) Aspi, K. K.; Surana, N. M.; Ethirajulu, K.; Vennila, V. Liquid− Liquid Equilibria for the Cyclohexane + Benzene + Dimethylformamide + Ethylene Glycol System. J. Chem. Eng. Data 1998, 43, 925− 927. (52) Johnson, G. C.; Francis, A. W. Ternary Liquid System, BenzeneHeptane−Diethylene Glycol. Ind. Eng. Chem. 1954, 46, 1662−1668. (53) Rodrigues Mesquita, F. M.; Pinheiro, R. S.; de Sant’Ana, H. B.; Santiago-Aguiar, R. S. Removal of Aromatic Hydrocarbons from Hydrocarbon Mixture Using Glycols at 303.15 and 333.15 K and Atmospheric Pressure: Experimental and Calculated Data by NRTL and UNIQUAC Models. Fluid Phase Equilib. 2015, 387, 135−142. (54) Mohsen-Nia, M.; Mohammad Doulabi, F. S.; Manousiouthakis, V. I. Liquid + Liquid) Equilibria for Ternary Mixtures of (ethylene Glycol + Toluene + N-Octane). J. Chem. Thermodyn. 2008, 40, 1269− 1273. (55) Radwan, G. M.; Al-Muhtaseb, S. A.; Fahim, M. A. Liquid-Liquid Equilibria for the Extraction of Aromatics from Naphtha Reformate by Dimethylformamide/ethylene Glycol Mixed Solvent. Fluid Phase Equilib. 1997, 129, 175−186. (56) Saha, M.; Rawat, B. S.; Khanna, M. K.; Nautiyal, B. R. Liquid− Liquid Equilibrium Studies on Toluene + Heptane + Solvent. J. Chem. Eng. Data 1998, 43, 422−426. (57) Salem, A. B. S. H. Liquid-Liquid Equilibria for the Systems Triethylene Glycol - Toluene - Heptane, Propylene Carbonate Toluene - Heptane and Propylene Carbonate -O-Xylene- Heptane. Fluid Phase Equilib. 1993, 86, 351−361. (58) Somekh, G. S.; Friedlander, B. O. Tetraethylene Glycol - A Superior Solvent for Aromatics Extraction. In Refining Petroleum for Chemicals; Advances in Chemistry; American Chemical Society: 1970; Vol. 97, pp 14−228. (59) Haghnazarloo, H.; Lotfollahi, M. N.; Mahmoudi, J.; Asl, A. H. Liquid−liquid Equilibria for Ternary Systems of (ethylene Glycol + Toluene + Heptane) at Temperatures (303.15, 308.15, and 313.15) K and Atmospheric Pressure: Experimental Results and Correlation with UNIQUAC and NRTL Models. J. Chem. Thermodyn. 2013, 60, 126− 131. (60) Mohsen-Nia, M.; Mohammad Doulabi, F. S. Separation of Aromatic Hydrocarbons (toluene or Benzene) from Aliphatic Hydrocarbon (n-Heptane) by Extraction with Ethylene Carbonate. J. Chem. Thermodyn. 2010, 42, 1281−1285. (61) Mohsen-Nia, M.; Doulabi, F. S. M. Liquid−liquid Equilibria for Mixtures of (ethylene Carbonate + Aromatic Hydrocarbon + Cyclohexane). Thermochim. Acta 2006, 445, 82−85. (62) Ali, S. H.; Lababidi, H. M. S.; Merchant, S. Q.; Fahim, M. A. Extraction of Aromatics from Naphtha Reformate Using Propylene Carbonate. Fluid Phase Equilib. 2003, 214, 25−38. (63) Annesini, M. C.; Gironi, F.; Marrelli, L.; Kikic, I. Liquid-Liquid Equilibria for Ternary Systems Containing Hydrocarbons and Propylene Carbonate. J. Chem. Eng. Data 1985, 30, 195−196. (64) Li, J.; Zhao, Q.; Tang, X.; Xiao, K.; Yuan, J. Liquid−Liquid Equilibria for the Systems: Heptane + Benzene + Solvent (Propylene Carbonate, N,N-Dimethylformamide, or Mixtures) at Temperatures from (303.2 to 323.2) K. J. Chem. Eng. Data 2014, 59, 3307−3313. (65) Al-Zayied, T. A.; Al-Sahhaf, T. A.; Fahim, M. A. Measurement of Phase Equilibrium in Multicomponent Systems of Aromatics with NMethylpyrrolidone and Predictions with Unifac. Fluid Phase Equilib. 1990, 61, 131−144. (66) Ferreira, P. O.; Ferreira, J. B.; Medina, A. G. Liquid-Liquid Equilibria for the System N-Methylpyrrolidone + Toluene + NHeptane: UNIFAC Interaction Parameters for N-Methylpyrrolidone. Fluid Phase Equilib. 1984, 16, 369−379. (67) Alkhaldi, K. H. A. E.; Fandary, M. S.; Al-Jimaz, A. S.; AlTuwaim, M. S.; Fahim, M. A. Liquid−liquid Equilibria of Aromatics Removal from Middle Distillate Using NMP. Fluid Phase Equilib. 2009, 286, 190−195. (68) Nagpal, J. M.; Rawat, B. S. Liquid-Liquid Equilibria for TolueneHeptane-N-Methyl Pyrrolidone and Benzene-Heptane Solvents. J. Chem. Technol. Biotechnol. 1981, 31, 146−150.

(69) DongChu, C.; HongQi, Y.; Hao, W. Liquid + Liquid) Equilibria of Three Ternary Systems: (heptane + Benzene + N-Formylmorpholine), (heptane + Toluene + N-Formylmorpholine), (heptane + Xylene + N-Formylmorpholine) from T = (298.15 to 353.15) K. J. Chem. Thermodyn. 2007, 39, 1182−1188. (70) Chen, D.; Ye, H.; Hao, W. Liquid + Liquid) Equilibria of {heptane + Xylene + N-Formylmorpholine} Ternary System. J. Chem. Thermodyn. 2007, 39, 1571−1577. (71) Seyedein Ghannad, S. M.; Lotfollahi, M. N.; Haghighi Asl, A. Measurement of (liquid + Liquid) Equilibria for Ternary Systems of (N-Formylmorpholine + Benzene + Cyclohexane) at Temperatures (303.15, 308.15, and 313.15) K. J. Chem. Thermodyn. 2011, 43, 938− 942. (72) Wang, Z.; Xia, S.; Ma, P. (Liquid + Liquid) Equilibria for the Ternary System of (N-Formylmorpholine + Ethylbenzene + 2,2,4Trimethylpentane) at Temperatures (303.15, 313.15, and 323.15) K. Fluid Phase Equilib. 2012, 328, 25−30. (73) Al Qattan, M. A.; Al-Sahhaf, T. A.; Fahim, M. A. Liquid-Liquid Equilibria in Some Binary and Ternary Mixtures with NFormylmorpholine. J. Chem. Eng. Data 1995, 40, 88−90. (74) Cincotti, A.; Murru, M.; Cao, G.; Marongiu, B.; Masia, F.; Sannia, M. Liquid−Liquid Equilibria of Hydrocarbons with NFormylmorpholine. J. Chem. Eng. Data 1999, 44, 480−483. (75) Ferreira, P. O.; Barbosa, D.; Medina, A. G. Phase Equilibria for the Separation of Aromatic and Nonaromatic Compounds Using Mixed Solvents. Part I. The System N-heptanetolueneNMethylpyrrolidone/monoethyleneglycol. Fluid Phase Equilib. 1984, 15, 309−322. (76) Gramajo de Doz, M. B.; Cases, A. M.; Sólimo, H. N. (Liquid + Liquid) Equilibria of the Quaternary System Methanol + Isooctane + Cyclohexane + Benzene at T = 303.15 K. Fluid Phase Equilib. 2010, 289, 15−19. (77) Salem, A. B. S. H.; Hamad, E. Z. Liquid-Liquid Equilibrium of the Five Component System of N-Hexane-N-Heptane-Toluene-OXylene-Propylene Carbonate. Fluid Phase Equilib. 1995, 108, 231− 241. (78) Chen, J.; Mi; Fei; Li. Liquid−Liquid Equilibria of Quaternary and Quinary Systems Including Sulfolane at 298.15 K. J. Chem. Eng. Data 2001, 46, 169−171. (79) Salem, A. B. S. H.; Hamad, E. Z.; Al-Naafa, M. A. Quaternary Liquid-Liquid Equilibrium of N-Heptane-Toluene-O-Xylene-Propylene Carbonate. Ind. Eng. Chem. Res. 1994, 33, 689−692. (80) Santiago, R. S.; Aznar, M. Liquid−liquid Equilibria for Quaternary Mixtures of Nonane + Undecane + (benzene or Toluene or M-Xylene) + Sulfolane at 298.15 and 313.15 K. Fluid Phase Equilib. 2007, 253, 137−141. (81) Santiago, R. S.; Aznar, M. Quinary Liquid−liquid Equilibria for Mixtures of Nonane + Undecane + Two Pairs of Aromatics (benzene/ toluene/m-Xylene) + Sulfolane at 298.15 and 313.15 K. Fluid Phase Equilib. 2007, 259, 71−76. (82) Arce, A.; Earle, M. J.; Rodriguez, H.; Seddon, K. R. Separation of Aromatic Hydrocarbons from Alkanes Using the Ionic Liquid 1-Ethyl3-Methylimidazoliumbis{(trifluoromethyl) Sulfonyl}amide. Green Chem. 2007, 9, 70−74. (83) Arce, A.; Earle, M. J.; Rodríguez, H.; Seddon, K. R. Separation of Benzene and Hexane by Solvent Extraction with 1-Alkyl-3Methylimidazolium Bis{(trifluoromethyl)sulfonyl}amide Ionic Liquids: Effect of the Alkyl-Substituent Length†. J. Phys. Chem. B 2007, 111, 4732−4736. (84) Arce, A.; Earle, M. J.; Rodriguez, H.; Seddon, K. R.; Soto, A. 1Ethyl-3-Methylimidazolium Bis{(trifluoromethyl)sulfonyl}amide as Solvent for the Separation of Aromatic and Aliphatic Hydrocarbons by Liquid Extraction - Extension to C7- and C8-Fractions. Green Chem. 2008, 10, 1294−1300. (85) Arce, A.; Earle, M. J.; Rodríguez, H.; Seddon, K. R.; Soto, A. Isomer Effect in the Separation of Octane and Xylenes Using the Ionic Liquid 1-Ethyl-3-Methylimidazolium Bis{(trifluoromethyl)sulfonyl}amide. Fluid Phase Equilib. 2010, 294, 180−186. K

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

(86) Arce, A.; Earle, M. J.; Katdare, S. P.; Rodriguez, H.; Seddon, K. R. Application of Mutually Immiscible Ionic Liquids to the Separation of Aromatic and Aliphatic Hydrocarbons by Liquid Extraction: A Preliminary Approach. Phys. Chem. Chem. Phys. 2008, 10, 2538−2542. (87) Cai, F.; Zhu, W.; Wang, Y.; Wang, T.; Xiao, G. Dialkylphosphate-Based Ionic Liquids as Solvents to Extract Toluene from Heptane. J. Chem. Eng. Data 2015, 60, 1776−1780. (88) Calvar, N.; Domínguez, I.; Gómez, E.; Palomar, J.; Domínguez, Á . Evaluation of Ionic Liquids as Solvent for Aromatic Extraction: Experimental, Correlation and COSMO-RS Predictions. J. Chem. Thermodyn. 2013, 67, 5−12. (89) Calvar, N.; Domínguez, I.; Gómez, E.; Domínguez, Á . Separation of Binary Mixtures Aromatic + Aliphatic Using Ionic Liquids: Influence of the Structure of the Ionic Liquid, Aromatic and Aliphatic. Chem. Eng. J. 2011, 175, 213−221. (90) Cassol, C. C.; Umpierre, A. P.; Ebeling, G.; Ferrera, B.; Chiaro, S. S. X.; Dupont, J. On the Extraction of Aromatic Compounds from Hydrocarbons by Imidazolium Ionic Liquids. Int. J. Mol. Sci. 2007, 8, 593. (91) Corderí, S.; González, E. J.; Calvar, N.; Domínguez, Á . Application of [HMim][NTf2], [HMim][TfO] and [BMim][TfO] Ionic Liquids on the Extraction of Toluene from Alkanes: Effect of the Anion and the Alkyl Chain Length of the Cation on the LLE. J. Chem. Thermodyn. 2012, 53, 60−66. (92) Corderí, S.; Gómez, E.; Calvar, N.; Domínguez, Á . Measurement and Correlation of Liquid−Liquid Equilibria for Ternary and Quaternary Systems of Heptane, Cyclohexane, Toluene, and [EMim][OAc] at 298.15 K. Ind. Eng. Chem. Res. 2014, 53, 9471−9477. (93) Corderí, S.; Calvar, N.; Gómez, E.; Domínguez, A. Capacity of Ionic Liquids [EMim][NTf2] and [EMpy][NTf2] for Extraction of Toluene from Mixtures with Alkanes: Comparative Study of the Effect of the Cation. Fluid Phase Equilib. 2012, 315, 46−52. (94) Corderí, S.; Calvar, N.; Gómez, E.; Domínguez, Á . Quaternary (liquid + Liquid) Equilibrium Data for the Extraction of Toluene from Alkanes Using the Ionic Liquid [EMim][MSO4]. J. Chem. Thermodyn. 2014, 76, 79−86. (95) Domańska, U.; Pobudkowska, A.; Ż ołek-Tryznowska, Z. Effect of an Ionic Liquid (IL) Cation on the Ternary System (IL + P-Xylene + Hexane) at T = 298.15 K†. J. Chem. Eng. Data 2007, 52, 2345− 2349. (96) Domínguez, I.; González, E. J.; González, R.; Domínguez, Á . Extraction of Benzene from Aliphatic Compounds Using Commercial Ionic Liquids as Solvents: Study of the Liquid−Liquid Equilibrium at T = 298.15 K. J. Chem. Eng. Data 2011, 56, 3376−3383. (97) Domínguez, I.; González, E. J.; Domínguez, Á . Liquid Extraction of Aromatic/cyclic Aliphatic Hydrocarbon Mixtures Using Ionic Liquids as Solvent: Literature Review and New Experimental LLE Data. Fuel Process. Technol. 2014, 125, 207−216. (98) Domínguez, I.; Calvar, N.; Gómez, E.; Domínguez, Á . Separation of Toluene from Cyclic Hydrocarbons Using 1-Butyl-3Methylimidazolium Methylsulfate Ionic Liquid at T = 298.15K and Atmospheric Pressure. J. Chem. Thermodyn. 2011, 43, 705−710. (99) Domínguez, I.; Calvar, N.; Gómez, E.; Domínguez, Á . Liquid− Liquid Extraction of Aromatic Compounds from Cycloalkanes Using 1-Butyl-3-Methylimidazolium Methylsulfate Ionic Liquid. J. Chem. Eng. Data 2013, 58, 189−196. (100) Domínguez, I.; González, E. J.; Palomar, J.; Domínguez, Á . Phase Behavior of Ternary Mixtures {aliphatic Hydrocarbon + Aromatic Hydrocarbon + Ionic Liquid}: Experimental LLE Data and Their Modeling by COSMO-RS. J. Chem. Thermodyn. 2014, 77, 222− 229. (101) Domínguez, I.; Requejo, P. F.; Canosa, J.; Domínguez, Á . Liquid + Liquid) Equilibrium at T = 298.15 K for Ternary Mixtures of Alkane + Aromatic Compounds + Imidazolium-Based Ionic Liquids. J. Chem. Thermodyn. 2014, 74, 138−143. (102) Dukhande, V. A.; Choksi, T. S.; Sabnis, S. U.; Patwardhan, A. W.; Patwardhan, A. V. Separation of Toluene from N-Heptane Using Monocationic and Dicationic Ionic Liquids. Fluid Phase Equilib. 2013, 342, 75−81.

(103) García, S.; Larriba, M.; García, J.; Torrecilla, J. S.; Rodríguez, F. (Liquid+liquid) Equilibrium for the Ternary Systems {heptane +toluene+1-Allyl-3-Methylimidazolium Bis(trifluoromethylsulfonyl)imide} and {heptane+toluene+1-Methyl-3-Propylimidazolium Bis(trifluoromethylsulfonyl)imide} Ionic Liquids. J. Chem. Thermodyn. 2011, 43, 1641−1645. (104) García, S.; Larriba, M.; García, J.; Torrecilla, J. S.; Rodríguez, F. Liquid−Liquid Extraction of Toluene from Heptane Using 1-Alkyl-3Methylimidazolium Bis(trifluoromethylsulfonyl)imide Ionic Liquids. J. Chem. Eng. Data 2011, 56, 113−118. (105) García, S.; Larriba, M.; García, J.; Torrecilla, J. S.; Rodríguez, F. 1-Alkyl-2,3-Dimethylimidazolium Bis(trifluoromethylsulfonyl)imide Ionic Liquids for the Liquid−Liquid Extraction of Toluene from Heptane. J. Chem. Eng. Data 2011, 56, 3468−3474. (106) García, J.; Fernández, A.; Torrecilla, J. S.; Oliet, M.; Rodríguez, F. Liquid−liquid Equilibria for {hexane + Benzene + 1-Ethyl-3Methylimidazolium Ethylsulfate} at (298.2, 313.2 and 328.2) K. Fluid Phase Equilib. 2009, 282, 117−120. (107) García, S.; García, J.; Larriba, M.; Torrecilla, J. S.; Rodríguez, F. Sulfonate-Based Ionic Liquids in the Liquid−Liquid Extraction of Aromatic Hydrocarbons. J. Chem. Eng. Data 2011, 56, 3188−3193. (108) García, J.; Fernández, A.; Torrecilla, J. S.; Oliet, M.; Rodríguez, F. Ternary Liquid−Liquid Equilibria Measurement for Hexane and Benzene with the Ionic Liquid 1-Butyl-3-Methylimidazolium Methylsulfate at T = (298.2, 313.2, and 328.2) K. J. Chem. Eng. Data 2010, 55, 258−261. (109) González, E. J.; Calvar, N.; Dominguez, I.; Domínguez, Á . Liquid + Liquid) Equilibrium Data for the Ternary Systems (cycloalkane + Ethylbenzene + 1-Ethyl-3-Methylimidazolim Ethylsulfate) at T = 0298.15 K and Atmospheric Pressure. J. Chem. Thermodyn. 2011, 43, 725−730. (110) González, E. J.; Calvar, N.; Gómez, E.; Domínguez, Á . Application of [EMim][ESO4] Ionic Liquid as Solvent in the Extraction of Toluene from Cycloalkanes: Study of Liquid−liquid Equilibria at T = 298.15K. Fluid Phase Equilib. 2011, 303, 174−179. (111) González, E. J.; Calvar, N.; González, B.; Domínguez, Á . Liquid Extraction of Benzene from Its Mixtures Using 1-Ethyl-3-Methylimidazolium Ethylsulfate as a Solvent. J. Chem. Eng. Data 2010, 55, 4931−4936. (112) González, E. J.; Calvar, N.; Gómez, E.; Domínguez, Á . Separation of Benzene from Linear Alkanes (C6−C9) Using 1-Ethyl-3Methylimidazolium Ethylsulfate at T = 298.15 K. J. Chem. Eng. Data 2010, 55, 3422−3427. (113) González, E. J.; Calvar, N.; Domínguez, I.; Domínguez, Á . Extraction of Toluene from Aliphatic Compounds Using an Ionic Liquid as Solvent: Influence of the Alkane on the (liquid+liquid) Equilibrium. J. Chem. Thermodyn. 2011, 43, 562−568. (114) González, E. J.; González, B.; Calvar, N.; Domínguez, Á . Study of [EMim][ESO4] Ionic Liquid as Solvent in the Liquid−liquid Extraction of Xylenes from Their Mixtures with Hexane. Fluid Phase Equilib. 2011, 305, 227−232. (115) Gutierrez, J. P.; Meindersma, W.; de Haan, A. B. Binary and Ternary (liquid + Liquid) Equilibrium for {methylcyclohexane (1) + Toluene (2) + 1-Hexyl-3-Methylimidazolium Tetracyanoborate (3)/1Butyl-3-Methylimidazolium Tetracyanoborate (3)}. J. Chem. Thermodyn. 2011, 43, 1672−1677. (116) Hansmeier, A. R.; Ruiz, M. M.; Meindersma, G. W.; de Haan, A. B. Liquid−Liquid Equilibria for the Three Ternary Systems (3Methyl-N-Butylpyridinium Dicyanamide + Toluene + Heptane), (1Butyl-3-Methylimidazolium Dicyanamide + Toluene + Heptane) and (1-Butyl-3-Methylimidazolium Thiocyanate + Toluene + Heptane) at T = (31. J. Chem. Eng. Data 2010, 55, 708−713. (117) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Selective Extraction of Toluene from N-Heptane Using [emim][SCN] and [bmim][SCN] Ionic Liquids as Solvents. J. Chem. Thermodyn. 2014, 79, 266−271. (118) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Liquid− Liquid Extraction of Toluene from Heptane Using [emim][DCA], L

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

[bmim][DCA], and [emim][TCM] Ionic Liquids. Ind. Eng. Chem. Res. 2013, 52, 2714−2720. (119) Letcher, T. M.; Deenadayalu, N. Ternary Liquid−liquid Equilibria for Mixtures of 1-Methyl-3-Octyl-Imidazolium Chloride+ Benzene + an Alkane at T = 298.2 K and 1 Atm. J. Chem. Thermodyn. 2003, 35, 67−76. (120) Letcher, T. M.; Reddy, P. Ternary (liquid + Liquid) Equilibria for Mixtures of 1-Hexyl-3-Methylimidazolium (tetrafluoroborate or Hexafluorophosphate) + Benzene + an Alkane at T = 298.2 K and p = 0.1 MPa. J. Chem. Thermodyn. 2005, 37, 415−421. (121) Lu, Y.; Yang, X.; Luo, G. Liquid−Liquid Equilibria for Benzene + Cyclohexane + 1-Butyl-3-Methylimidazolium Hexafluorophosphate. J. Chem. Eng. Data 2010, 55, 510−512. (122) Maduro, R. M.; Aznar, M. Liquid−liquid Equilibrium of Ternary Systems 1-Octyl-3-Methylimidazolium Hexafluorophosphate + Aromatic + Aliphatic Hydrocarbons. Fluid Phase Equilib. 2010, 296, 88−94. (123) Maduro, R. M.; Aznar, M. Liquid−liquid Equilibrium of Ternary Systems 1-Butyl-3-Methylimidazolium Hexafluorophosphate + Aromatic + Aliphatic. Fluid Phase Equilib. 2008, 265, 129−138. (124) Revelli, A.-L.; Mutelet, F.; Jaubert, J.-N. Extraction of Benzene or Thiophene from N-Heptane Using Ionic Liquids. NMR and Thermodynamic Study. J. Phys. Chem. B 2010, 114, 4600−4608. (125) Selvan, M. S.; McKinley, M. D.; Dubois, R. H.; Atwood, J. L. Liquid-Liquid Equilibria for Toluene + Heptane + 1-Ethyl-3Methylimidazolium Triiodide and Toluene + Heptane + 1-Butyl-3Methylimidazolium Triiodide. J. Chem. Eng. Data 2000, 45, 841−845. (126) Wang, R.; Wang, J.; Meng, H.; Li, C.; Wang, Z. Liquid−Liquid Equilibria for Benzene + Cyclohexane + 1-Methyl-3-Methylimidazolium Dimethylphosphate or + 1-Ethyl-3-Methylimidazolium Diethylphosphate. J. Chem. Eng. Data 2008, 53, 1159−1162. (127) Wang, R.; Li, C.; Meng, H.; Wang, J.; Wang, Z. Ternary Liquid−Liquid Equilibria Measurement for Benzene + Cyclohexane + N-Methylimidazole, or N-Ethylimidazole, or N-Methylimidazolium Dibutylphosphate at 298.2 K and Atmospheric Pressure. J. Chem. Eng. Data 2008, 53, 2170−2174. (128) Zhou, T.; Wang, Z.; Ye, Y.; Chen, L.; Xu, J.; Qi, Z. Deep Separation of Benzene from Cyclohexane by Liquid Extraction Using Ionic Liquids as the Solvent. Ind. Eng. Chem. Res. 2012, 51, 5559. (129) Abu-Eishah, S. I.; Dowaidar, A. M. Liquid−Liquid Equilibrium of Ternary Systems of Cyclohexane + (Benzene, + Toluene, + Ethylbenzene, or + O-Xylene) + 4-Methyl-N-Butyl Pyridinium Tetrafluoroborate Ionic Liquid at 303.15 K. J. Chem. Eng. Data 2008, 53, 1708−1712. (130) Al-Tuwaim, M. S.; Alkhaldi, K. H. A. E.; Fandary, M. S.; AlJimaz, A. S. Extraction of Propylbenzene or Butylbenzene from Dodecane Using 4-Methyl-N-Butylpyridinium Tetrafluoroborate, [mebupy][BF4], as an Ionic Liquid at Different Temperatures. J. Chem. Thermodyn. 2011, 43, 1804−1809. (131) Arce, A.; Earle, M. J.; Rodriguez, H.; Seddon, K. R.; Soto, A. Bis{(trifluoromethyl)sulfonyl}amide Ionic Liquids as Solvents for the Extraction of Aromatic Hydrocarbons from Their Mixtures with Alkanes: Effect of the Nature of the Cation. Green Chem. 2009, 11, 365−372. (132) García, J.; García, S.; Torrecilla, J. S.; Rodríguez, F. Solvent Extraction of Toluene from Heptane with the Ionic Liquids NEthylpyridinium Bis(trifluoromethylsulfonyl)imide and Z-Methyl-NEthylpyridinium Bis(trifluoromethylsulfonyl)imide (z = 2, 3, or 4) at T = 313.2 K. J. Chem. Eng. Data 2010, 55, 4937−4942. (133) García, J.; García, S.; Torrecilla, J. S.; Oliet, M.; Rodríguez, F. Separation of Toluene and Heptane by Liquid−liquid Extraction Using Z-Methyl-N-Butylpyridinium Tetrafluoroborate Isomers (z = 2, 3, or 4) at T = 313.2 K. J. Chem. Thermodyn. 2010, 42, 1004−1008. (134) García, J.; García, S.; Torrecilla, J. S.; Rodríguez, F. NButylpyridinium Bis-(trifluoromethylsulfonyl)imide Ionic Liquids as Solvents for the Liquid−liquid Extraction of Aromatics from Their Mixtures with Alkanes: Isomeric Effect of the Cation. Fluid Phase Equilib. 2011, 301, 62−66.

(135) García, J.; García, S.; Torrecilla, J. S.; Oliet, M.; Rodríguez, F. Liquid−Liquid Equilibria for the Ternary Systems {Heptane + Toluene + N-Butylpyridinium Tetrafluoroborate or N-Hexylpyridinium Tetrafluoroborate} at T = 313.2 K. J. Chem. Eng. Data 2010, 55, 2862−2865. (136) Gómez, E.; Domínguez, I.; Calvar, N.; Domínguez, Á . Separation of Benzene from Alkanes by Solvent Extraction with 1Ethylpyridinium Ethylsulfate Ionic Liquid. J. Chem. Thermodyn. 2010, 42, 1234−1239. (137) Gómez, E.; Domínguez, I.; González, B.; Domínguez, Á . Liquid−Liquid Equilibria of the Ternary Systems of Alkane + Aromatic + 1-Ethylpyridinium Ethylsulfate Ionic Liquid at T = (283.15 and 298.15) K. J. Chem. Eng. Data 2010, 55, 5169−5175. (138) Gómez, E.; Domínguez, I.; Calvar, N.; Palomar, J.; Domínguez, Á . Experimental Data, Correlation and Prediction of the Extraction of Benzene from Cyclic Hydrocarbons Using [Epy][ESO4] Ionic Liquid. Fluid Phase Equilib. 2014, 361, 83−92. (139) González, E. J.; Calvar, N.; Canosa, J.; Domínguez, Á . Effect of the Chain Length on the Aromatic Ring in the Separation of Aromatic Compounds from Methylcyclohexane Using the Ionic Liquid 1-Ethyl3-Methylpyridinium Ethylsulfate. J. Chem. Eng. Data 2010, 55, 2289− 2293. (140) González, E. J.; Domínguez, I.; González, B.; Canosa, J. Liquid−liquid Equilibria for Ternary Systems of {cyclohexane + Aromatic Compounds + 1-Ethyl-3-Methylpyridinium Ethylsulfate}. Fluid Phase Equilib. 2010, 296, 213−218. (141) González, E. J.; Calvar, N.; González, B.; Domínguez, Á . Measurement and Correlation of Liquid−liquid Equilibria for Ternary Systems {cyclooctane + Aromatic Hydrocarbon + 1-Ethyl-3Methylpyridinium Ethylsulfate} at T = 298.15 K and Atmospheric Pressure. Fluid Phase Equilib. 2010, 291, 59−65. (142) González, E. J.; Calvar, N.; González, B.; Domínguez, Á . Liquid−Liquid Equilibrium for Ternary Mixtures of Hexane + Aromatic Compounds + [EMpy][ESO4] at T = 298.15 K. J. Chem. Eng. Data 2010, 55, 633−638. (143) González, E. J.; Calvar, N.; González, B.; Domínguez, Á . Separation of Toluene from Alkanes Using 1-Ethyl-3-Methylpyridinium Ethylsulfate Ionic Liquid at T = 298.15 K and Atmospheric Pressure. J. Chem. Thermodyn. 2010, 42, 752−757. (144) González, E. J.; Calvar, N.; Gómez, E.; Domínguez, Á . Separation of Benzene from Alkanes Using 1-Ethyl-3-Methylpyridinium Ethylsulfate Ionic Liquid at Several Temperatures and Atmospheric Pressure: Effect of the Size of the Aliphatic Hydrocarbons. J. Chem. Thermodyn. 2010, 42, 104−109. (145) González, E. J.; Calvar, N.; González, B.; Domínguez, Á . Liquid + Liquid) Equilibria for Ternary Mixtures of (alkane + Benzene + [EMpy] [ESO4]) at Several Temperatures and Atmospheric Pressure. J. Chem. Thermodyn. 2009, 41, 1215−1221. (146) Hansmeier, A. R.; Jongmans, M.; Wytze Meindersma, G.; de Haan, A. B. LLE Data for the Ionic Liquid 3-Methyl-N-Butyl Pyridinium Dicyanamide with Several Aromatic and Aliphatic Hydrocarbons. J. Chem. Thermodyn. 2010, 42, 484−490. (147) Meindersma, G. W.; Podt, A.; de Haan, A. B. Ternary Liquid− Liquid Equilibria for Mixtures of an Aromatic + an Aliphatic Hydrocarbon + 4-Methyl-N-Butylpyridinium Tetrafluoroborate. J. Chem. Eng. Data 2006, 51, 1814−1819. (148) Meindersma, G. W.; Simons, B. T. J.; de Haan, A. B. Physical Properties of 3-Methyl-N-Butylpyridinium Tetracyanoborate and 1Butyl-1-Methylpyrrolidinium Tetracyanoborate and Ternary LLE Data of [3-mebupy]B(CN)4 with an Aromatic and an Aliphatic Hydrocarbon at T = 303.2 and 328.2 K and. J. Chem. Thermodyn. 2011, 43, 1628−1640. (149) Mirkhani, S. A.; Vossoughi, M.; Pazuki, G. R.; Safekordi, A. A.; Heydari, A.; Akbari, J.; Yavari, M. Liquid + Liquid) Equilibrium for Ternary Mixtures of {heptane + Aromatic Compounds + [EMpy][ESO4]} at T = 298.15 K. J. Chem. Thermodyn. 2011, 43, 1530−1534. (150) Requejo, P. F.; Gómez, E.; Calvar, N.; Domínguez, Á . Application of Pyrrolidinium-Based Ionic Liquid as Solvent for the M

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

Mixtures: Liquid−Liquid Phase Split Separation and Modelling. Philos. Trans. R. Soc., A 2015, 373, 20150011. (170) Müller, E.; Berger, R.; Blass, E.; Sluyts, D.; Pfennig, A. Liquid− Liquid Extraction. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: 2000. (171) Kislik, V. S. Advances in Development of Solvents for Liquid− Liquid Extraction. In Solvent Extraction; Kislik, V. S., Ed.; Elsevier: Amsterdam, 2012; Chapter 13, pp 451−481. (172) Tilstam, U. Sulfolane: A Versatile Dipolar Aprotic Solvent. Org. Process Res. Dev. 2012, 16, 1273−1278. (173) Hand, D. B. Dineric Distribution. J. Phys. Chem. 1929, 34, 1961−2000. (174) Othmer, D.; Tobias, P. Liquid-Liquid Extraction Data - The Line Correlation. Ind. Eng. Chem. 1942, 34, 693−696. (175) Carniti, P.; Cori, L.; Ragaini, V. A Critical Analysis of the Hand and Othmer-Tobias Correlations. Fluid Phase Equilib. 1978, 2, 39−47. (176) García, S.; Larriba, M.; García, J.; Torrecilla, J. S.; Rodríguez, F. Liquid−Liquid Extraction of Toluene from Heptane Using 1-Alkyl-3Methylimidazolium Bis(trifluoromethylsulfonyl)imide Ionic Liquids. J. Chem. Eng. Data 2011, 56, 113−118. (177) Ferreira, A. R.; Freire, M. G.; Ribeiro, J. C.; Lopes, F. M.; Crespo, J. G.; Coutinho, J. A. P. Overview of the Liquid−Liquid Equilibria of Ternary Systems Composed of Ionic Liquid and Aromatic and Aliphatic Hydrocarbons, and Their Modeling by COSMO-RS. Ind. Eng. Chem. Res. 2012, 51, 3483−3507. (178) González, E. J.; Requejo, P. F.; Maia, F. M.; Domínguez, Á .; Macedo, E. A. Solubility, Density and Excess Molar Volume of Binary Mixtures of Aromatic Compounds and Common Ionic Liquids at T = 283.15 K and Atmospheric Pressure. Phys. Chem. Liq. 2015, 53, 419− 428. (179) Zhong, Y.; Wang, H.; Diao, K. Densities and Excess Volumes of Binary Mixtures of the Ionic Liquid 1-Butyl-3-Methylimidazolium Hexafluorophosphate with Aromatic Compound at T = (298.15 to 313.15) K. J. Chem. Thermodyn. 2007, 39, 291−296. (180) Marczak, W.; Verevkin, S.; Heintz, A. Enthalpies of Solution of Organic Solutes in the Ionic Liquid 1-Methyl-3-Ethyl-Imidazolium Bis(trifluoromethyl-Sulfonyl) Amide. J. Solution Chem. 2003, 32, 519− 526. (181) Gutel, T.; Santini, C. C.; Pádua, A. A. H.; Fenet, B.; Chauvin, Y.; Canongia Lopes, J. N.; Bayard, F.; Costa Gomes, M. F.; Pensado, A. S. Interaction between the π-System of Toluene and the Imidazolium Ring of Ionic Liquids: A Combined NMR and Molecular Simulation Study. J. Phys. Chem. B 2008, 113, 170−177. (182) Hanke, C. G.; Johansson, A.; Harper, J. B.; Lynden-Bell, R. M. Why Are Aromatic Compounds More Soluble than Aliphatic Compounds in Dimethylimidazolium Ionic Liquids? A Simulation Study. Chem. Phys. Lett. 2003, 374, 85−90. (183) Holbrey, J. D.; Reichert, W. M.; Nieuwenhuyzen, M.; Sheppard, O.; Hardacre, C.; Rogers, R. D. Liquid Clathrate Formation in Ionic Liquid-Aromatic Mixtures. Chem. Commun. 2003, 476−477. (184) Domańska, U.; Marciniak, A.; Królikowski, M. Phase Equilibria and Modeling of Ammonium Ionic Liquid, C2NTf2, Solutions†. J. Phys. Chem. B 2008, 112, 1218−1225. (185) Domańska, U.; Królikowski, M.; Ramjugernath, D.; Letcher, T. M.; Tumba, K. Phase Equilibria and Modeling of Pyridinium-Based Ionic Liquid Solutions. J. Phys. Chem. B 2010, 114, 15011−15017. (186) Lachwa, J.; Szydlowski, J.; Makowska, A.; Seddon, K. R.; Esperanca, J. M. S. S.; Guedes, H. J. R.; Rebelo, L. P. N. Changing from an Unusual High-Temperature Demixing to a UCST-Type in Mixtures of 1-Alkyl-3-Methylimidazolium Bis{(trifluoromethyl)sulfonyl}amide and Arenes. Green Chem. 2006, 8, 262−267. (187) Domańska, U.; Marciniak, A. Liquid Phase Behaviour of 1Hexyloxymethyl-3-Methyl-Imidazolium-Based Ionic Liquids with Hydrocarbons: The Influence of Anion. J. Chem. Thermodyn. 2005, 37, 577−585. (188) Haynes, W. M. CRC Handbook of Chemistry and Physics; CRC Press: 2015.

Liquid Extraction of Benzene from Its Mixtures with Aliphatic Hydrocarbons. Ind. Eng. Chem. Res. 2015, 54, 1342−1349. (151) Domańska, U.; Pobudkowska, A.; Królikowski, M. Separation of Aromatic Hydrocarbons from Alkanes Using Ammonium Ionic Liquid C2NTf2 at T = 298.15 K. Fluid Phase Equilib. 2007, 259, 173− 179. (152) Zhang, F.; Li, Y.; Zhang, L.; Sun, W.; Ren, Z. Selective Separation of Aromatics from Paraffins and Cycloalkanes Using Morpholinium-Based Ionic Liquid. J. Chem. Eng. Data 2015, 60, 1634−1641. (153) Pereiro, A. B.; Rodriguez, A. Application of the Ionic Liquid Ammoeng 102 for Aromatic/aliphatic Hydrocarbon Separation. J. Chem. Thermodyn. 2009, 41, 951−956. (154) Larriba, M.; Navarro, P.; González, E. J.; García, J.; Rodríguez, F. Dearomatization of Pyrolysis Gasolines from Mild and Severe Cracking by Liquid−liquid Extraction Using a Binary Mixture of [4empy][Tf2N] and [emim][DCA] Ionic Liquids. Fuel Process. Technol. 2015, 137, 269−282. (155) Larriba, M.; Navarro, P.; González, E. J.; García, J.; Rodríguez, F. Separation of BTEX from a Naphtha Feed to Ethylene Crackers Using a Binary Mixture of [4empy][Tf2N] and [emim][DCA] Ionic Liquids. Sep. Purif. Technol. 2015, 144, 54−62. (156) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Liquid− Liquid Extraction of BTEX from Reformer Gasoline Using Binary Mixtures of [4empy][Tf2N] and [emim][DCA] Ionic Liquids. Energy Fuels 2014, 28, 6666−6676. (157) Sakal, S. A.; Shen, C.; Li, C. Liquid + Liquid) Equilibria of {benzene + Cyclohexane + Two Ionic Liquids} at Different Temperature and Atmospheric Pressure. J. Chem. Thermodyn. 2012, 49, 81−86. (158) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Extraction of Benzene, Ethylbenzene, and Xylenes from N-Heptane Using Binary Mixtures of [4empy][Tf2N] and [emim][DCA] Ionic Liquids. Fluid Phase Equilib. 2014, 380, 1−10. (159) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Liquid− Liquid Extraction of Toluene from N-Alkanes Using {[4empy][Tf2N] + [emim][DCA]} Ionic Liquid Mixtures. J. Chem. Eng. Data 2014, 59, 1692−1699. (160) Marciniak, A. Influence of Cation and Anion Structure of the Ionic Liquid on Extraction Processes Based on Activity Coefficients at Infinite Dilution. A Review. Fluid Phase Equilib. 2010, 294, 213−233. (161) Wytze Meindersma, G.; Podt, A. J. G.; de Haan, A. B. Selection of Ionic Liquids for the Extraction of Aromatic Hydrocarbons from Aromatic/aliphatic Mixtures. Fuel Process. Technol. 2005, 87, 59−70. (162) Meindersma, G. W.; de Haan, A. B. Conceptual Process Design for Aromatic/aliphatic Separation with Ionic Liquids. Chem. Eng. Res. Des. 2008, 86, 745−752. (163) Jongmans, M. T. G.; Trampé, J.; Schuur, B.; de Haan, A. B. Solute Recovery from Ionic Liquids: A Conceptual Design Study for Recovery of Styrene Monomer from [4-mebupy][BF4]. Chem. Eng. Process. 2013, 70, 148−161. (164) Scurto, A. M.; Aki, S. N. V. K.; Brennecke, J. F. CO2 as a Separation Switch for Ionic Liquid/Organic Mixtures. J. Am. Chem. Soc. 2002, 124, 10276−10277. (165) Scurto, A. M.; Aki, S. N. V. K.; Brennecke, J. F. Carbon Dioxide Induced Separation of Ionic Liquids and Water. Chem. Commun. 2003, 572−573. (166) Aki, S. N. V. K.; Scurto, A. M.; Brennecke, J. F. Ternary Phase Behavior of Ionic Liquid (IL)−Organic−CO2 Systems. Ind. Eng. Chem. Res. 2006, 45, 5574−5585. (167) Mellein, B. R.; Brennecke, J. F. Characterization of the Ability of CO2 To Act as an Antisolvent for Ionic Liquid/Organic Mixtures†. J. Phys. Chem. B 2007, 111, 4837−4843. (168) Canales, R. I.; Brennecke, J. F. Liquid-Liquid Phase Split in Ionic Liquid + Toluene Mixtures Induced by CO2. AIChE J. 2015, 61, 2968−2976. (169) Canales, R. I.; Lubben, M. J.; Gonzalez-Miquel, M.; Brennecke, J. F. Solubility of CO2 in [1-N-butylthiolanium][Tf2N] + Toluene N

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Review

(189) Hayduk, W.; Malik, V. K. Density, Viscosity, and Carbon Dioxide Solubility and Diffusivity in Aqueous Ethylene Glycol Solutions. J. Chem. Eng. Data 1971, 16, 143−146. (190) Azizian, S.; Hemmati, M. Surface Tension of Binary Mixtures of Ethanol + Ethylene Glycol from 20 to 50 °C. J. Chem. Eng. Data 2003, 48, 662−663. (191) Kahl, H.; Wadewitz, T.; Winkelmann, J. Surface Tension of Pure Liquids and Binary Liquid Mixtures. J. Chem. Eng. Data 2003, 48, 580−586. (192) Markarian, S. A.; Terzyan, A. M. Surface Tension and Refractive Index of Dialkylsulfoxide + Water Mixtures at Several Temperatures. J. Chem. Eng. Data 2007, 52, 1704−1709. (193) Srivastava, A. K.; Samant, R. A.; Patankar, S. D. Ionic Conductivity in Binary Solvent Mixtures. 2. Ethylene Carbonate + Water at 25 °C. J. Chem. Eng. Data 1996, 41, 431−435. (194) PubChem. http://pubchem.ncbi.nlm.nih.gov/ (accessed Apr. 13, 2016). (195) Naejus, R.; Damas, C.; Lemordant, D.; Coudert, R.; Willmann, P. Excess Thermodynamic Properties of the Ethylene Carbonate− trifluoroethyl Methyl Carbonate and Propylene Carbonate−trifluoroethyl Methyl Carbonate Systems at T = (298.15 or 315.15) K. J. Chem. Thermodyn. 2002, 34, 795−806. (196) Muhuri, P. K.; Hazra, D. K. Density and Viscosity of Propylene Carbonate + 2-Methoxyethanol at 298.15, 308.15, and 318.15 K. J. Chem. Eng. Data 1995, 40, 582−585. (197) Begum, S. K.; Clarke, R. J.; Ahmed, M. S.; Begum, S.; Saleh, M. A. Densities, Viscosities, and Surface Tensions of the System Water + Diethylene Glycol. J. Chem. Eng. Data 2011, 56, 303−306. (198) EMD Millipore. http://www.emdmillipore.com/US/en/ product/N-Formylmorpholine,MDA_CHEM-818582 (accessed Apr. 14, 2016). (199) Yang, T.; Xia, S.; Song, S.; Fu, X.; Ma, P. Densities and Viscosities of N-Formylmorpholine (NFM) + P-Xylene, + O-Xylene, + M-Xylene at Different Temperatures and Atmospheric Pressure. J. Chem. Eng. Data 2007, 52, 2062−2066. (200) Kelayeh, S. A.; Jalili, A. H.; Ghotbi, C.; Hosseini-Jenab, M.; Taghikhani, V. Densities, Viscosities, and Surface Tensions of Aqueous Mixtures of Sulfolane + Triethanolamine and Sulfolane + Diisopropanolamine. J. Chem. Eng. Data 2011, 56, 4317−4324. (201) Green, D.; Perry, R. Perry’s Chemical Engineers’ Handbook, 8th ed.; McGraw Hill professional; McGraw-Hill Education: 2007. (202) Domańska, U.; Królikowska, M.; Królikowski, M. Phase Behaviour and Physico-Chemical Properties of the Binary Systems {1Ethyl-3-Methylimidazolium Thiocyanate, or 1-Ethyl-3-Methylimidazolium Tosylate + Water, or + an Alcohol}. Fluid Phase Equilib. 2010, 294, 72−83. (203) Quijada-Maldonado, E.; van der Boogaart, S.; Lijbers, J. H.; Meindersma, G. W.; de Haan, A. B. Experimental Densities, Dynamic Viscosities and Surface Tensions of the Ionic Liquids Series 1-Ethyl-3Methylimidazolium Acetate and Dicyanamide and Their Binary and Ternary Mixtures with Water and Ethanol at T = (298.15 to 343.15 K). J. Chem. Thermodyn. 2012, 51, 51−58. (204) Domańska, U.; Królikowska, M.; Walczak, K. Effect of Temperature and Composition on the Density, Viscosity Surface Tension and Excess Quantities of Binary Mixtures of 1-Ethyl-3Methylimidazolium Tricyanomethanide with Thiophene. Colloids Surf., A 2013, 436, 504−511. (205) Tenney, C. M.; Massel, M.; Mayes, J. M.; Sen, M.; Brennecke, J. F.; Maginn, E. J. A Computational and Experimental Study of the Heat Transfer Properties of Nine Different Ionic Liquids. J. Chem. Eng. Data 2014, 59, 391−399. (206) Banerjee, T.; Ramalingam, A. Physiochemical Properties of Aromatic Sulfur and Nitrogen Compounds with Imidazolium-Based Ionic Liquids. In Desulphurization and Denitrification of Diesel Oil Using Ionic Liquids; Elsevier: Amsterdam, 2015; Chapter 5, pp 151−211. (207) Wan Normazlan, W. M. D.; Sairi, N. A.; Alias, Y.; Udaiyappan, A. F.; Jouyban, A.; Khoubnasabjafari, M. Composition and Temperature Dependence of Density, Surface Tension, and Viscosity of EMIM DEP/MMIM DMP + Water + 1-Propanol/2-Propanol Ternary

Mixtures and Their Mathematical Representation Using the Jouyban− Acree Model. J. Chem. Eng. Data 2014, 59, 2337−2348. (208) Makino, T.; Kanakubo, M.; Masuda, Y.; Umecky, T.; Suzuki, A. CO2 Absorption Properties, Densities, Viscosities, and Electrical Conductivities of Ethylimidazolium and 1-Ethyl-3-Methylimidazolium Ionic Liquids. Fluid Phase Equilib. 2014, 362, 300−306. (209) Martino, W.; de la Mora, J. F.; Yoshida, Y.; Saito, G.; Wilkes, J. Surface Tension Measurements of Highly Conducting Ionic Liquids. Green Chem. 2006, 8, 390−397.

O

DOI: 10.1021/acs.jced.6b00077 J. Chem. Eng. Data XXXX, XXX, XXX−XXX