Deep Eutectic Solvents As Azeotrope Breakers LLE

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Research Article

pubs.acs.org/journal/ascecg

Deep Eutectic Solvents as Azeotrope Breakers: Liquid−Liquid


Extraction and COSMO-RS Prediction
Andreia S. L. Gouveia,†,‡ Filipe S. Oliveira,† Kiki A. Kurnia,§ and Isabel M. Marrucho*,†,‡

Instituto de Tecnologia Química e Biológica António Xavier, Universidade Nova de Lisboa, 2780-157 Oeiras, Portugal

Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Avenida Rovisco Pais, 1049-001 Lisboa, Portugal
§
Department of Chemical Engineering, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Perak, Malaysia
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The efficient and sustainable separation of azeotropic


mixtures remains a challenge in chemical engineering. In this work,
the performance of benign solvents, namely deep eutectic solvents
(DES), in the separation of aromatic−aliphatic hydrocarbon
Downloaded via INHA UNIV on September 22, 2022 at 02:19:58 (UTC).

azeotropic mixtures via liquid−liquid extraction (LLE) was evaluated.


The DES studied in this work were based on different ammonium
salts (cholinium chloride, [Ch]Cl, benzylcholinium chloride, [BzCh]-
Cl, and tetrabutylammonium chloride, [N4444]Cl) as hydrogen bond
acceptor (HBA) and one organic acid (levulinic acid, LevA) as
hydrogen bond donor (HBD), always in the mole ratio of 1 HBA:2
HBD. The thermophysical properties, namely density and viscosity,
of the three used DES were measured in the temperature range T =
(293.15 up to 353.15) K and at atmospheric pressure. The phase
equilibria diagrams of all ternary systems were determined at T =
298.15 K and at atmospheric pressure using 1H NMR spectroscopy. The results showed that the introduction of a more
hydrophobic HBA in the DES promotes the improvement of the distribution coefficient, while playing with the aromaticity of the
DES leads to higher selectivity. In addition, the performance of the predictive conductor-like screening model for real solvent
(COSMO-RS) model in the description of these systems was also evaluated. COSMO-RS is capable of quantitatively predicting
the phase behavior and tie-lines for ternary mixtures containing DES as well as of estimating the trend of distribution ratio and
selectivity.
KEYWORDS: Deep eutectic solvents, Aromatic−aliphatic azeotropic mixtures, Liquid−liquid extraction, COSMO-RS

■ INTRODUCTION
In many areas of industry, particularly in the petrochemical
considered toxic, volatile, and/or flammable.7 Consequently,
their replacement by greener alternatives has been a topic of
industry, solvent mixtures accumulate due to recycling interest over the past several years. In this context, ionic liquids
difficulties. Therefore, their efficient separation into pure (ILs), which are known for their exceptional combination of
components is mandatory so that they can be reused. However, properties, such as negligible vapor pressure, low flammability,
many of these solvent mixtures contain azeotropes, whose and high solvation capacity for a wide range of inorganic and
separation is still an issue in chemical engineering. In particular, organic materials, have been proposed as candidates for the
the separation of aromatic hydrocarbons, namely toluene, replacement of conventional organic extractants. Due to their
benzene, xylenes, and others, from mixtures with aliphatic properties, it is expected that the use of ILs in the extraction of
hydrocarbons, still remains an enormous challenge, since these aromatics from aromatic/aliphatic mixtures leads to more
compounds have close boiling points and several combinations sustainable separation with less energy consumption and
of them can form azeotropes.1 smaller number of process steps. A large number of articles
The conventional industrial processes commonly used in the have reported the use of different families of ILs, such as
separation of aromatic and aliphatic hydrocarbon mixtures are imidazolium-, pyridinium-, or ammonium-based ILs. In
azeotropic or extractive distillation and liquid−liquid extraction particular, imidazolium-based ILs combined with several
(LLE). LLE presents economic advantages compared to the anions, such as [OAc]−, [SCN]− , [NTf2 ]−, [HSO 4]− ,
other processes, since it requires low energy demand. However, [CH3SO4]−, [C2H5SO4]−, [CF3SO3]− and others, have been
the conventional extraction solvents typically used in industry
are organic solvents, such as sulfolane,2−4 tetraethylene glycol,5 Received: July 5, 2016
and N-methylpyrrolidone (NMP),6 which require additional Revised: August 25, 2016
investments and energy consumption for their recovery and are Published: August 31, 2016

© 2016 American Chemical Society 5640 DOI: 10.1021/acssuschemeng.6b01542


ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

Table 1. Chemical Structures of the HBA and HBD in the DES Studied in This Work as Well as the Mole Ratio and Respective
Molar Mass, M (g·mol−1)

deeply explored as suitable extraction solvents in the separation in the LLE of BTEX aromatics (benzene, toluene, ethyl-
of different aromatic/aliphatic hydrocarbon mixtures, and some benzene, and xylenes) from n-octane.
ILs showed higher separation factors than those of the DES based on choline chloride and three different HBD,
conventional sulfolane.8−17 Nevertheless, and despite their such as ethylene glycol, levulinic acid, and glycerol, were already
promising and clear advantages, the environmental impact of reported by our group as promising alternatives to ILs in the
ILs has been shown as mostly dependent on the chemical separation of ethanol/n-heptane mixtures, since remarkably
structure of the IL, which means that these fluids are not high selectivities and distribution coefficients were obtained.24
universally green.18 Moreover, the complex ILs synthesis and In this context, in this work we explore the performance of
purification lead to a disadvantageous high cost of production sustainable DES based on three different salts, namely choline
of some ILs. chloride, benzylcholine chloride, and tetrabutylammonium
Deep eutectic solvents (DES) have recently emerged as chloride, and levulinic acid as hydrogen bond donor, in the
potential alternatives to ILs, since their physicochemical separation of toluene from n-heptane via liquid−liquid
properties are similar to those of ILs.19,20 DES are mixtures extraction at T = 298.15 K. The thermophysical properties
of a hydrogen bond acceptor (HBA), such as halide salts and, (density and viscosity) were also measured and discussed, since
typically, a hydrogen bond donor (HBD), viz. carboxylic acids, they influence the mass transfer operations in extraction
alcohols, or amines, that present a much lower melting point processes.
than the pure components. In addition, DES can be synthesized In addition, the capability of the conductor-like screening
from nontoxic, cheap, biodegradable, and biocompatible model for real solvent (COSMO-RS) to predict the phase
materials, using simple methods or preparation, heating and equilibrium of a ternary mixture containing DES was evaluated.
mechanical stirring, or grinding, without the need for further COSMO-RS is a quantum chemical based prediction
purification steps. This fact is probably the main advantage of method25,26 that has been widely used to predict liquid−liquid
DES over ILs. Recently, DES have been tested as extraction phase equilibria of binary and ternary mixtures containing ionic
solvents for the separation of aromatic/aliphatic mixtures. For liquids and hydrocarbon.27,28 However, to our knowledge, this
instance, DES based on tetrabutylphosphonium bromide and work is the first to address the capability of the model to predict
ternary phase diagrams containing DES.


ethyltriphenylphosphonium iodide as HBA and ethylene glycol
or sulfolane as HBD in the separation of toluene from n-
EXPERIMENTAL SECTION
heptane were reported by Kareem et al.1,21 The same group
also tested DES based on methyltriphenylphosphonium Materials. Choline chloride ((2-hydroxyethyl)trimethylammonium
chloride, ≥98 wt %), benzylcholine chloride (benzyl(2-hydroxyethyl)-
bromide as a salt and also ethylene glycol as HBD for the dimethylammonium chloride, ≥97 wt %), [N4444]Cl (tetrabutylam-
separation of benzene from hexane.22 Generally, a similar or monium chloride, ≥97 wt %), and levulinic acid (98 wt %) were
higher performance of DES was observed compared to purchased from Sigma-Aldrich. The chemical structures, respective
conventional solvents or ILs. Moreover, Mulyono et al.23 acronyms, and mole ratio of the DES used in this work are presented
studied DES based on an ammonium salt and sulfolane as HBD in Table 1. The components of the azeotropic mixture, n-heptane and

5641 DOI: 10.1021/acssuschemeng.6b01542


ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

toluene, were purchased from Carlo Erba (99 wt %) and Sigma- literature. In general, the model has been proven to correctly describe
Aldrich (99.9 wt %), respectively, and were used without any further the solvent−solute solvation behavior in a real solvent system
purification. qualitatively and, to some extent, quantitatively.28 At the present
Synthesis and Characterization of DES. In this work, three time, COSMO-RS represents one of the most efficient models to
DES, namely [Ch]Cl:LevA (DES1), [BzCh]Cl:LevA (DES2), and predict the thermodynamic properties of a pure compound, such as an
[N4444]Cl:LevA (DES3), were synthesized in a mole ratio of 1 HBA:2 ionic liquid, and its mixture with organic compounds. In this work, we
HDB using the grinding method,29 in which both components were used this model to predict the LLE behavior between the studied DES
first mixed and then ground in a mortar with a pestle at room and aromatic and aliphatic compounds and to better understand the
temperature in order to obtain a homogeneous liquid. In order to mechanism underlying these separations.
reduce the water and other volatile substance contents, the DES were The first step for the COSMO-RS calculation is to produce the
dried under vacuum (1 Pa) at room temperature for at least 3 days. required .cosmo file for all compounds present in the ternary mixtures.
The water contents of the dried DES were measured by Karl Fischer In this step, the continuum solvation COSMO calculations of
coulometric titration, and all the DES contained less than 0.5 wt % electronic density and molecular geometry were performed using the
H2O. TURBOMOLE 6.1 program package on the density functional theory
The thermophysical properties, namely density, ρ (kg·m−3), and level, utilizing the BP functional B88-P86 with a triple-ζ valence
viscosity, η (Pa·s), of the synthesized DES were measured within the polarized basis set (TZVP) and the resolution of identity standard
temperature range (293.15 and 353.15) K and at atmospheric pressure (RI) approximation.31 After the calculation was finished, the produced
using an SVM 3000 Anton Paar rotational Stabinger viscometer- .cosmo file was then used as input for the next calculation.
densimeter. The standard uncertainty for the temperature is 0.02 K. In the second step, the estimation of the liquid−liquid equilibrium
The repeatability values of density and viscosity of this equipment are of the studied ternary mixtures was performed with the COSMOtherm
0.5 kg·m−3 and 3.5 × 10−6 Pa·s, respectively. Triplicates of each sample program utilizing the parameter file BP_TZVP_C30_1302 (COSMO-
were performed to ensure accuracy, and the reported results are logic GmbH & Co KG, Leverkusen, Germany). The equilibrium for a
average values. Furthermore, the relative standard uncertainty in ternary liquid−liquid system is defined by the following equation:
density and viscosity was calculated by the ratio of the standard
deviation and the average of the three replicates of each sample, where xiIγi I = xiIIγi II (1)
the highest relative standard uncertainty registered for the density and
dynamic viscosity measurements was 0.2 and 2 × 10−5, respectively. where γi is the activity coefficient of component i in phase I or II, and
LLE Measurements. The LLE consisted of two steps: in the first xi is the mole fraction of component i in phase I or II. Both γi and xi
step, preparation of calibration curves to determine the concentration are solely predicted using COSMO-RS. It should be highlighted that,
of toluene in n-heptane or the DES phase was carried out, while, in the during all calculations, the DES is treated as a complex neutral
second step, the tie-lines for each one of the ternary mixtures were molecule.
determined.
The calibration curves were constructed using a binary mixture of n-
heptane + toluene and toluene + DES at different mole fractions,
depending on the solubility of toluene in the respective solvent. All the
■ RESULTS AND DISCUSSION
Density and Viscosity Measurements. The thermophys-
binary mixtures were stirred using a magnetic stirrer for at least 1 h. ical properties, namely density and viscosity, are significant
After the equilibrium, each sample was analyzed using 1H NMR parameters that influence the selection of proper extraction
spectroscopy (Bruker Avance II+ 500 MHz NMR spectrometer) in solvent in LLE, since, for instance, high viscosities lead to mass
order to obtain the calibration curves, integration of the area under the transfer limitations and also operational costs related to liquid
peak vs mole fraction. For the NMR analysis, the samples were placed pumping and dispersion problems. Bearing this in mind, the
in NMR spectroscopy sealed tubes with dimethyl sulfoxide-d6 thermophysical characterization of the studied DES was
(DMSO-d6) as solvent. The deuterated solvent was not mixed with performed in this work, except for DES1, whose thermophys-
the sample but kept inside a sealed capillary tube, which was then ical properties were previously reported by our group.32 The
introduced inside the NMR tube, minimizing the amount of solvent
required. The DMSO-d6 was used for the external lock of the NMR density, ρ (kg·m−3), and viscosity, η (Pa·s), values of the three
magnetic field, and the greatest quantitative agreement was found DES studied in this work are given in Table S1 in the
when selecting, for [C2mim][OAc], the peaks corresponding to the Supporting Information. Since the studied DES are composed
imidazolium cation (for instance, the methyl group bonded to the by the same HBD (levulinic acid), it allows us to study the
nitrogen atom of the imidazolium ring), and for the studied DES, the impact of HBA chemical structures toward the density and
peaks corresponding to their HBA (for DES1 and DES2, the peaks of viscosity of the obtained DES. As can be seen from Figure 1a,
the methyl groups bonded to the nitrogen atom and, for DES3, the DES2 has the highest density, closely followed by DES1, while
peaks of the terminal methyl groups). DES3 has the lowest density. The observed difference in the
After the calibration curves were established, the next step was density of the DES might be attributed to the different HBA.
determination of tie-lines for each ternary mixture. In this step, ternary
mixtures of known composition were prepared in vials and were
DES1 and DES2 are formed by HBAs that contain one
vigorously stirred by magnetic agitation for 1 h and left to reach hydroxyl group that enables the further establishment of
equilibrium for at least 24 h in an incubator at 25 °C to ensure the hydrogen bonds with HBD, leading to more close arrangement
complete phase separation. This period of time was checked to be between HBA and HBD and, ultimately, to higher density. On
sufficient for reaching the equilibrium. Then, each phase was carefully the other hand, the lack of a hydroxyl group of the HBA in
sampled to avoid cross-contamination of the phases, and both phases DES3 is responsible for its lower density when compared to the
were analyzed by 1H NMR spectroscopy using the capillary method other two DES. Similar behavior was also observed previously
previously described to establish the calibration curves. The by Kroon et al.33 for DES based on choline chloride and
composition of each phase was calculated by the previously obtained tetramethylammonium chloride as HBAs. Thus, the presence of
calibration curves.
the hydroxyl group in the HBA forming the DES might have a
COSMO-RS Prediction. The conductor-like screening model for
real solvent developed by Klamt and co-worker30 is a predictive model significant impact on its density.
that integrates the concepts of quantum chemistry, dielectric The density of the studied DES decreases with increasing
continuum models, electrostatic interactions, and statistical thermody- temperaturea general trend observed for organic compounds
namics. Numerous theoretical documents and evaluation of the and ILs. The density decrement shows a linear behavior with
performance of many COSMO-RS applications can be found in the the temperature increment, and thus the density values were
5642 DOI: 10.1021/acssuschemeng.6b01542
ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

Table 2. Fitted Parameters of the Linear Expression Given


by Eq 2, Relative Standard Deviation, σ, and Respective
Correlation Coefficient, R2a
a (g·cm−3) b × 104 (g·cm−3−1)K σ R2
−4
DES 1 1.3324 −6.53 2.44 × 10 0.9994
DES 2 1.3472 −6.58 1.47 × 10−4 0.9998
DES 3 1.2034 −6.33 1.15 × 10−4 0.9999
a
Standard uncertainties u are u(a) = 0.002 and u(b) = 5 × 10−6.

where ρ corresponds to the density (g·cm−3) and M is the


molar mass (g·mol−1) of the DES. The calculated molar volume
values are given in Table S1 in the Supporting Information and
are depicted in Figure 1b. In general, the molar volumes of all
DES slightly increase with increasing temperature. Within the
studied temperature range, the molar volume can be ranked as
DES1 < DES2 < DES3. This rank of molar volumes of the
studied DES is similar to the rank of their molar mass, where
DES1 (123.95 g·mol−1) < DES2 (149.32 g·mol−1) < DES3
(170.05 g·mol−1). Thus, it is interesting to note that the
presence of different HBAs plays a minimal role in the molar
volume, since the differences between the molar volumes of the
studied DES are similar to those observed for the molar mass.
Regarding the viscosity of the studied DES, from Figure 1c, it
can be seen that DES2 presents the highest viscosity, followed
by DES3, and DES1 has the lowest viscosity values. The highest
viscosity of DES2 in comparison to DES1 might be attributed
to the presence of the aromatic ring in its structure that, besides
largely increasing the DES molar volume, might promote
additional π−π interactions. Unlike density, it seems that the
Figure 1. Experimental (a) density, (b) molar volume, and (c) presence of a hydroxyl group in the HBA does not have a
viscosity values for (diamonds) [Ch]Cl:LevA (1:2); (circles) [BzCh]- significant impact on viscosity, since the viscosity of DES3 lies
Cl:LevA (1:2); and (triangles) [N4444]Cl:LevA (1:2) as a function of in between that of DES1 and DES2. These intermediate
temperature (T). The density and viscosity values of [Ch]Cl:LevA
(1:2) were taken from Florindo et al.29
viscosity values might be attributed to the longer alkyl chains of
ammonium cation in the HBA of DES3, since and as already
observed in the literature,34 an increase in the alkyl chain length
fitted as a linear function of temperature, T (K), by the least- of the quaternary ammonium cation increases the area of
squares method using the linear expression given by eq 2: contact between the molecules, which also leads to an increase
of the dispersive forces between the molecules and,
ρ = a + b(T ) (2) consequently, high viscosities.
where a and b are adjustable parameters and ρ is density. The Similar to density, the viscosity of the studied DES decreases
reliability of eq 1 was evaluated based on the correlation with increasing temperature. A very significant decrement of
coefficient, R2, and the relative standard deviation, σ, which was viscosity is observed for DES2, from 2.026 Pa·s (at 293.15 K)
calculated using the following equation: to 0.04 Pa·s (at 353.15 K) in the studied temperature range.
The experimental viscosity values were fitted as a function of
N ⎛ ρ − ρ ⎞2 temperature, using the Vogel−Fulcher−Tammann (VFT)
1
∑ ⎜⎜ ⎟
exp, i calc, i
σ= ⎟ model described in eq 5:
N i=1 ⎝
ρcalc, i ⎠ (3)

where ρexp,i is the experimental data, ρcalc,i is calculated from the ln η = A η +
calculation, and N is the total number of experimental points. (T − Cη) (5)
The a and b parameters, along with the R2 and σ values, are
given in Table 2. From the correlation coefficient, R2, listed in where η is the viscosity in mPa·s, T is the temperature in K, and
Table 2, and also the standard deviation values, it can be Aη, Bη, and Cη are adjustable parameters. The adjustable
concluded that the use of a linear function adequately describes parameters, which were determined from the fitting of the
the measured density data for the studied temperature range. experimental data, are listed along with the relative standard
From the temperature dependence of density, the molar deviation, σ, and the energy barrier of a fluid to shear stress, E
volumes (Vm) of the studied DES were calculated using eq 4: (kJ·mol−1) at T = 298.15 K, in Table 3. The σ values were
calculated using eq 3, and the energy barrier was determined
M
Vm = based on the viscosity dependence with temperature using eq 6,
ρ (4) as follows:35
5643 DOI: 10.1021/acssuschemeng.6b01542
ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

Table 3. Fitted Parameters of the VFT Expression Given by Eq 5, Relative Standard Deviation, σ, Respective Correlation
Coefficient, R2, and Energy Barrier Values at T = 298.15 K
Parameters
Aη (mPa·s) Bη (K) Cη (K) σ R2 E298.15K (kJ·mol‑1)
−4
DES 1 −2.198 869.6 184.0 3.53 × 10 1.0 40.8
DES 2 −2.767 1020.5 194.9 8.07 × 10−4 1.0 55.8
DES 3 −2.915 1000.4 185.9 9.44 × 10−4 1.0 47.5

Figure 2. Literature16 (black dots) and this work (blue dots) tie-lines of the n-heptane (1) + toluene (2) + [C2mim][OAc] (3) ternary system at
298.15 K.

⎛ ⎞ the literature and experimental (this work) tie-lines of this


∂(ln η) ⎜ B ⎟ ternary system at T = 298.15 K. As can be seen from Figure 2,
E=R = R ⎜ η
⎟ the tie-lines determined by 1H NMR spectroscopy are in good
∂ T
1
() ⎜ ⎛⎜ Cη2 2C ⎞⎟
⎜ 2 − T η + 1⎟ ⎟ agreement with those reported in the literature determined by
⎝ ⎝T ⎠⎠ (6) the conventional method at the same temperature (rmsd =
where η is the viscosity, T is the temperature, Bη and Cη are the 0.9%), which allow us to validate 1H NMR spectroscopy as
adjustable parameters obtained from eq 5, and R is the universal quantification method. In this context, the performance of
gas constant. DES1, DES2, and DES3 in the specific separation of toluene
The higher the energy barrier value, the more difficult it is for from n-heptane was studied using the validated methodology.
the DES molecules to flow. This fact can be a direct For all the studied ternary systems, the composition of each
consequence of the size or entanglement of the molecules one of the DES in the n-heptane-rich phases could be
and/or of the presence of stronger interactions, for instance, H- considered to be negligible, since 1H NMR spectra showed
bonding interactions and π−π interactions, in the fluid. From no detectable signals corresponding to DES. The 1H NMR
Table 3, it can be observed that DES2, containing [BzCh]Cl as spectra of typical upper (heptane-rich phase) and bottom
HBA, presents a higher E value, due to the presence of the phases (DES-rich phase) of the studied ternary systems are
aromatic ring in its structure, than DES1, which contains provided in the Supporting Information (Figures S1−S6).
[Ch]Cl as HBA. The E values of the three DES studied can be From these spectra, it can also be observed that the proportion
ordered as follows: DES1 < DES3 < DES2, which are in of all the DES is maintained for the three ternary systems and
agreement with the experimental viscosity values obtained. also for all the studied compositions. This leads to the
LLE Measurements. The use of 1H NMR spectroscopy as conclusion that, contrary to what happens in aqueous
quantification method for other ternary systems containing systems,38,39 dissociation of the DES in these organic solvent
ionic liquids was already tested and reported in the mixtures is not observed, and thus, DES are suitable solvents to
literature.8,9,36,37 To further validate the NMR method used carry out the proposed application.
in this work, the tie-lines of the ternary system n-heptane + The experimental liquid−liquid phase equilibria data of the
toluene + [C2mim][OAc] measured in this work were three ternary systems were measured at T = 298.15 K and
compared to literature data that was obtained using a atmospheric pressure and are presented in Table 4 and plotted
conventional method.16 Figure 2 shows a comparison between through a ternary diagram in Figures 3, 4, and 5. It should be
5644 DOI: 10.1021/acssuschemeng.6b01542
ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

Table 4. Composition of the Experimental Tie-Lines, contamination in DES3 is very high, thus requiring a difficult
Toluene Distribution Coefficient (β2), and Selectivity (S) for solvent recovery stage. The recovery of entrainers is usually
the Ternary Systems at T = 298.15 Ka carried out using distillation. Therefore, thermal studies will be
required before the introduction of DES in a refinery.
Heptane-rich phase DES-rich phase
It can also be observed from Figures 3−5 that the tie-lines
x1 x2 x1 x2 β2 S obtained for all the studied ternary systems exhibit negative
Heptane (1) + Toluene (2) + DES1 (3) slopes, leading to distribution coefficients lower than unity;
0.9207 0.0793 0.0038 0.0079 0.100 23.90 consequently, a high amount of DES will be required for this
0.8205 0.1795 0.0089 0.0184 0.103 9.47 extraction. Despite the fact that DES are cheap and easy to
0.7380 0.2620 0.0056 0.0277 0.106 13.93 synthesize, this drawback can be overcome by recovering and
0.6765 0.3235 0.0057 0.0367 0.113 13.46 reusing the extraction solvent.
0.6263 0.3737 0.0046 0.0474 0.127 17.27 Distribution coefficient, β, and selectivity, S, are the two
0.5504 0.4496 0.0059 0.0553 0.123 11.47 crucial parameters broadly used in the evaluation of extraction
0.4968 0.5032 0.0050 0.0608 0.121 12.01 solvent suitability to perform LLE. Thus, these two parameters
0.4392 0.5608 0.0023 0.0640 0.114 21.79 were calculated from the experimental data as follows:
0.3632 0.6368 0.0020 0.0672 0.106 19.16
0.3090 0.6910 0.0020 0.0713 0.103 16.18 x 2downphase
β2 =
0.2609 0.7391 0.0014 0.0742 0.100 18.97 x 2upperphase (7)
0.2125 0.7875 0.0011 0.0746 0.095 18.30
Heptane (1) + Toluene (2) + DES2 (3) ⎛ x upperphase ⎞⎛ x downphase ⎞
0.9314 0.0686 0.0080 0.0231 0.337 39.11 S = ⎜⎜ 1downphase ⎟⎟⎜⎜ 2upperphase ⎟⎟
0.8358 0.1642 0.0092 0.0324 0.198 17.90 ⎝ x1 ⎠⎝ x 2 ⎠ (8)
0.7503 0.2497 0.0102 0.0418 0.167 12.31 where xupper phase
and xupper phase
are the mole fractions of heptane
1 2
0.6844 0.3156 0.0063 0.0500 0.158 17.21 and toluene in the upper phase (heptane-rich phase),
0.6199 0.3801 0.0058 0.0625 0.164 17.57 respectively, and xdown phase
and xdown phase
are the mole fractions
1 2
0.5637 0.4363 0.0057 0.0745 0.171 16.92 of heptane and toluene in the lower phase (DES-rich phase).
0.5075 0.4925 0.0055 0.0875 0.178 16.39 The distribution coefficient and selectivity values are also listed
0.4520 0.5480 0.0048 0.1036 0.189 17.80 in Table 4 and represented as a function of the toluene mole
0.3703 0.6297 0.0038 0.1171 0.186 18.12
fraction in the aliphatic-rich phase in Figures 6 and 7,
0.3127 0.6873 0.0035 0.1211 0.176 15.74
respectively.
Heptane (1) + Toluene (2) + DES3 (3)
Taking into account that the distribution coefficient
0.9455 0.0545 0.0524 0.0405 0.742 13.41
determines the quantity of DES required for the extraction
0.8634 0.1366 0.0506 0.0671 0.491 8.38
process and that selectivity is related to the ability of DES to
0.7792 0.2208 0.0533 0.1060 0.480 7.02
extract the toluene in preference to n-heptane, the ideal
0.7276 0.2724 0.0514 0.1354 0.497 7.04
extraction solvent should have both high distribution coefficient
0.6693 0.3307 0.0557 0.1588 0.480 5.77
and selectivity values, since high distribution coefficients
0.5963 0.4037 0.0544 0.1960 0.486 5.32
correspond to lower operating costs and high selectivity values
0.5483 0.4517 0.0598 0.2211 0.489 4.49
lead to fewer number of extraction stages required in the
0.4813 0.5187 0.0642 0.2554 0.492 3.69
process.
0.4038 0.5962 0.0647 0.3055 0.512 3.20
From Figures 6 and 7, it can be seen that DES3 presents
0.3390 0.6610 0.0650 0.3374 0.510 2.66
lower selectivity but higher distribution coefficient values
0.2819 0.7181 0.0681 0.3816 0.531 2.20
compared to DES1 and DES2. The selectivity values can be
0.2302 0.7698 0.0707 0.4246 0.552 1.80
ordered as follows: DES3 < DES1 < DES2, which are in
0.1439 0.8561 0.0706 0.4944 0.578 1.18
agreement with the trend observed for the experimental density
0.0778 0.9222 0.0622 0.5884 0.638 0.80
values. In addition, the density difference between both phases
0.0447 0.9553 0.0481 0.6141 0.643 0.60
a
is very important, since large density differences permit higher
Standard uncertainties u are u(T) = 0.02 K, u(x) = 0.0001. equipment capacities. Although in the present work we did not
measure the density of the phases in equilibrium but only of the
highlighted that the compositions of both phases in equilibrium pure DES, the composition of the DES rich phase is very high
were calculated using calibration curves established in this in DES (usually around 0.85−0.95). Thus, the density
work. difference between the DES-rich phase and the phase rich in
From Figures 3 and 4, it can be seen that DES1 and DES2 toluene and n-heptane is most probably sufficiently large, since
are nearly immiscible to n-heptane. Also, although the solubility the DES density values, at 298.15 K, range from 1015 kg·m−3 to
of these two DES in toluene is very similar, DES2 presents a 1151 kg·m−3 for the three systems studied while the density
slightly higher solubility value, due to the presence of the values of the phase rich in toluene and n-heptane are in
aromatic ring in the HBA. Moreover, a different behavior was between those of the pure compounds, at 298.15 K (ρn‑heptane =
found when the HBA was changed to a quaternary ammonium 679 kg·m−3 and ρtoluene = 872 kg·m−3). Moreover, the presence
salt with higher hydrophobicity, [N4444]Cl. DES3 containing of the aromatic ring in the HBA of DES2 leads to high
this salt as HBA presents higher solubility in toluene compared distribution coefficient and selectivity values when compared to
to DES1 and DES2, as can be seen in Figure 5. The very low DES1, containing [Ch]Cl as HBA. In this case only the
solubility of the studied DES1 and DES2 in the heptane-rich aromaticity is under discussion because the two DES were
phase minimizes the loss of the DES and, consequently, the prepared at the same mole ratio and also have the same HBD.
contamination of the refined stream. However, the cross- Moreover, the introduction of hydrophobicity in the DES,
5645 DOI: 10.1021/acssuschemeng.6b01542
ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

Figure 3. Tie-lines for a ternary mixture of {heptane + toluene + DES1} at T = 298.15 K and p = 0.1 MPa. Symbols: (circles, ), experimental data;
(squares, - - -), COSMO-RS.

Figure 4. Tie-lines for a ternary mixture of {heptane + toluene + DES 2} at T = 298.15 K and p = 0.1 MPa. Symbols: (circles, ), experimental
data; (squares, - - -), COSMO-RS.

selecting a HBA based on a quaternary ammonium salt, caused and b were obtained at a different temperature (T = 313.15 K),
an enhancement of the distribution coefficient values. it can be observed that the distribution coefficients and
The experimental selectivity and distribution coefficient
selectivities of the systems studied in this work are of the same
values were compared to the literature data and are plotted
in Figures 8a and b, respectively. Although a direct comparison order of magnitude of the literature values. Nevertheless, the
cannot be exactly made, since the distribution coefficients and selectivity and distribution coefficients of the DES tested in this
selectivities of DES and most of the ILs presented in Figures 8a work are higher than the literature values for DES.
5646 DOI: 10.1021/acssuschemeng.6b01542
ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

Figure 5. Tie-lines for a ternary mixture of {heptane + toluene + DES3} at T = 298.15 K and p = 0.1 MPa. Symbols: (circles, ), experimental data;
(squares, - - -), COSMO-RS.

Figure 6. Distribution coefficient (β2) values as a function of the mole Figure 7. Selectivity (S) values as a function of the mole fraction of
fraction of toluene in the n-heptane-rich phase at 298.15 K for the toluene in the n-heptane-rich phase at 298.15 K for the ternary
ternary systems: (diamonds) n-heptane + toluene + DES1); (circles) systems: (diamonds) n-heptane + toluene + DES1; (circles) n-heptane
n-heptane + toluene + DES2); and (triangles) n-heptane + toluene + + toluene + DES2; and (triangles) n-heptane + toluene + DES3. The
DES3). The closed and open symbols represent experimental and closed and open symbols represent experimental and COSMO-RS
COSMO-RS prediction, respectively. prediction, respectively.

COSMO-RS Prediction. The LLE data of the three systems


studied in this work was predicted using COSMO-RS the ternary mixtures at both phases predicted using COSMO-
calculations. The COSMO-RS model combines a number of RS is given in Table S2 in the Supporting Information. In order
quantum chemical calculations with statistical thermodynamics to evaluate the performance of COSMO-RS to predict the
in order to determine and predict the thermophysical liquid−liquid phase equilibria of the studied ternary systems,
properties and phase behavior of fluids.40 Two different the root-mean-square deviation (rmsd) between the exper-
approaches can be considered for DES using COSMO-RS imental and predicted data was determined using the following
calculations: the independent-ion model ([C] + [A]), in which equation:
the DES is treated as an equimolar mixture, or the ion-paired
I
model ([CA]), in which the DES is considered as a whole rmsd = 100[∑ ∑ (xiexp,
,n − xipred
,n
,I 2
)
neutral molecule. In this work, the COSMO-RS calculations i n
were performed assuming that DES are a whole neutral II 1/2
molecule. The concentration of each one of the compounds in + (xiexp,
,n − xipred
,n
, II 2
) /2NR ] (9)

5647 DOI: 10.1021/acssuschemeng.6b01542


ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

The change in the distribution ratio and selectivity with the


concentration of toluene in the heptane-rich phase was also
estimated with COSMO-RS using eqs 7 and 8, respectively.
The estimated distribution ratio and selectivity predicted using
COSMO-RS are plotted in Figures 6 and 7, respectively,
together with the values obtained from experimental data.
COSMO-RS was able to capture the low and inferior to 1
distribution ratio values, as obtained from experimental data. As
mentioned before, these values resulted from negative tie-line
slopes (cf. Figures 3−5), meaning that large volumes of DES
are required for the extraction process.
Regarding selectivity, COSMO-RS was also able to produce a
similar trend of selectivity values to that obtained from
experimental data (provided in Table S2 in the Supporting
Information). In general, the selectivity values estimated using
COSMO-RS are superior to unity, thus confirming the
potential use of the studied DES as solvent from the extraction
process when applied to aromatic−aliphatic hydrocarbon
separation. Furthermore, the model has correctly described
the higher selectivity of DES1 and DES2 toward the aromatic−
aliphatic separation than of DES3, as observed experimentally.
It seems that the presence of a polar group, namely hydroxyl, is
responsible for an increase in the selectivity of the studied DES
toward aromatic−aliphatic separation.

■ CONCLUSIONS
In this work, DES based on three different salts as hydrogen
bond acceptors and a common hydrogen bond donor were
Figure 8. (a) Selectivity (S) and (b) distribution coefficient (β2) values synthesized and their thermophysical properties, namely
of the systems studied in this work (green bars) and systems reported density and viscosity, were analyzed, since these two parameters
in the literature (gray bars),13,14,16,17,21 at T = 298.15 K (for DES1, influence the selection of a suitable extraction solvent in
DES2, DES3, and [C2mim][OAc]) and T = 313.15 K (for the remain liquid−liquid extraction processes. It was observed that the
DES and ILs) and toluene mole fraction in the upper phase ≈0.1. DES containing a −OH group in the HBA present higher
densities compared to quaternary ammonium-based DES.
where x is the mole fraction of compound i, R is the total Afterward, their performance as extraction solvents in the
number of compounds (R = 3), n is the tie-line number, and N separation of toluene from n-heptane via liquid−liquid
is the total number of experiments. The rmsd values for the extraction was evaluated, and it was concluded that the
studied systems are also shown in Table S2 in the Supporting presence of an aromatic ring in the HBA of DES leads to
Information. high distribution coefficients and selectivities. However, the
The results for COSMO-RS prediction for the studied distribution coefficient values can be enhanced with the
ternary systems are depicted in Figures 3−5 along with the introduction of a more hydrophobic HBA, for instance, a
experimental data. COSMO-RS provides an excellent quanti- quaternary ammonium salt. Also, it was shown that the DES3
tative prediction of the shape and size of the miscibility gap and presents higher solubility in toluene compared to the other two
of the binodal curves and tie-lines for the ternary systems studied DES. A balance between the selectivity and the
composed by DES1 and DES2. The differences between the coefficient distribution is always needed in order to select the
predicted values and the experimental data are more significant ideal extraction solvent for a specific extraction process. For the
for the ternary systems with DES3 (rmsd = 3.9%), as the studied ternary systems, COSMO-RS can satisfactorily describe
immiscibility region decreases when compared with the systems the trend observed for the phase diagrams and tie-lines slopes,
containing DES1 (rmsd = 0.9%) and DES2 (rmsd = 0.6%). The with rmsd varied between 0.5 and 3.9%. In addition, COSMO-
obtained phase diagrams, using both experimental data and RS was also able to capture the trend of distribution ratio and
COSMO-RS, are of type 2, presenting a partial miscibility in selectivity as observed experimentally. The prediction results
two pairs of compoundsthat is DES + heptane and DES + here reported show that COSMO-RS model can be successfully
toluene. For the ternary phase containing DES1 and DES2, the applied to the a priori screening of DES to be used as solvent in
predicted binodal curves at the heptane-rich phase (upper separations of aromatic−aliphatic hydrocarbon mixtures.


phase) lie close to the edge of the phase diagram due to low
solubilities displayed by the studied DES, as also evident from ASSOCIATED CONTENT
the NMR method used in this work. On the other hand, the
phase diagram containing DES3 shows a noticeable miscibility
* Supporting Information
S

gap domain when compared to the other two DES. This might The Supporting Information is available free of charge on the
be addressed due to its low polarity, which results from the ACS Publications website at DOI: 10.1021/acssusche-
large alkyl chains and the absence of a hydroxyl group in the meng.6b01542.
HBA of DES3, and a subsequent increase of the dispersion Experimental density, viscosity, and calculated molar
forces, thus enhancing the miscibility of DES3 and n-heptane. volume of the synthesized DES. Typical 1H NMR spectra
5648 DOI: 10.1021/acssuschemeng.6b01542
ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

of both upper and bottom phases of the ternary system: (11) Maduro, R. M.; Aznar, M. Liquid−liquid equilibrium of ternary
n-heptane + toluene + DES. Composition of the tie-lines, systems 1-octyl-3-methylimidazolium hexafluorophosphate + aromatic
toluene distribution coefficient (β2) and selectivity (S), + aliphatic hydrocarbons. Fluid Phase Equilib. 2010, 296 (2), 88−94.
(12) Rodríguez, H.; Francisco, M.; Soto, A.; Arce, A. Liquid−liquid
predicted using COSMO-RS, as well as the root-mean- equilibrium and interfacial tension of the ternary system heptane +
square deviation (rmsd) between the experimental and thiophene + 1-ethyl-3-methylimidazolium bis-
predicted data (PDF) (trifluoromethanesulfonyl)imide. Fluid Phase Equilib. 2010, 298 (2),


240−245.
(13) García, S.; García, J.; Larriba, M.; Torrecilla, J. S.; Rodríguez, F.
AUTHOR INFORMATION Sulfonate-Based Ionic Liquids in the Liquid−Liquid Extraction of
Corresponding Author Aromatic Hydrocarbons. J. Chem. Eng. Data 2011, 56 (7), 3188−3193.
*Tel: +351-21-4469724. Fax: +351-21-4411277. E-mail ad- (14) García, S.; Larriba, M.; García, J.; Torrecilla, J. S.; Rodríguez, F.
dress: imarrucho@itqb.unl.pt. Liquid−Liquid Extraction of Toluene from Heptane Using 1-Alkyl-3-
methylimidazolium Bis(trifluoromethylsulfonyl)imide Ionic Liquids. J.
Notes Chem. Eng. Data 2011, 56 (1), 113−118.
The authors declare no competing financial interest. (15) Varma, N. R.; Ramalingam, A.; Banerjee, T. Experiments,


correlations and COSMO-RS predictions for the extraction of
benzothiophene from n-hexane using imidazolium-based ionic liquids.
ACKNOWLEDGMENTS
Chem. Eng. J. 2011, 166 (1), 30−39.
I.M.M. acknowledges FCT/MCTES (Portugal) for a contract (16) Corderí, S.; Gómez, E.; Calvar, N.; Domínguez, Á . Measure-
under Investigador FCT 2012. This work was partially ment and Correlation of Liquid−Liquid Equilibria for Ternary and
supported by FCT through the project PTDC/QEQ-FTT/ Quaternary Systems of Heptane, Cyclohexane, Toluene, and [EMim]-
1686/2012 and R&D unit, UID/Multi/04551/2013 (Green- [OAc] at 298.15 K. Ind. Eng. Chem. Res. 2014, 53 (22), 9471−9477.
IT). The NMR spectrometers are part of the National NMR (17) Larriba, M.; Navarro, P.; García, J.; Rodríguez, F. Selective
Facility supported by Fundação para a Ciência e a Tecnologia extraction of toluene from n-heptane using [emim][SCN] and
[bmim][SCN] ionic liquids as solvents. J. Chem. Thermodyn. 2014,
(RECI/BBB-BQB/0230/2012).


79, 266−271.
(18) Petkovic, M.; Seddon, K. R.; Rebelo, L. P. N.; Silva Pereira, C.
REFERENCES Ionic liquids: a pathway to environmental acceptability. Chem. Soc. Rev.
(1) Kareem, M. A.; Mjalli, F. S.; Hashim, M. A.; Hadj-Kali, M. K. O.; 2011, 40 (3), 1383−1403.
Bagh, F. S. G.; Alnashef, I. M. Phase equilibria of toluene/heptane with (19) Abbott, A. P.; Boothby, D.; Capper, G.; Davies, D. L.; Rasheed,
tetrabutylphosphonium bromide based deep eutectic solvents for the R. K. Deep Eutectic Solvents Formed between Choline Chloride and
potential use in the separation of aromatics from naphtha. Fluid Phase Carboxylic Acids: Versatile Alternatives to Ionic Liquids. J. Am. Chem.
Equilib. 2012, 333, 47−54. Soc. 2004, 126 (29), 9142−9147.
(2) Choi, Y.-J.; Kwon, T.-I.; Yeo, Y.-K., Optimization of the sulfolane (20) Ruβ, C.; Konig, B. Low melting mixtures in organic synthesis -
extraction plant based on modeling and simulation. Korean J. Chem. an alternative to ionic liquids? Green Chem. 2012, 14 (11), 2969−
Eng. 2000, 17 (6), 712−718.10.1007/BF02699122 2982.
(3) Chen, J.; Duan, L.-P.; Mi, J.-G.; Fei, W.-Y.; Li, Z.-C. Liquid− (21) Kareem, M. A.; Mjalli, F. S.; Hashim, M. A.; Hadj-Kali, M. K. O.;
liquid equilibria of multi-component systems including n-hexane, n- Ghareh Bagh, F. S.; Alnashef, I. M. Phase equilibria of toluene/heptane
octane, benzene, toluene, xylene and sulfolane at 298.15 K and with deep eutectic solvents based on ethyltriphenylphosphonium
atmospheric pressure. Fluid Phase Equilib. 2000, 173 (1), 109−119. iodide for the potential use in the separation of aromatics from
(4) Chen, J.; Li, Z.; Duan, L. Liquid−Liquid Equilibria of Ternary naphtha. J. Chem. Thermodyn. 2013, 65, 138−149.
and Quaternary Systems Including Cyclohexane, 1-Heptene, Benzene, (22) Kareem, M. A.; Mjalli, F. S.; Hashim, M. A.; AlNashef, I. M.
Toluene, and Sulfolane at 298.15 K. J. Chem. Eng. Data 2000, 45 (4), Liquid−liquid equilibria for the ternary system (phosphonium based
689−692. deep eutectic solvent−benzene−hexane) at different temperatures: A
(5) Wang, W.; Gou, Z.; Zhu, S. Liquid−Liquid Equilibria for new solvent introduced. Fluid Phase Equilib. 2012, 314, 52−59.
Aromatics Extraction Systems with Tetraethylene Glycol. J. Chem. Eng. (23) Mulyono, S.; Hizaddin, H. F.; Alnashef, I. M.; Hashim, M. A.;
Data 1998, 43 (1), 81−83. Fakeeha, A. H.; Hadj-Kali, M. K. Separation of BTEX aromatics from
(6) Al-Jimaz, A. S.; Fandary, M. S.; Alkhaldi, K. H. A. E.; Al-Kandary, n-octane using a (tetrabutylammonium bromide + sulfolane) deep
J. A.; Fahim, M. A. Extraction of Aromatics from Middle Distillate eutectic solvent - experiments and COSMO-RS prediction. RSC Adv.
Using N-Methyl-2-pyrrolidone: Experiment, Modeling, and Optimi- 2014, 4 (34), 17597−17606.
zation. Ind. Eng. Chem. Res. 2007, 46 (17), 5686−5696. (24) Oliveira, F. S.; Pereiro, A. B.; Rebelo, L. P. N.; Marrucho, I. M.
(7) Wytze Meindersma, G.; Podt, A.; de Haan, A. B. Selection of Deep eutectic solvents as extraction media for azeotropic mixtures.
ionic liquids for the extraction of aromatic hydrocarbons from Green Chem. 2013, 15 (5), 1326−1330.
aromatic/aliphatic mixtures. Fuel Process. Technol. 2005, 87 (1), 59− (25) Klamt, A.; Eckert, F. COSMO-RS: a novel and efficient method
70. for the a priori prediction of thermophysical data of liquids. Fluid Phase
(8) Arce, A.; Earle, M. J.; Rodriguez, H.; Seddon, K. R. Separation of Equilib. 2000, 172 (1), 43−72.
aromatic hydrocarbons from alkanes using the ionic liquid 1-ethyl-3- (26) Klamt, A. COSMO-RS from quantum chemistry to fluid phase
methylimidazoliumbis{(trifluoromethyl) sulfonyl}amide. Green Chem. thermodynamics and drug design; Elsevier: 2005.
2007, 9 (1), 70−74. (27) Ferreira, A. R.; Freire, M. G.; Ribeiro, J. C.; Lopes, F. M.;
(9) Arce, A.; Earle, M. J.; Rodríguez, H.; Seddon, K. R. Separation of Crespo, J. G.; Coutinho, J. A. P. An Overview of the Liquid−Liquid
Benzene and Hexane by Solvent Extraction with 1-Alkyl-3- Equilibria of (Ionic Liquid + Hydrocarbon) Binary Systems and Their
methylimidazolium Bis{(trifluoromethyl)sulfonyl}amide Ionic Modeling by the Conductor-like Screening Model for Real Solvents.
Liquids: Effect of the Alkyl-Substituent Length. J. Phys. Chem. B Ind. Eng. Chem. Res. 2011, 50 (9), 5279−5294.
2007, 111 (18), 4732−4736. (28) Ferreira, A. R.; Freire, M. G.; Ribeiro, J. C.; Lopes, F. M.;
(10) Arce, A.; Earle, M. J.; Rodríguez, H.; Seddon, K. R.; Soto, A. Crespo, J. G.; Coutinho, J. A. P. Overview of the Liquid−Liquid
Isomer effect in the separation of octane and xylenes using the ionic Equilibria of Ternary Systems Composed of Ionic Liquid and
liquid 1-ethyl-3-methylimidazolium bis{(trifluoromethyl)sulfonyl}- Aromatic and Aliphatic Hydrocarbons, and Their Modeling by
amide. Fluid Phase Equilib. 2010, 294 (1−2), 180−186. COSMO-RS. Ind. Eng. Chem. Res. 2012, 51 (8), 3483−3507.

5649 DOI: 10.1021/acssuschemeng.6b01542


ACS Sustainable Chem. Eng. 2016, 4, 5640−5650
ACS Sustainable Chemistry & Engineering Research Article

(29) Florindo, C.; Oliveira, F. S.; Rebelo, L. P. N.; Fernandes, A. M.;


Marrucho, I. M. Insights into the Synthesis and Properties of Deep
Eutectic Solvents Based on Cholinium Chloride and Carboxylic Acids.
ACS Sustain. ACS Sustainable Chem. Eng. 2014, 2 (10), 2416−2425.
(30) Klamt, A.; Eckert, F.; Arlt, W. COSMO-RS: An Alternative to
Simulation for Calculating Thermodynamic Properties of Liquid
Mixtures. Annu. Rev. Chem. Annu. Rev. Chem. Biomol. Eng. 2010, 1
(1), 101−122.
(31) University of Karlsruhe and Forschungszentrum Karlsruhe
GmbH, T. V., 1989−2007, 25 GmbH, since 2007; available from
http://www.turbomole.com.
(32) Florindo, C.; Oliveira, F. S.; Rebelo, L. P. N.; Fernandes, A. M.;
Marrucho, I. M. Insights into the Synthesis and Properties of Deep
Eutectic Solvents Based on Cholinium Chloride and Carboxylic Acids.
ACS Sustainable Chem. Eng. 2014, 2 (10), 2416−2425.
(33) Rodriguez, N. R.; Ferre Guell, J.; Kroon, M. C. Glycerol-Based
Deep Eutectic Solvents as Extractants for the Separation of MEK and
Ethanol via Liquid−Liquid Extraction. J. Chem. Eng. Data 2016, 61
(2), 865−872.
(34) Zubeir, L. F.; Lacroix, M. H. M.; Kroon, M. C. Low Transition
Temperature Mixtures as Innovative and Sustainable CO2 Capture
Solvents. J. Phys. Chem. B 2014, 118 (49), 14429−14441.
(35) Neves, C. M. S. S.; Kurnia, K. A.; Coutinho, J. A. P.; Marrucho,
I. M.; Lopes, J. N. C.; Freire, M. G.; Rebelo, L. P. N. Systematic Study
of the Thermophysical Properties of Imidazolium-Based Ionic Liquids
with Cyano-Functionalized Anions. J. Phys. Chem. B 2013, 117 (35),
10271−10283.
(36) Arce, A.; Rodríguez, O.; Soto, A. Experimental Determination of
Liquid−Liquid Equilibrium Using Ionic Liquids: tert-Amyl Ethyl Ether
+ Ethanol + 1-Octyl-3-Methylimidazolium Chloride System at 298.15
K. J. Chem. Eng. Data 2004, 49 (3), 514−517.
(37) 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 (3), 365−372.
(38) Dai, Y.; Witkamp, G.-J.; Verpoorte, R.; Choi, Y. H. Tailoring
properties of natural deep eutectic solvents with water to facilitate their
applications. Food Chem. 2015, 187, 14−19.
(39) Esquembre, R.; Sanz, J. M.; Wall, J. G.; del Monte, F.; Mateo, C.
R.; Ferrer, M. L. Thermal unfolding and refolding of lysozyme in deep
eutectic solvents and their aqueous dilutions. Phys. Chem. Chem. Phys.
2013, 15 (27), 11248−11256.
(40) Palomar, J.; Torrecilla, J. S.; Ferro, V. R.; Rodríguez, F.
Development of an a Priori Ionic Liquid Design Tool. 1. Integration of
a Novel COSMO-RS Molecular Descriptor on Neural Networks. Ind.
Eng. Chem. Res. 2008, 47 (13), 4523−4532.

5650 DOI: 10.1021/acssuschemeng.6b01542


ACS Sustainable Chem. Eng. 2016, 4, 5640−5650

You might also like