Вы находитесь на странице: 1из 15

Journal of Molecular Liquids 275 (2019) 323–337

Contents lists available at ScienceDirect

Journal of Molecular Liquids

journal homepage: www.elsevier.com/locate/molliq

Assessment of carbon dioxide solubility in ionic liquid/toluene/water


systems by extended PR and PC-SAFT EOSs: Carbon capture implication
Mahsa Aghaie, Nima Rezaei, Sohrab Zendehboudi ⁎
Faculty of Engineering and Applied Science, Memorial University, St. John's, NL, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Recently, there has been a growing interest in CO2 capture using ionic liquids (ILs). To determine the CO2 absorp-
Received 5 September 2018 tion potential of ILs and their selectivity in the presence of other gaseous components in the mixture, the solubil-
Received in revised form 5 November 2018 ity of ILs at various operating conditions and the influence of impurities/additives such as water and toluene
Accepted 10 November 2018
should be evaluated. Carbon capture capacity of the ILs can be examined using various methods such as experi-
Available online 13 November 2018
ments, thermodynamic approaches including Equations of State (EOSs), and molecular-based modeling. In this
Keywords:
research work, the extended Peng–Robinson (PR) and Perturbed-Chain Statistical Associating Fluid Theory
Carbon capture (PC-SAFT) EOSs are utilized to assess the solubility and selectivity of CO2 in ILs through comparing the modeling
Ionic liquids results with experimental data. PC-SAFT and PR parameters are determined by employing the experimental den-
Selectivity sity data. Modeling results reveal that the solubility values estimated by the PC-SAFT due to considering the as-
Solubility sociation effects have a lower deviation (or a better performance) than PR EOS based on the magnitudes of
PC-SAFT Absolute Average Deviation (AAD %). The AAD (%) for [bmim][BF4], [bmim][PF6], [bmim][Tf2N], [hmim][Tf2N],
Peng-Robinson [hmim][FAP], and [hmim][FAP] are calculated using PR EOS and PC-SAFT EOS, which are 2–5.7% and 3–7.5%, re-
spectively. Furthermore, ternary systems of CO2 + ILs + water and CO2 + ILs + toluene are modeled to deter-
mine the effect of water and toluene on the gas solubility in ILs and viscosity of ILs with PC-SAFT EOS. Based
on the results, low concentrations of water (0.1 wt%) have a negligible influence on the CO2 solubility in ILs. How-
ever, with increasing the water concentration, the solubility of water reduces significantly. On the other hand, the
viscosity of ILs is reduced with increasing the water concentration. Viscosity reduction in the hydrophilic ILs is
significant. It seems promising to add water up to 10 wt% to hydrophilic ILs, since a decrease in the viscosity to
an amount close to the viscosity of water and a decrease in solubility by 9% are experienced. Finally, the selectivity
of [hmim][Tf2N] in separation of CO2 from mixtures containing H2S, SO2, CH4, and H2 are reported. Based on the
results, [hmim][Tf2N] is not appropriate for separation of CO2 from streams with a high concentration of H2S and
SO2 gases.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction solvents for organic and polymer synthesis, biphasic separations (liq-
uid-liquid extraction), and as a catalyst and catalyst support for reac-
Ionic liquids (ILs) are molten salts at temperatures below 100 °C. tions [4]. One of the important characteristics of ILs is their potential
Their unique properties such as having a low vapor pressure in separating a gas from mixtures of gases. ILs have an exceptionally
(e.g., b10−8 bar [1]) and a high thermal stability (473–673 K [2]) high solubility for some gases (such as CO2 or/and H2S). Hence, they
make them promising materials in different processes such as separa- can be regulated to selectively capture one gas from a mixture of
tion and heat storage. Common ILs are those with cations based on gases [5]. Moreover, due to their unusual low vapor pressure, the sol-
imidazolium, phosphonium, pyridinium, ammonium, and gaunidinium. vent loss can be minimized. These characters give ILs a great potential
The tunability of ILs is another attractive feature, enabling design of in capturing CO2 from flue gases. As ILs do not contaminate the gas
engineered liquids for specific applications (e.g., batteries, fuel cells, phase due to their non-volatility, the selectivity of gases is greater
and lubricants) by selecting different cations, anions, and functionality. with ILs, compared to other solvents. The thermal stability of ILs also
For instance, a specific IL can be tuned to promote a desired viscosity, makes them suitable for high temperature applications.
miscibility (in water), and reactivity with other compounds [3]. ILs are Experimental measurements of CO2 solubility in ILs are extensive in
suitable substitutions for volatile organic solvents. They can be used as the literature, using gravimetric, isochoric saturation, synthetic, chro-
matography, and thermodynamic consistency tests [6–18]. Aki et al.
⁎ Corresponding author. [19] investigated the solubility of CO2 in 10 different imidazolium-
E-mail address: szendehboudi@mun.ca (S. Zendehboudi). based ILs at temperature levels of 25 °C, 40 °C, and 60 °C and pressures

https://doi.org/10.1016/j.molliq.2018.11.038
0167-7322/© 2018 Elsevier B.V. All rights reserved.
324 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

up to 150 bar. They studied the effect of anion, cation alkyl chain length, In this work, the solubility data for CO2 in different ILs is modeled
and the addition of alkyl substituents (on the cation ring) on the using PR and PC-SAFT EOSs in broad ranges of temperature and pres-
solubility of CO 2 in ILs at high pressures. The experiments were sure. The accuracy of PR and PC-SAFT EOSs in predicting the solubility
conducted at the stoichiometric equilibrium condition. The ILs of CO2 in various ILs is compared. Then, the influence of water and tolu-
were charged into the cell which was kept at a constant temperature, ene in the IL phase on the solubility of CO2 in IL-water and IL-toluene
using a water bath. A known amount of CO2 was added to the cell, mixtures and the IL mixture viscosities are investigated. Also, the
after stirring when they reached the equilibrium. The solubility of Henry's constant and the selectivity of CO2 over other gases in [hmim]
CO 2 in ILs was then measured. Shariati and Peters [20] conducted [Tf2N] IL are calculated. Finally, some concluding remarks are given in
experiments based on the synthetic method (in this method, the the last section.
composition of the mixture is known and temperature or pressure
is adjusted until a homogenous phase is achievEd.) at high pressures 2. Thermodynamic modeling
in an autoclave cell for high concentrations of CO2 and in the Cailletet
apparatus for lower concentrations of CO2 in 1-hexyl-3- In this study, the Vaper-Liquid Equilibria (VLE) for the CO2-IL system
methylimidazolium hexafluorophosphate. Jacek Kumełan [21] and Vapor-Liquid-Liquid Equilibria (VLLE) for the CO2-IL-Water system
carried out tests using a high- pressure view-cell technique based are modeled by two thermodynamic models, namely: PRWS EOS (Peng-
on the synthetic method for solubility of CO 2 in [hmim][Tf2N]. Robinson with Wong-Sandler mixing rules) and PC-SAFT (Perturbed
Unfortunately, the experimental data are available in a narrow Chain Statistical Associating Fluid Theory). The PRWS model serves as
region of temperatures and pressures. To study the phase behaviours a reference to show the importance of association in the CO2-IL systems.
of CO2-ILs in a wider range of temperature and pressure, a thermody-
namic model (or an equation of state) is required. 2.1. Peng-Robinson EOS
Lawien et al. [22] reported experimental results for the solubility and
diffusivity of CO2 in [Cnmim][TCM] ILs, using a magnetic suspension bal- The Peng-Robinson EOS [33] is one of the appropriate cubic EOSs in
ance at pressures up to 2 MPa. They employed electrolyte Perturbed- the thermodynamics, which can be written as follows:
Chain Statistical Associating Fluid Theory (ePC-SAFT) [23] EOS to
model the system's behavior and verified it by the real data. Karakatsani RT aðT Þ
P¼ − ð1Þ
et al. [24,25] used the correlated density data at various temperatures to v−b v2 þ 2bv−b2
obtain the parameters of truncated Perturbed Chain-Polar Statistical As-
sociating Fluid Theory (tPC-PSAFT) model [26]. Using the correlated bi- R2 T 2c
aðT Þ ¼ 0:45724 α ðT Þ ð2Þ
nary interactions from tuning solubility data, they determined the Pc
solubility of CO2 and O2 in [bmim][PF6]. They reported an excellent
agreement between the model and experimental results. The tPC- RT c
b ¼ 0:07780 ð3Þ
PSAFT is suitable for systems with weakly dispersed and highly oriented Pc
polar interactions and short-range strong hydrogen bonding interac-
where Pc and Tc are the critical pressure and critical temperature of pure
tions. They modeled the ILs as highly polar and neutral ion pairs. Shiflett
compounds; v stands for the molar volume; a refers to the attraction pa-
et al. [6,27] generated the solubility data of CO2 in three common ILs in-
rameter; b represents the hard-core or co-volume parameter [34]; and
cluding, [bmim][PF6], [bmim][BF4], and [hmim][Tf2N] through the
α(T) denotes the temperature-dependant term, which is expressed as
gravimetric microbalance experiments over a temperature range of
follows:
283.15 K to 348.15 K and pressures up to 2.0 MPa. They simulated the
phase equilibrium behaviors by employing Redlich-Kwong (RK) EOS h   pffiffiffiffiffii2
[28]. Shin et al. [29] reported the solubility of CO2 in [Cn-mim][TfO] α ðT Þ ¼ 1 þ 0:37464 þ 1:54226ω−0:26992ω2 1− T r ð4Þ
(for n = 2, 4, 6, and 8) at pressures up to about 40 MPa and temperature
levels of 303.85 K and 344.55 K, using Peng-Robinson (PR) EOS and van where ω represents the acentric factor and Tr symbolizes the reduced
der Waals (vdW) mixing rules. Cavaliho et al. used PR EOS with Wong- pressure. PR EOS can be utilized for mixtures by applying a suitable
Sandler mixing rule (PRWS EOS) to model the phase behaviors of CO2 mixing rule. The most common mixing rules are the van der Waals
and ILs ([bmim][Tf2N], [C5-mim][Tf2N], [THTDP][NTf2], and [THTDP] mixing rules as given below:
[Cl]) at temperatures up to 363 K and pressures up to 50 MPa [30,31].
c X
X c
There are several modeling data for solubility of CO2 in some ILs; how- am ¼ xi x j aij ð5Þ
ever, they examined EOSs that are not as much accurate as PC-SAFT. i¼1 j¼1
In this work, we investigate the solubility of new ILs with PC-SAFT;
the parameters are determined with high precision. The final solubility X
c X
c

correlated data are more accurate and reliable than the data reported in bm ¼ xi x j bij ð6Þ
i¼1 j¼1
the open sources while comparing the Average Absolute Deviation
(AAD%) results.
where am and bm signify the attraction and co-volume parameters of PR
Fu et al. [32] studied the effect of water content from 0.0067 wt% to
EOS for mixture, respectively; c is the number of components; xi and xj
1.6 wt% on the solubility of CO2 in [bmim][PF6]. They concluded that the
stand for the mole fractions of component i and j, respectively; and aij
deviation in solubility is not N15%. Most researchers only reported the
and bij are the ij pair interaction parameters between molecules i and j
effect of water on the solubility of gases in ILs without considering the
which can be determined from a set of combining rules such as:
impact of water content on the viscosity and solubility of CO2 in hydro-
philic and hydrophobic ILs. In this research study, the influence of water  1=2  
aij ¼ ai a j 1−kij ð7Þ
on decreasing viscosity of ILs and solubility of CO2 in mixture of ILs and
water is evaluated.    
SAFT based EOSs for CO2 separation processes have not been studied bij ¼ bi þ b j =2 1−K 0ij ð8Þ
and developed enough; however, they have shown accurate results for
CO2-ILs and CO2-Amines systems. To the best of our knowledge, the PC- For cubic EOSs, using van der Waals mixing and combining rules,
SAFT EOS has not been utilized in determining solubility of CO2 in some there are two problems. First, the predicted error for the liquid density
of the ILs, yet. of pure fluids and mixtures exceeds 5%. Second, for system of molecules
M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 325

with size or chemical nature dissimilarity, the predictions may not be in where AE∞ signifies the excess Helmholtz free energy at the limit of in-
a good agreement with real data. To overcome these problems, the finite pressure, and GE is the excess Gibbs free energy.
modified mixing rules are introduced, among which Wong-Sandler
[35] has been verified promising in estimating the liquid density [36]. 2.2. PC-SAFT
The Wong-Sandler mixing rule combines the cubic EOS with excess
free-energy (or liquid activity coefficient) models. The PC-SAFT EOS [37] can add the contribution of different terms to
Wong Sandler [35] derived a general form of mixing rules first by the residual Helmholtz free energy of the system, as listed below:
combining the quadratic dependence of the second virial coefficient
on composition and the relationship between the second virial coeffi- ~res ¼ a
a ~hs þ a
~chain þ a
~disp þ a
~assoc ð17Þ
cient and the parameters in a cubic EOS. The second equation in their
mixing rules was derived by taking the limit of the excess Helmholtz ~
where a
res
is the residual Helmholtz free energy of the system. The terms
free energy for a cubic EOS mixture at an infinite pressure. Helmholtz a
hs
~ ,a ~ chain
~disp , and a
,a ~assoc introduce the reference hard sphere, chain
free energy is less pressure dependent and can be approximated by (segment) formation, dispersion, and association contributions to the
the excess Gibbs energy at low pressures where most experimental residual Helmholtz free energy, respectively. a ~ can be obtained from
data are collected. The resultant mixing rules are the pressure indepen- the following equation:
dent, which satisfy the quadratic requirement:
~ ¼ a=RT
a ð18Þ
a
BðT Þ ¼ b− ð9Þ
RT
where T is the absolute temperature and R represents the universal gas
in which, a and b denote the attraction and co-volume parameters of PR constant. The hard-sphere term in Eq. (17) can be calculated using the
EOS. Since the composition dependence of the second virial coefficient Wertheim's theories [38]:
is quadratic, the following relationships are obtained from the statistical
mechanics: ahs 4η−3η2
¼m ð19Þ
RT ð1−ηÞ2
a XX  a
b− ¼ xi x j b− ð10Þ
RT RT ij where m denotes the number of spherical segments per molecule and η
 stands for the reduced density, as defined below:
XX a
RT xi x j b−
i j
RT ij η ¼ τρmv0 ð20Þ
b¼ " # ð11Þ
X ai
GE
RT− xi þ  in which, v0 is the segment molar volume of the fluid that can be
i
bi C
expressed as follows:
" #
GE X ai

a¼b þ xi ð12Þ 3u 3
C bi v0 ¼ v00 1−C exp − ð21Þ
kT

where xi and xj denotes the composition of component i and j, and C⁎ = where u/k stands for the dispersive energy parameter per segment and
0.34657 for the PR EOS. The expression (b − a/RT)ij, cross second virial C = 0.12 for all elements except for the hydrogen [39]. The chain forma-
coefficient, is given by the EOS as follows: tion and association contributions of the Helmholtz free energy are cal-

culated as follows [40]:
a 1  a  aj  
b− ¼ bi − i þ b j − 1−kij ð13Þ " #
RT ij 2 RT RT
achain X
hs
∂ ln g ii ðdii Þ
¼ ð
xi 1−mi Þρ ð22Þ
RT ∂ρ
or i T;x j

 pffiffiffiffiffiffiffiffiffi
a 1  ai a j   " ! #
b− ¼ bi þ b j − 1−kij ð14Þ aassoc X X X X Aiα M iα
RT ij 2 RT ¼ xi ln X Aiα − þ ð23Þ
RT i α A
2 2

In Eqs. (13) and (14), kij stands for the second virial coefficient bi-
nary interaction parameter. The values of the parameter are given in where,
Appendix A for various CO2/IL cases. This parameter has to be obtained " #
hs
experimentally near the conditions of interests. The experimental de- ∂ lng ii ðdii Þ i ξ3 3 dii ξ2
¼ ½ þ
pendence of kij renders the WS model, not a fully predictive scheme. ∂ρ hs
g ii ðdii Þ ð1−ξ3 Þ
2 2 ð1−ξ Þ2
3
Another relationship between a and b is defined by the excess Helm-
T;X j
2 2
ð24Þ
2 2
3dii ξ2 dii ξ2 3dii ξ2 ξ3
holtz free energy AE for the liquid phase, as given below: þ 3
þ 3
þ 4

ð1−ξ3 Þ ð1−ξ3 Þ 2ð1−ξ3 Þ
" #
AE  X  AE ðxÞ
limP→∞ ¼ σ α− xi α i ¼ ð15Þ in which, Miα signifies the number of association sites of segment type α
RT RT
per molecule and XA symbolizes the mole fraction of molecules not
bonded at the association interaction site A. The summation is over all
where σ is an EOS dependant constant and α represents the reduced
association sites on the molecule. The non-bonded fraction XA is calcu-
attractive-term parameter, α = a/bRT. The subscript i corresponds to
lated from the following equation:
pure component values of component i. Wong and Sandler assumed
that AE for the liquid solutions is the same as GE at low pressures, as !−1
M
shown below: X Aiα ¼ 1 þ ρ ∑ x j ∑ ∑ ρX Bjβ ΔAiα Bjβ ð25Þ
j β B¼1
AE ∞ ðxÞ GE ðxÞ
¼ ð16Þ
RT RT where ΔAiβBjβ refers to the association strength which is introduced by
326 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

the following expression: X


6
I1 ðη; mÞ ¼ ai ðmÞηi ð29Þ
Aiα Bjβ
i¼0
  ε
ΔAiβ Bjβ ¼ g iαiβ diαiβ κ Aiα Bjβ σ 3iαiβ exp −1 ð26Þ
kT
X
6
where, εAiαBjβ denotes the association energy and κAiαBjβ is the effective I2 ðη; mÞ ¼ bi ðmÞηi ð30Þ
association volume between site A of molecule i and site B of molecule j. i¼0

The dispersion contribution to the Helmholtz free energy is given by


the following equation [37]: where the ai and bi coefficients are expressed below:
XX XX

adisp ε iαjβ
¼ −2πρI 1 ðη; mÞ xi x j mi m j ziα zjβ σ iαjβ 3 −
RT kT m−1 m−1 m−2
i
!−1
j i j ai ðmÞ ¼ a0i þ a1i þ a2i ð31Þ
XX XX
m m m
∂Z hc εiαjβ
πρm 1 þ Z hc þ ρ I 2 ðη; mÞ xi x j mi m j ziα zjβ σ iαjβ 3
∂ρ i j i j
kT
m−1 m−1 m−2
ð27Þ bi ðmÞ ¼ b0i þ b þ b ð32Þ
m 1i m m 2i
hc −1
where ð1 þ Z hc þ ρ ∂Z∂ρ Þ , and I1 and I2 integrals are given by the follow-
The model constants a0i, a1i, a2i, b0i, b1i, and b2i are reported in refer-
ing relationships:
ence [37].
!−1
∂Z hc 2.3. Modeling algorithm and assumptions
1 þ Z hc þ ρ
∂ρ
!−1 The solubility of CO2 in ILs, using the PR and PC-SAFT EOSs, are obtained
8η−2η2 20η−27η2 þ 12η3 −2η4
¼ 1þm 4
þ ð1−mÞ 2
ð28Þ by applying the equilibrium condition at a constant temperature and pres-
ð1−ηÞ ½ð1−ηÞð2−ηÞ sure through the equality of fugacity of CO2 (Component 1) in the gas phase

Input T, P, and Initial Compositions

Guess Reduced Density ηα, ηβ , KA, and KB

New ρα
Calculate Pcalc with PC-SAFT
and ρβ

False
Pcalc=Psys

True

New KA and KB Calculate Molar Density ρˆ and


Compressibility Factor Z

Calculate Chemical Potential μ


res
and
Fugacity Coefficient ϕ k

(
Equality of Fugacity Δf iαβ = f α i − f β i )

False True
Nonlinear Equation
<ɛ Final P and {xiβ }
Solver

Fig. 1. Flash algorithm to obtain solubility using EOSs such as PC-SAFT.


M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 327

Table 1
Experimental information of pure-IL density data used for parameter fitting (P ≤ 1000 bar).

IL T (K) Ref

[BMIM][BF4] 278–373 [47]


[BMIM][PF6] 278–373 [47]
[BMIM][Tf2N] 278–373 [48]
[hmim][Tf2N] 273–413 [49]
[hmim][PF6] 278–365 [50]
[hmim][FAP] 278–373 [51,52]

gas IL
(^f 1 ) and ionic liquid phase (^f 1 ) for the VLE system, as given below [37]:

^f gas ¼ ^f IL ð33Þ
1 1

Using Eq. (33) and conducting flash calculations with the PC-SAFT
and PR EOSs, the solubility of CO2 in different IL mixtures containing Fig. 2. Comparison of the density of pure [hmim][Tf2N] with temperature, using
water and toluene are determined. experimental data [49] (scatter points) and PC-SAFT EOS (solid line).
Fig. 1 shows the computational procedure to calculate the solubility
of gas components in the liquid phase. The main steps are as follows: acceptor site with a negative charge; only the donor acceptor associa-
(1) Enter pressure (P), temperature (T), and initial compositions of tion bonding A-B type is allowed. In this case, the only non-zero interac-
CO2 and IL mixtures; (2) Guess the reduced density and K-values (distri- tions AB and BA are equivalent, as shown below [39]:
bution coefficients); (3) Input EOS parameters obtained from the den-
sity regression; (4) Obtain Z compressibility factor and fugacity ΔBiα Ajβ ¼ ΔAiα Bjβ ð34Þ
coefficients for all components; (5) Determine the fugacity of each com-
ponent in both phases; (6) Calculate the difference between the fugacity
in which, Δ refers to the association strength. Thus, the mole fraction of
of the component in the phases (Δfvli ); (7) If the values of old and new
molecules at the types of A and B will be equal, as follows:
pressures and magnitudes of fugacity of each component in both phases
are equal, the correct extent of the solubility will be attained, otherwise,
the reduced density and K- values should be updated; and (8) Deter- X Aiα ¼ X Biα ð35Þ
mine the final pressure and solubility of gases in the liquid after the
stop conditions are met. The proton donor can be one of the hydrogen atoms bonded to the
For three phases in equilibrium, the (VLLE) multi-flash calculations carbons on the imidazolium ring. The proton acceptor can be any
need to be performed. An isothermal flash with multiple phases is con- atoms of oxygen and fluorine in the anion [39].
sidered so that two functions/equations including fraction of each phase
in the equilibrium and stability of each phase are solved simultaneously. 3. Parameter estimation
Minimization of the Gibbs free energy is implemented to solve these
two functions. The pure component parameters for PR EOS such as Tc and Pc are not
In this modeling approach, ILs composed of cation and anions are available for the ionic liquids, but they can be determined from the
considered as a single chain molecule with two associating sites. group contribution method such as modified Lyderson-Joback-Reid
Association or self-association occurs when complex formations due technique [42]. The acentric factor ω for PR EOS can be calculated
to the chemical forces act between like molecules and solvation. Cross from the Ambrose-Walton corresponding-states method [43]. For the
association happens between unlike molecules. It is important to define ILs used in this study, the parameters Tc, Pc, and ω are obtained based
how molecules interact with each other using number of association on the research work conducted by Valderrama et al. [42,44]. The binary
sites and their type (electron donor and/or electron acceptor) [41]. interaction coefficients and activity coefficients are determined from
Associating molecules, which tend to form hydrogen bonds, and two fitting the experimental P-xy data for the CO2-IL binary systems within
association parameters are considered in the modeling. The first one is the temperature range of interest. The optimized parameters are
the association energy between sites A and B on molecule i and the sec- attained from minimizing the objective function which is the relative
ond one is the effective association volume between sites A and B on the absolute error between the experimental and correlated solubility
same molecule. data for PR EOS. The objective function (OF) is defined as follows:
In this work, CO2 gas is taken into account as a non-associating and
also the cross association between CO2 and ILs is neglected. ILs are con- X
N
P exp −P calc
i
sidered as associating molecules with the type of 2B. It implies that each OF ¼ min i
ð36Þ
IL molecule have one site type A as a donor site and one site type B as an i
P exp
i

Table 2
Optimized PC-SAFT parameters used for different ILs.

PC-SAFT parameters Absolute deviation (%)


IL Mw (g/gmol) No. data
m σ (A) ɛ/k (K) κAB ɛAB/k (K) min average max

[bmim][BF4] 226 2.2969 5.037 483.53 0.00225 3450 0.00017 0.0056 0.0190 22
[bmim][PF6] 284 2.2820 5.201 456.29 0.00225 3450 0.00027 0.0080 0.0150 22
[bmim][Tf2N] 419 2.7150 5.451 350.50 0.00225 3450 0.00046 0.0044 0.0140 28
[hmim][Tf2N] 447 11.000 3.342 305.25 0.10000 5000 0.00120 0.0141 0.0350 18
[hmim][PF6] 312 2.5500 5.301 487.13 0.00225 3450 0.00240 0.0152 0.0250 18
[hmim][FAP] 239 4.4060 5.145 357.91 0.00225 3450 0.00096 0.0140 0.0410 21
328 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

Table 3 Table 4
Calculated AAD (%) for the solubility of CO2 in ILs based on PR and PC-SAFT EOSs. Magnitudes of AAD (%) achieved in estimating CO2 solubility in IL with PC-SAFT for
[hmim][Tf2N], [hmim][PF6], and [hmim][FAP] at various temperatures.
AAD (%)
EOS IL Data ref. AAD (%)
298 K 313 K 333 K IL Data ref.
298 K 313 K 333 K
PR [bmim][BF4] 3.630 5.806 4.699 [10,19]
[bmim][PF6] 3.164 5.700 4.510 [6] [hmim][PF6] 5.26 – 2.97 [20,59]
[bmim][Tf2N] 6.473 7.513 6.710 [57] [hmim][Tf2N] 3.52 2.29 1.20 [19]
PC-SAFT [bmim][BF4] 2.013 4.163 2.827 [10,19] [hmim][FAP] 1.70 2.08 4.45 [60–62]
[bmim][PF6] 3.010 4.300 5.420 [6]
[bmim][Tf2N] 5.710 5.730 5.120 [57]

where Piexp is the experimental pressure, Picalc represents the calculated


pressure from PR, and N stands for the number of data points.
The parameters of PC-SAFT EOS include the association and non-
association parameters. Three non-association parameters include the
segment number m, the segment diameter σ, and the segment energy

Fig. 3. Solubility of CO2 in different ILs at 298 K, 313 K, and 333 K for (a) [hmim][PF6]
(scatter data from [20,59]), (b) [hmim][Tf2N] (scatter data from [19]), and (c) [hmim]
[FAP] (scatter data from [60–62]). Solid lines show the PC-SAFT predictions and scatter Fig. 4. Effect of water content on solubility of CO2 in [hmim][FAP] at (a) 298 K, (b) 313 K,
points are taken from the experiments. and (c) 333 K.
M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 329

parameter kεB . The two association-based parameters are the association PC-SAFT parameters of pure ILs are obtained by fitting the calculated
energy ε AB and the association volume κ AB. The PC-SAFT pure- density data of ILs to the experimental density data.
component parameters for CO2 are reported in the literature [26]. The The optimal parameters of PC-SAFT that are determined for various
non-association parameters for some ILs are also found in reference ILs (used in this research work) are presented in Table 2. Absolute Aver-
[45]. The association parameters for the Lewis acid-base association be- age Deviation Percentage (AAD %) is used as a measure of model accu-
tween the CO2 (Lewis acid) and the anion site on IL (Lewis base) can be racy, as shown below:
estimated from the following equations [25]:

N !
100 X ρexp
i −ρi
calc
AADð%Þ ¼ ð39Þ
εAB N ρexp
≈ −ΔH dissol ð37Þ i¼1 i
k

! where ρexp
i represents the experimental density; ρcalc
i is the calculated
dissol density; and N denotes the number of data points. Fig. 2 compares the
ΔS
κ AB ≈ exp ð38Þ PC-SAFT and experimental results to forecast pure IL ([hmim][Tf2N])
R
density with temperature. For brevity, we do not show the results for
other ILs here; they are presented in Appendix A. The resultant
The values of the association energy (ε AB/k) and association volume density-temperature diagram shows that the density is well correlated
AB
(κ ) are reported as 3540 K and 0.00225 for CO2-IL system, respectively within the range of temperature. The magnitudes of AAD (%) achieved
[45]. Values such as ε AB/k = 5000 K and κ AB = 0.1 are also given in the in estimating the density of pure ILs are tabulated in Table 2. These
open sources for the system of CO2-IL, which exhibited accurate results values are in the order of 0.01%, which show a great agreement between
for association parameters using the density correlation in some ILs [46]. the PC-SAFT predictions and experimental density data.
Both values are adjusted for each IL and the best magnitudes are The heat capacities of [hmim][Tf2N], [hmim][PF6], and [hmim][FAP]
selected. at different temperatures and pressures are reported in references
PC-SAFT parameters for pure ILs are obtained by fitting the pure [54–56]. The binary interaction coefficients between ILs and CO2 are
component density data in a range of temperature (260–370 K). The bi- also presented in Appendix A.
nary interaction coefficients are also calculated from fitting the VLE data
for CO2-IL system. In the regression of density data for pure ILs to find 4. Results and discussion
the PC-SAFT parameters, it is noticed that the effect of segment energy
parameter is negligible, compared to other parameters. The segment 4.1. Solubility of CO2 in pure ionic liquids (ILs)
number m and the segment diameter σ are found to have a considerable
effect. It means that a slight change in these parameters leads to a signif- The solubility of CO2 in [bmim][BF4], [bmim][PF6], and [bmim][Tf2N]
icant variation in the estimation of density, using the fitted parameters. at 298 K, 313 K, and 333 K is modeled, using PR and PC-SAFT EOSs in a
The references for the density data of used ILs are listed in Table 1. broad pressure range of 1–100 bar. The average absolute deviation per-
In fitting the experimental data of pure IL density to the PC-SAFT centage, AAD (%), between the calculated and experimental equilibrium
EOS, the association parameters are considered constant. The values pressure data is calculated at different temperatures for various CO2-IL
for these parameters are taken from the literature for the systems of systems. The results are shown in Table 3.
CO2-IL [45,53]. A better fit for [hmim][Tf2N] is found with κ AB = 0.1 Based on AAD (%) for PR and PC-SAFT EOSs presented in Table 3, the
and (ε AB/k) = 5000 K as suggested by Ji et al. [46]. For [bmim][BF4], PC-SAFT estimates the solubility with more accuracy for all ILs, com-
[bmim][PF6], [bmim][Tf2N], [hmim][PF6], and [hmim][FAP] ILs, the asso- pared to the PR. The SAFT EOS considers the effect of molecular struc-
ciation parameters are fitted with the association parameters as κ AB = ture and interactions on the bulk properties and phase behaviors,
0.00225 and (ε AB/k) = 3540 K. while the PR EOS assumes the molecules to behave as hard spheres.
The objective function, which is used for the density data regression, The SAFT parameters have the physical significance, featuring diameter
is the maximum-likelihood function using the Britt-Luecke algorithm. of a segment, number of segments in a chain, and the segment-segment

Table 5
Solubility of CO2 in IL + W mixtures and the viscosity of IL and IL + W mixtures at various levels of water content.

IL + W mixture (wt% of W)
Parameter T (K) IL
0 0.1% 1% 2% 5% 10%
a
Viscosity (cP) 298 [hmim][Tf2N] 69.48 69.18 66.54 63.71 55.95 45.05
[hmim][FAP] 88.49 88.09 84.54 80.76 70.40 56.01
[bmim][Ac] 416 389 225.86 135.5 44.08 14.34
313 [hmim][Tf2N] 36.86 36.72 35.42 34.02 30.17 24.69
[hmim][FAP] 43.64 43.61 43.59 43.53 43.36 43.02
[bmim][Ac] 139.86 131.96 81.92 56.53 19.73 7.41
333 [hmim][Tf2N] 18.88 18.6 18.2 17.54 15.7 13.066
[hmim][FAP] 20.72 20.71 20.70 20.66 20.60 20.47
[bmim][Ac] 47.64 45.30 30.04 20.45 8.78 3.77

Solubility of CO2a (mol/mol) 298 [hmim][Tf2N] 0.304 0.301 0.283 0.267 0.238 0.219
[hmim][FAP] 0.478 0.473 0.429 0.390 0.304 0.215
[bmim][Ac] 0.456 0.45 0.436 0.429 0.419 0.41
313 [hmim][Tf2N] 0.265 0.262 0.243 0.224 0.203 0.172
[hmim][FAP] 0.417 0.412 0.37 0.31 0.272 0.177
[bmim][Ac] 0.424 0.421 0.405 0.391 0.378 0.363
333 [hmim][Tf2N] 0.217 0.215 0.197 0.184 0.163 0.131
[hmim][FAP] 0.344 0.339 0.304 0.281 0.235 0.14
323 [bmim][Ac] 0.389 0.388 0.378 0.367 0.334 0.318
a
At P = 20 bar.
330 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

van der Waals interactions [58]. Furthermore, the PC-SAFT EOS incorpo- The mixing rule used to determine the viscosity of the mixture of ILs
rates the influence of molecular association on the volume and energy, and water is based on the Katti and Chaudhri mixing law as shown
which can potentially predict a more precise and reliable phase behav- below [65]:
iors when the association between the molecules becomes important.
The solubility of CO2 in [hmim][Tf2N], [hmim][PF6], and [hmim][FAP]
    ΔGE
at temperatures of 298.15 K, 313.15 K, and 333.15 K (common temper- log ðV ηÞ ¼ x1 log V 1 η1 þ ð1−x1 Þ log V 2 η2 þ ð40Þ
atures for experiments in the literatures) and a broad range of pressure RT
(1–100 bar) is determined using PC-SAFT. The solubility results (P-x) di-
agram is given in terms of mole fraction (xCO2) for different isotherms. where V denotes the molar volume; η symbolizes the absolute viscosity
For each isotherm, the maximum pressure is chosen around the bubble of the mixture; η1 and η2 stand for the viscosity of component 1 and
point pressure of the mixture. The temperature-dependent binary inter- component 2, respectively; x1 represents the mole fraction of compo-
action coefficients are tuned by fitting to the experimental VLE data for nent 1; R is the universal gas constant; T introduces the absolute tem-
the binary systems of CO2-IL. The solubility of CO2 in different ILs is stud- perature; and GE is the excess molar Gibbs free energy. If a mixture
ied at temperatures 298.15 K, 313.15 K, and 333.15 K and pressure does not deviate considerably from the ideal mixture model, GE = 0.
range 1–100 bar, using PC-SAFT. Fig. 3 depicts the solubility values for
[hmim][PF6], [hmim][Tf2N], and [hmim][FAP]. The modeling results
are compared with the experimental data in Fig. 3, as well, exhibiting
an excellent agreement between the experimental data of solubility
and those estimated from the PC-SAFT model. However, at higher pres-
sures and temperatures, some deviations in both systems, particularly
at 313 K are observed. As it is expected, the solubility of CO2 in IL in-
creases with increasing pressure and decreases with increasing temper-
ature. The AAD (%) values for each IL at various temperatures are
reported in Table 4, showing that the modeling outputs are in a very
good match with the experimental data of the solubility for the CO2-IL
systems.

4.2. Solubility of CO2 in mixtures of ILs and water

Adding other solvents to ILs is performed for two main purposes:


1) to reduce the viscosity of ILs by using a less viscous solvent such as
water, and 2) to decrease the overall cost. In this section, the solubility
of CO2 in ILs + water and in ILs + toluene systems are investigated
and compared to that in pure ILs. The solubility of CO2 in hydrophilic
IL [bmim][Ac] and two hydrophobic ILs including [hmim][Tf2N] and
[hmim][FAP] at different water contents (0.1, 1.0, and 10 wt%) and var-
ious temperatures and pressures is investigated. In general, the hydro-
phobicity of the ILs increases with increasing the alkyl cation chain
length [50,63]. Fig. 4 demonstrates the influence of water content in
the liquid medium on the solubility of CO2 at 298 K, 313 K, and 323 K
for [hmim][FAP], being used as the IL. According to Fig. 4, the presence
of water at low concentrations (e.g., b1 wt%) does not have a significant
impact on the solubility of CO2 when compared to the case, using pure IL
([hmim][FAP) at the same temperature and pressure. However, at a
higher water content, the solubility of CO2 is lowered considerably.
When the concentration of water reaches 10 wt%, there is a substantial
decrease (by about 45%) in the solubility of CO2 in IL + W solution (IL
and water mixture), compared to the solubility of CO2 in pure IL at
298 K. On the other hand, the mixture viscosity decreases with increas-
ing the water content, as shown in Table 5. To evaluate the solubility of
CO2 in the mixtures of ILs and water, the Vapor-Liquid-Liquid Equilib-
rium (VLLE) is performed. There are two liquid phases; the first phase
is mostly water at which a minor amount of CO2 (e.g., 3 × 10−5
mol/mol) is dissolved in, and the second phase is mainly IL where
most of the CO2 is dissolved in this phase.
With increasing the concentration of water from 0 to 10 wt%, the vis-
cosity of the system decreases from 88.49 to 56.01 cP at a temperature
and a pressure of 298 K and 20 bar. Adding water to reduce the viscosity
of hydrophobic ILs is not the best solution, since it lowers the solubility
of CO2 significantly. Similarly, we investigate the effect of water content
on the solubility of CO2 in [hmim][Tf2N] /water mixture in Fig. 5 and in
[bmim][Ac]/water mixture in Fig. 6.
The temperature dependency of the viscosity for pure liquids is
expressed by the Vogel-Tammann-Fulcher equation, where various Fig. 5. Effect of water content on the solubility of CO2 in [hmim][Tf2N] at (a) 298 K,
mixing rules can be applied for the mixtures [64]. (b) 313 K, and (c) 333 K.
M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 331

hydrophobic IL. Also, at a higher temperature and pressure, the reduc-


tion in the solubility is more pronounced. By increasing the concentra-
tion of water from 0 to 10 wt%, the mixture viscosity ([hmim][Tf2N]
and water) declines from 69.48 to 45.2 cP at a temperature and a pres-
sure of 298 K and 20 bar. Hence, water with low concentrations such as
0.1 wt% case can be employed as a suitable additive to decrease the vis-
cosity of hydrophilic ILs that has a minimal effect on the solubility at
lower concentrations. The optimum amount of water in the IL + W mix-
tures depends on the composition of the IL systems.
As shown in Table 5, the viscosity of the IL + W mixture decreases
and the solubility of CO2 in the mixture of IL + W also decreases with
increasing the water content. The optimum amount of water needs to

Fig. 6. Effect of water content on the solubility of CO2 in [bmim][Ac] at (a) 298 K, and
(b) 323 K.

There is another equation, based on a similar hypothesis, called the


“ideal” Grunberg and Nissan mixing law [66]. The corresponding equa-
tion is given by the following expression:
   
log ðηÞ ¼ x1 log η1 þ ð1−x1 Þ log η2 ð41Þ

Eq. (41) is also suitable for systems which do not exhibit large devi-
ations from the ideal mixture model.
It is found that the organic solutes including water and toluene stud-
ied in this work alter the viscosity of the IL (ηIL) as shown by the follow-
ing exponential equation:

η ¼ ηIL expð−xs =aÞ ð42Þ

where η and ηIL represent the viscosity of the mixture and the viscosity
of the pure ionic liquid, respectively; xs signifies the mole fraction of the
organic solutes; and a is a constant with a value of 0.231 ± 0.003 for
[hmim][Tf2N] and [hmim][FAP] and 0.022 ± 0.011 for [bmim][Ac].
This confirms the findings attained by Seddon et al. [67]. Therefore,
the viscosity of a mixture can be estimated as a function of the concen-
tration of the dissolved compounds, independent of their polarity [68].
The result for adding water into the moderately hydrophobic IL
[hmim][Tf2N] reveals that at a low concentration of water (0.1 wt%)
there is no effect on the solubility of CO2, similar to the case for
[hmim][FAP]. However, by increasing the concentration of water in
[hmim][Tf2N] to 10 wt%, the solubility of CO2 is reduced by about 20%
at 298 K and 20 bar, as demonstrated in Fig. 5 (and also Table 5). The re- Fig. 7. The effect of water content on the solubility of CO2 in the mixtures of IL with water,
using PC-SAFT EOS for (a) [hmim][FAP], (b) [hmim][Tf2N], and (c) [bmim][Ac]. In panels
duction in the solubility of CO2 in [hmim][Tf2N] by adding 10 wt% water (a) and (b), temperatures of 298, 313, and 333 K and the pressure range 1–100 bar are
is less than that with [hmim][FAP]. The influence of water on the IL + W used; while in panel (c) temperatures 298 and 323 K and pressure range 1–30 bar are
mixture viscosity shows the same trend as that with the strongly utilized.
332 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

Table 6
Viscosity, vapor pressure, and solubility of CO2 in IL + T mixtures for [hmim][Tf2N].

% Toluene (T) in IL + T mixture (mol%)


Parameter 0% 1% 10% 30% 50% 70% 100%
(0 wt%) (0.2 wt%) (1.1 wt%) (8.1 wt%) (17 wt%) (32.4 wt%) (100 wt%)

Viscosity (cP) 69.48 68.79 62.37 46.92 30.45 14.48 0.554


Psat⁎ (bar) 6 × 10−17 0.0011 0.0109 0.0288 0.0399 0.042 0.0366
Solubility⁎ (mol/mol) 0.394 0.39 0.386 0.369 0.351 0.328 0.25
⁎ At T = 298 K and P = 20 bar.

be obtained by considering the thermodynamic characteristics (solubil- (in Fig. 7(b)), and [bmim][Ac] (in Fig. 7(c)). The solubility behavior is in-
ity and selectivity), transport properties (rate of mass transfer), unit op- teresting. For Fig. 6(a) and (b), the solubility values are compiled for dif-
erations (number of equilibrium stages), process design (equipment ferent temperature (298, 313, and 333 K) and pressure (up to 100 bar)
sizing), and energy requirement (for pumping based on the viscosity). conditions. In Fig. 7(c), the ranges of operating conditions for the solu-
For instance, in the case of [hmim][Tf2N] at 298 K and 20 bar, there is bility are temperatures of 298 and 323 K and pressures up to 30 bar.
b2% reduction in the viscosity by adding 10 wt% water, while the solu- Fig. 7 depicts that there is a minimum in the ratio of solubility in IL
bility lowers by about 20%. A careful optimization process requires the + W mixtures and that in the pure ILs, which appears at a lower solubil-
contribution of water content to a reduction in the capital and operating ity limit (0.1–0.3 mol/mol). This will occur at higher temperatures and
costs (due to the viscosity reduction and an increase in the dissolution lower pressures. Among the three ILs shown in Fig. 7, [hmim][FAP]
rate) to be evaluated. and [hmim][Tf2N] are hydrophobic and [bmim][Ac] is hydrophilic. In
Fig. 6 illustrates the effect of water content in the IL on the solubility addition, [hmim][FAP] is more hydrophobic than [hmim][Tf2N]. Based
of CO2 at temperatures 298 K, 313 K, and 333 K within a pressure range on Fig. 7, it is concluded that the effect of water content is different for
of up to 30 bar with [bmim][Ac], which is a hydrophilic IL. As clear from hydrophobic and hydrophilic IL cases. The extent of hydrophobicity
Fig. 6, at low water concentrations (e.g., 0.1 wt%), water content does also affects the contribution/role of water (as an additive to the IL) to
not exhibit an appreciable influence on the solubility of CO2. With in- the solubility of CO2. Comparing Fig. 7 panels (a) and (b) conveys the
creasing the concentration of water to 10 wt%, the solubility of CO2 is re- message that the performance of IL in CO2 capture process is not af-
duced by about 9% at 298 K and 20 bar. The effect of water concentration fected by the water content at concentration 0.1 wt%. By adding 1%
on the IL + W mixture viscosity implies that by increasing the concen- water, a negative impact on the solubility for most of solubility range
tration of water from 0 to 10 wt%, the viscosity of the mixture lowers is noticed and there is no strong correlation between the performance
dramatically from 416 to 14.5 cP at a temperature and a pressure of and temperature. However, by adding 10 wt% water, there is a signifi-
298 K and 20 bar. Thus, water can be considered as a viable additive cant reduction in the solubility of CO2 in the IL + W mixture, compared
to hydrophilic ILs for CO2 separation, as a significant viscosity reduction to that in the pure IL. Moreover, we observe a minimum performance
is experienced without sacrificing solubility. This helps to lower the compared to the baseline, which occurs at an equivalent solubility in
pressure drop in the CO2 capture process that will affect the amount pure IL at 0.1 mol/mol for [hmim][FAP] and about 0.2 for [hmim]
of energy requirement and Greenhouse Gas (GHG) emissions. [Tf2N]. This minimum performance (in terms of solubility) is experi-
Although with the addition of water to the IL, the solubility de- enced at 0.25 mol/mol for [bmim][Ac]. Therefore, there is a shift in the
creases (slightly), the rate of CO2 mass transfer in the IL + W mixture minimum solubility towards higher solubility values with lowering
will be increased as it is proportional to the inverse of viscosity. There- the hydrophobicity characteristic. Also, the ratio of solubility in the IL
fore, it will contribute to a decrease in the capital cost due to smaller + W mixture to that in the pure IL at 10 wt% water is not correlated
process equipment as a result of shorter processing time. with temperature for the most hydrophobic IL ([hmim][FAP]), while
The effect of water on the solubility of CO2 in the mixtures of IL with there is a significant correlation to temperature for the other two ILs
water is also explored by plotting the ratio of solubility in IL with and (in Fig. 7(b) and (c)) for which, the performance decreases due to an in-
without water as a function of solubility in pure IL (without water). crease in the temperature. A shift in the location of the minimum con-
This assessment is shown for [hmim][FAP] (in Fig. 7(a)), [hmim][Tf2N] centration towards lower solubility values (with respect to the

Fig. 8. Solubility of CO2 in pure IL ([hmim][Tf2N]) [19,69], pure toluene [70], and their mixtures [71]. Scatter points show the experimental values and the solid lines represent the PC-SAFT
model results, using the parameters listed in Table 2, and kij = −0.01 for [hmim][Tf2N] + CO2 and kij = 0.105 for toluene + CO2 at 298 K.
M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 333

solubility in the pure IL) is also observed. The drop in the solubility of
CO2 as a result of the water presence is much more predominant for
the hydrophilic IL ([hmim][Ac]); however, it affects a narrower range
of solubility, beyond which the solubility is not much different than
the pure IL (without water). This analysis (as described in Fig. 7(c)) sug-
gests that the addition of a fixed amount of water to the IL will not sig-
nificantly alter the performance of the IL + W mixture beyond the
solubility value of 0.3 mol/mol. This will provide screening criteria on
operating temperatures and pressures, which are suitable for CO2
capture.

4.3. Solubility of CO2 in mixtures of ILs and toluene

Toluene can be added to ILs to decrease their viscosity and make


them more suitable for high temperature carbon capture systems. To
study the influence of toluene (T) content on the viscosity and vapor
pressure of IL + T mixture and on the solubility of CO2 in IL + T mixture, Fig. 10. Effect of additives (water (W) or/and toluene(T)) on the viscosity of IL-additive
[hmim][Tf2N] is utilized. The toluene content in the IL + T mixture in mixtures for [hmim][Tf2N], [hmim][FAP], and [hmim][Ac] at 298 K.
the range of 0–100% is investigated as listed in Table 6. Through adding
toluene, the viscosity of mixture reduces. In this research work, the ef-
fect of adding toluene on the vapor pressure of IL + T mixtures is also slightly while the viscosity is significantly reduced. This reduction in
determined. Based on the results, the vapor pressure of the mixture is the viscosity is expected to lower the dissolution rate of CO2 and also
similar to the vapor pressure of toluene (as a volatile component). the energy input for the processing/transferring of IL. The effect of tolu-
Even by adding only 1 mol% toluene to the ionic liquid, the vapor pres- ene on the solubility of CO2 in the IL + T mixture is also depicted in Fig. 9
sure of the mixture will be 0.0011 bar, which is much higher than the as the ratio of solubility in the IL with and without toluene versus the
vapor pressure of the pure IL. Hence, the effect of ILs in reducing the vol- CO2 solubility in the pure IL (without toluene). Fig. 9 illustrates a differ-
atility of toluene is not important. The solubility of CO2 in pure [hmim] ent trend, compared to the behavior of water as an additive, as seen in
[Tf2N] is higher than that in pure toluene. Hence, to decrease the cost Fig. 7. In this research investigation, a monotonic increase in the perfor-
of using ILs and lower the viscosity, we are able to add some toluene mance of toluene as an additive is observed at higher levels of solubility,
into the system at lower temperatures. The optimum amount of compared to the CO2 solubility in the pure IL as the baseline. In other
added toluene should be determined. At 298 K and 20 bar, the solubility words, the negative effect of toluene as an additive on the solubility of
of CO2 in the mixture is reported in Table 6. It appears that having up to CO2 is less pronounced at higher solubility values (higher pressures
8.1 wt% of toluene in IL + T mixture is acceptable. However, adding and lower temperatures). Moreover, with increasing the toluene con-
N30 mol% toluene causes an increase in the volatility of the mixture centration, the solubility of CO2 is decreased, which is expected. How-
close to the volatility of the toluene. Due to this addition, solubility ever, this difference becomes less at a greater range of CO2 solubility
will be also declined by N6.3%. The validity of the model is tested by (in pure IL).
comparing the experimental data in the literature with those from the The effect of an additive (water or toluene) on the solubility of IL-
PC-SAFT model. As demonstrated in Fig. 8, the solubility predictions additive mixture is illustrated in Fig. 10. It implies that the solubility be-
from PC-SAFT EOS show a great match with the real data for the solubil- haviour for W (water) as an additive to the hydrophobic IL and the sol-
ity of CO2 in pure IL ([hmim][Tf2N]), pure toluene, and mixtures of ubility behaviour of toluene (T) in the hydrophilic IL follow the similar
[hmim][Tf2N] + toluene. However, the solubility values estimated by trend, while adding water to the hydrophilic IL shows a significantly dif-
the PC-SAFT model deviate from the experimental data for pure IL at ferent behaviour. There is a substantial decrease in the solubility of hy-
higher pressures (e.g., N40 bar). drophilic IL ([bmim][Ac] mixture with increasing the water
According to Fig. 8, it is clear that the IL has a higher solubility at the concentration upon the addition of water. At about 10% water content,
same pressure (at 298 K). Toluene exhibits a good solubility in [hmim] the solubility of CO2 in the mixture is about 0.05 of that without the
[Tf2N]. With the addition of toluene to the IL, the solubility decreases water phase. This IL has the highest viscosity among the studied cases.
At 298 K, its viscosity is about 416 cP, which decreases to about 14 cP
by adding 10 wt% water.

4.4. Selectivity

In order to assess the solvent ability to gas separation, the selectivity


of gases into the solvent should be examined. The selectivity can be cal-
culated from the Henry's constant of various gases. Henry's constant

Table 7
PC-SAFT parameters for different gases used in the selectivity study phase [72].

PC-SAFT parameters
Gas
m σ (A) ɛ/k (K) κAB ɛAB/k (K)

CO2 2.604 2.550 151.04 – –


CO2(4C) 2.228 2.731 157.25 0.0287 307.41
CH4 1.000 3.704 150.03 – –
H2S 1.630 3.075 230.35 0.00689 273.55
N2 1.205 3.313 90.96 – –
SO2 2.444 2.680 228.30 – –
Fig. 9. Influence of toluene (T) on the solubility of CO2 in mixtures of [hmim][Tf2N] and
H2 0.487 4.240 33.85 – –
toluene at a temperature 298 K and a pressure range 1–100 bar, using PC-SAFT EOS.
334 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

Table 8 It is worth noting that the Henry's constant is not a suitable measure
Henry's constant (bar) for CO2, CH4, H2, H2S, and SO2 in [hmim][Tf2N], estimated through for the selectivity evaluation when the extent of solubility is at moder-
using PC-SAFT EOS at various temperatures.
ate to high level.
Henry's constant H (bar) Since various gases are present in real flue gases, it appears to be in-
Gas T (K) Ref.
PC-SAFT Experiment % Abs error evitable to study the solubility of each gas in the presence of other gases.
In this research work, the selectivity of CO2 over other gases is deter-
CO2 293 19.11 12.44 + 0.024 53.62 [19,21]
298 20.20 18.91 6.82 mined. Based on the modeling results, it is not easy to separate CO2,
303 21.35 – n/a SO2, and H2S from a flue gas in one single separation stage. Variations
313 26.57 23.64 12.39 of gas selectivity in recycling ILs to use in the regeneration operation
333 33.80 28.56 + 0.029 18.34 of ILs should be analysed carefully. Energy consumption of the desorp-
H2S 303 17.67 17.4 ± 0.01 1.55 [73]
313 21.52 21.7 ± 0.01 0.83
tion column needs to be considered while using the CO2 capture
323 26.19 26.2 ± 0.07 0.04 methods in pilot-scale applications.
333 34.06 33.7 ± 0.03 1.07 In general, the model presented in this research investigation is able
H2 293 886.74 863 ± 0.7 2.75 [74] to simulate the solubility of the common gas mixtures in quantitative
333 704.77 701 ± 0.6 0.54
agreement with the available experimental data.
SO2 298 1.64 1.64 ± 0.01 0.00 [3]
313 2.62 2.29 ± 0.02 14.41
333 4.36 4.09 ± 0.06 6.60 5. Conclusions
CH4 298 328.33 329 ± 23 0.20 [3]
313 380.34 380 ± 31 0.09 In this research work, the reliability and accuracy of PC-SAFT and PR
333 360.09 359 ± 28 0.30
EOSs to forecast the solubility of CO2 in ILs are investigated over practi-
cal ranges of temperature, pressure, and composition. The influences of
relates the fugacity of CO2 in the vapor phase to the molality of CO2 in adding water and toluene on the carbon capture capacity of ILs are ther-
the liquid phase, which is represented by the following expression: modynamically simulated using PC-SAFT EOS. The main conclusion
drawn on the basis of the modeling outputs are listed below:
f CO2 ðT; P Þ
H CO2 ðT Þ ¼ lim ð43Þ • PC-SAFT EOS is more accurate and applicable than PR while predicting
P→0 mCO2 =m 0
the CO2 solubility in ILs. This precision comes from the characteristic/
where m0 = 1 mol/kg; fCO 2(T, P) denotes the fugacity of CO2 in the vapor nature of SAFT EOSs where they take into account the effect of molec-
phase; and mCO 2 stands for the molality of CO2 in the liquid phase. ular structure and interactions on the bulk properties and phase be-
The PC-SAFT EOS is used to compute the Henry's constant as the haviors unlike the cubic EOSs that assume molecules as a hard sphere.
pressure approaches zero. The PC-SAFT parameters for various gases • PC-SAFT EOS estimates the solubility of CO2 in the mixtures of IL
are reported in Table 7 based on the literature. The magnitudes of + water with a great accuracy. Water can decrease the viscosity of
Henry's constant at different temperatures for [hmim][Tf2N] are listed ILs, however, its effect on the solubility of CO2 in the ILs is different
in Table 8. Henry's law constants for CO2, H2S, and SO2 are intensely for the cases of hydrophilic and hydrophobic ILs. The hydrophobic IL
lower than other gases, which is due to the higher solubility of CO2, shows a greater solubility reduction in the presence of water, com-
H2S, and SO2. Using the Henry's constants of all gases, the gas selectivity pared to hydrophilic ILs.
can be calculated as the ratio of Henry's constants of two gases, as writ- • ILs with high hydrophilicity such as [bmim][Ac] are miscible with
ten below: water so that it leads to a significant decrease in the viscosity of the
mixture of [bmim][Ac] and water, although the solubility of CO2 in
Hgas the mixture of [bmim][Ac] + water is not decreased as much as hy-
SCO2 =gas ¼ ð44Þ drophobic ILs.
H CO2
• The solubility of CO2 in the mixtures of IL + toluene is also modeled by
The results of the selectivity of gases are displayed in Table 9. [hmim] utilizing PC-SAFT with a high accuracy. The presence of toluene with
[Tf2N] has a high selectivity for CO2 separation from H2 and CH4, which small concentrations (e.g., 0.002 wt%) can increase the vapor pressure
is due to the very low solubility of H2 and CH4 in [hmim][Tf2N]. Hence, of the ILs, which is not favorable for using mixtures of toluene and ILs.
carbon dioxide capture from gaseous streams containing these gases However, adding up to 0.081 wt% toluene to [hmim][Tf2N] does not
are recommended where the IL is employed. On the other hand, due exhibit a remarkable negative effect on the solubility of CO2. Since tol-
to the low selectivity of CO2 with respect to H2S and SO2, [hmim] uene is much cheaper than ILs, employing mixtures of toluene and ILs
[Tf2N] is not appropriate for separation of CO2 from the stream with a is still an option to be considered in potential CO2 capture operations.
high concentration of H2S and SO2 gases. This is logical as H2S and SO2 • CO2 separation in the presence of other gases is studied in terms of se-
have a high solubility and low Henry's constant in [hmim][Tf2N]. How- lectivity. The Henry's constant and selectivity of CO2 with CH4, H2, H2S,
ever, the low values of selectivity show that the solvent can be used to and SO2 gases are estimated. CO2 separation can be achieved in the
remove these acid gases from the gaseous effluents and to leave the presence of CH4 and H2, while H2S and SO2 are absorbed favourably
CO2 in the gas phase without contaminants. and CO2 remains in the gas phase.

Table 9 Acknowledgements
Selectivity of CO2 over H2S, H2, SO2, and CH4, using [hmim][Tf2N] through thermodynamic
modeling with PC-SAFT EOS.
The authors would like to thank Memorial University, InnovateNL,
Selectivity (S) and the Natural Sciences and Engineering Research Council of Canada
T (K)
CO2/H2S CO2/H2 CO2/SO2 CO2/CH4 (NSERC) for the financial assistance of this research project.
293 n/a 46.40188 n/a n/a
298 n/a n/a 0.08119 16.2540 Appendix A
303 0.82764 n/a n/a n/a
313 0.80994 n/a 0.09861 14.3146 Density data for different ILs correlated with PC-SAFT EOS are com-
333 1.00769 20.85118 0.12899 10.6536
pared with the experimental data available in the literature. Fig. A1
M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 335

represents a very good match between the experimental and correlated [bmim][Ac]
data.Fig. A1Density of a variety of pure ILs versus temperature and com- 1-butyl-3-methylimidazolium acetate
parison of predictions and experimental data. The scatter points and [bmim][BF4]
solid lines represent the real data and PC-SAFT EOS results, respectively. 1-butyl-3-methylimidazolium tetrafluoroborate
[bmim][PF6]
Table A1 lists the magnitudes of the binary interactions of CO2 and 1-butyl-3-methylimidazolium hexafluorophosphate
ILs (kij) at different temperatures where various ILs are utilized. [bmim][Tf2N]
1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide
Table A1 [FAP] tris(pentafluoroalkyl)-trifluorophosphate
Binary interactions of CO2 and IL systems (kij) versus temperature for different ILs used in
[hmim][FAP]
this research work.
1-hexyl-3-methylimidazolium tris (pentafluoroethyl) trifluoro
T (K) phosphate
IL
298 313 333 [hmim][PF6]
[bmim][BF4] −0.016 −0.014 −0.011
1-hexyl-3-methylimidazolium hexafluorophosphate
[bmim][PF6] 0.080 0.090 0.100 [hmim][Tf2N]
[bmim][Tf2N] −0.031 −0.028 −0.025 1-hexyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide
[hmim][PF6] 0.199 0.220 0.240 [THTDP][Cl])
[hmim][FAP] 0.175 0.179 0.182
trihexyltetradecylphosphonium chloride
[hmim][Tf2N] 0.010 0.015 0.019
[bmim][Ac] −0.19 −0.15 −0.12 [THTDP][NTf2]
trihexyltetradecylphosphonium bis(trifluoromethylsulfonyl)imide
CH4 methane
Nomenclatures CO2 carbon dioxide
H2 hydrogen
H2S hydrogen sulfide
SO2 sulphur dioxide
AE∞ Helmholtz excess free energy at the limit of infinite pressure
GE Gibbs excess free energy Acronyms
gas
^f fugacity of comp-1 (solute) in the gas phase
1
^f IL fugacity of comp-1 (solute) in the IL phase
1
ΔBiαAjβ association strength AAD average absolute deviation
XAiα mole fraction of molecules bonded at interaction site A CCS carbon capture and storage
gii site-site radial distribution function EOS equation of state
ε/k segment energy parameter GHG greenhouse gas
a the attraction parameters of PR EOS IL ionic liquid
aassoc Helmholtz free energy of the association reference PC-SAFT perturbed chain statistical associating fluid theory
contribution PR Peng-Robinson EoS
achain Helmholtz free energy of the hard-chain reference VLE vapor liquid equilibria
contribution VLLE vapor liquid liquid equilibria
adisp Helmholtz free energy of the dispersion reference WS Wong Sandler
contribution
ahs Helmholtz free energy of the hard-sphere reference Greek letters
contribution
b co-volume parameters of PR EOS
d segment diameter ε AB association energy
H Henry's constant μ viscosity
I integral γ surface tension
k Boltzmann constant κ conductivity
Kij′ binary interaction coefficient parameter κ AB association volume
kij binary interaction parameter for the attractive parameter ρ density
m number of segments σ viscosity, and also diameter of segment
Pc critical pressure ω acentric factor
P sat vapor pressure η mixture viscosity
R universal gas constant
Tb boiling point References
Tc critical temperature
V molar volume [1] L.P. Rebelo, et al., On the critical temperature, normal boiling point, and vapor pres-
sure of ionic liquids, J. Phys. Chem. B 109 (13) (2005) 6040–6043.
x mole fraction [2] C.P. Fredlake, et al., Thermophysical properties of imidazolium-based ionic liquids, J.
Z compressibility factor Chem. Eng. Data 49 (4) (2004) 954–964.
Z assoc association contribution to the compressibility factor [3] J.L. Anderson, J.K. Dixon, J.F. Brennecke, Solubility of CO2, CH4, C2H6, C2H4, O2, and N2
in 1-Hexyl-3-methylpyridinium Bis (trifluoromethylsulfonyl) imide: comparison to
Z chain hard chain contribution to the compressibility factor other ionic liquids, Acc. Chem. Res. 40 (11) (2007) 1208–1216.
Z disp dispersion contribution to the compressibility factor [4] R.D. Rogers, K.R. Seddon, Ionic liquids–solvents of the future? Science 302 (5646)
Zhs hard sphere contribution to the compressibility factor (2003) 792–793.
[5] M. Aghaie, N. Rezaei, S. Zendehboudi, A systematic review on CO2 capture with ionic
liquids: current status and future prospects, Renew. Sust. Energ. Rev. 96 (2018)
502–525.
Chemical formula [6] M.B. Shiflett, A. Yokozeki, Solubilities and diffusivities of carbon dioxide in ionic liq-
uids:[bmim][PF6] and [bmim][BF4], Ind. Eng. Chem. Res. 44 (12) (2005) 4453–4464.
336 M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337

[7] J.L. Anthony, E.J. Maginn, J.F. Brennecke, Solubilities and thermodynamic properties [41] M.B. Oliveira, et al., Evaluation of the CO2 behavior in binary mixtures with alkanes,
of gases in the ionic liquid 1-n-butyl-3-methylimidazolium hexafluorophosphate, J. alcohols, acids and esters using the Cubic-Plus-Association Equation of State, J.
Phys. Chem. B 106 (29) (2002) 7315–7320. Supercrit. Fluids 55 (3) (2011) 876–892.
[8] Z. Lei, C. Dai, B. Chen, Gas solubility in ionic liquids, Chem. Rev. 114 (2) (2013) [42] J.O. Valderrama, R.E. Rojas, Critical properties of ionic liquids. Revisited, Ind. Eng.
1289–1326. Chem. Res. 48 (14) (2009) 6890–6900.
[9] C.S. Pomelli, et al., Influence of the interaction between hydrogen sulfide and ionic [43] B.E. Poling, et al., The Properties of Gases and Liquids, vol. 5, Mcgraw-Hill, New York,
liquids on solubility: experimental and theoretical investigation, J. Phys. Chem. B 2001.
111 (45) (2007) 13014–13019. [44] J.O. Valderrama, W.W. Sanga, J.A. Lazzús, Critical properties, normal boiling temper-
[10] J.L. Anthony, et al., Anion effects on gas solubility in ionic liquids, J. Phys. Chem. B ature, and acentric factor of another 200 ionic liquids, Ind. Eng. Chem. Res. 47 (4)
109 (13) (2005) 6366–6374. (2008) 1318–1330.
[11] M.E. Zakrzewska, M. Nunes da Ponte, Influence of water on the carbon dioxide sol- [45] Y. Chen, F. Mutelet, J.-N.l. Jaubert, Modeling the solubility of carbon dioxide in
ubility in [OTf]-and [eFAP]-based ionic liquids, J. Chem. Eng. Data 63 (4) (2017) imidazolium-based ionic liquids with the PC-SAFT equation of state, J. Phys. Chem.
907–912. B 116 (49) (2012) 14375–14388.
[12] D. Almantariotis, et al., Influence of fluorination on the solubilities of carbon dioxide, [46] X. Ji, C. Held, G. Sadowski, Modeling imidazolium-based ionic liquids with ePC-SAFT,
ethane, and nitrogen in 1-n-Fluoro-alkyl-3-methylimidazolium Bis (n- Fluid Phase Equilib. 335 (2012) 64–73.
fluoroalkylsulfonyl) amide ionic liquids, J. Phys. Chem. B 121 (2) (2017) 426–436. [47] J. Salgado, et al., Density and viscosity of three (2, 2, 2-trifluoroethanol+ 1-butyl-3-
[13] F. Llovell, et al., Solubility of greenhouse and acid gases on the [C4mim][MeSO4] ionic methylimidazolium) ionic liquid binary systems, J. Chem. Thermodyn. 70 (2014)
liquid for gas separation and CO2 conversion, Catal. Today 255 (2015) 87–96. 101–110.
[14] L.M. Pereira, et al., High pressure solubility of CH4, N2O and N2 in 1-butyl-3- [48] H.F. Almeida, et al., Densities and viscosities of mixtures of two ionic liquids contain-
methylimidazolium dicyanamide: solubilities, selectivities and soft-SAFT modeling, ing a common cation, J. Chem. Eng. Data 61 (8) (2016) 2828–2843.
J. Supercrit. Fluids 110 (2016) 56–64. [49] J.M. Esperança, et al., Pressure−density−temperature (p−ρ−T) surface of [C6mim]
[15] M.B. Shiflett, E.J. Maginn, The solubility of gases in ionic liquids, AICHE J. 63 (11) [NTf2], J. Chem. Eng. Data 53 (3) (2008) 867–870.
(2017) 4722–4737. [50] J. Klomfar, M. Součková, J. Pátek, Low temperature densities from (218 to 364) K and
[16] M.F. Friedrich, et al., Measuring diffusion and solubility of slightly soluble gases in [C up to 50 MPa in pressure and surface tension for Trihexyl (tetradecyl) phosphonium
n MIM][NTf2] ionic liquids, J. Chem. Eng. Data 61 (4) (2016) 1616–1624. Bis (trifluoromethylsulfonyl) imide and dicyanamide and 1-hexyl-3-
[17] L.F. Zubeir, et al., Effect of oxygenation on carbon dioxide absorption and methylimidazolium hexafluorophosphate, J. Chem. Eng. Data 59 (7) (2014)
thermophysical properties of ionic liquids: experiments and modeling using elec- 2263–2274.
trolyte PC-SAFT, Ind. Eng. Chem. Res. 55 (32) (2016) 8869–8882. [51] D. Almantariotis, et al., Absorption of carbon dioxide, nitrous oxide, ethane and ni-
[18] M. Ramdin, T.W. de Loos, T.J. Vlugt, State-of-the-art of CO2 capture with ionic liquids, trogen by 1-Alkyl-3-methylimidazolium (C n mim, n = 2, 4, 6) Tris
Ind. Eng. Chem. Res. 51 (24) (2012) 8149–8177. (pentafluoroethyl) trifluorophosphate ionic liquids (eFAP), J. Phys. Chem. B 116
[19] S.N. Aki, et al., High-pressure phase behavior of carbon dioxide with imidazolium- (26) (2012) 7728–7738.
based ionic liquids, J. Phys. Chem. B 108 (52) (2004) 20355–20365. [52] M. Součková, J. Klomfar, J. Pátek, Temperature dependence of the surface tension
[20] A. Shariati, C.J. Peters, High-pressure phase behavior of systems with ionic liquids: and 0.1 MPa density for 1-C n-3-methylimidazolium tris (pentafluoroethyl)
part III. The binary system carbon dioxide+ 1-hexyl-3-methylimidazolium trifluorophosphate with n = 2, 4, and 6, J. Chem. Thermodyn. 48 (2012) 267–275.
hexafluorophosphate, J. Supercrit. Fluids 30 (2) (2004) 139–144. [53] R.I. Canales, et al., Predicting the solubility of CO2 in toluene + ionic liquid mixtures
[21] J. Kumełan, et al., Solubility of CO2 in the ionic liquid [hmim][Tf2N], J. Chem. with PC-SAFT, Ind. Eng. Chem. Res. 56 (35) (2017) 9885–9894.
Thermodyn. 38 (11) (2006) 1396–1401. [54] E. Gómez, et al., Thermal analysis and heat capacities of 1-alkyl-3-
[22] L.F. Zubeir, et al., Carbon dioxide solubilities and diffusivities in 1-Alkyl-3- methylimidazolium ionic liquids with NTf2–, TFO–, and DCA–anions, Ind. Eng.
methylimidazolium tricyanomethanide ionic liquids: an experimental and model- Chem. Res. 52 (5) (2013) 2103–2110.
ing study, J. Chem. Eng. Data 61 (12) (2016) 4281–4295. [55] Y.U. Paulechka, et al., Heat capacity of ionic liquids: experimental determination and
[23] L.F. Cameretti, G. Sadowski, J.M. Mollerup, Modeling of aqueous electrolyte solutions correlations with molar volume, J. Chem. Eng. Data 55 (8) (2010) 2719–2724.
with perturbed-chain statistical associated fluid theory, Ind. Eng. Chem. Res. 44 (9) [56] J. Safarov, et al., Viscosity, density, heat capacity, speed of sound and other derived
(2005) 3355–3362. properties of 1-butyl-3-methylimidazolium tris (pentafluoroethyl)
[24] E.K. Karakatsani, et al., tPC-PSAFT modeling of gas solubility in imidazolium-based trifluorophosphate over a wide range of temperature and at atmospheric pressure,
ionic liquids, J. Phys. Chem. C 111 (43) (2007) 15487–15492. J. Chem. Eng. Data 62 (10) (2017) 3620–3631.
[25] M.C. Kroon, et al., Modeling of the carbon dioxide solubility in imidazolium-based [57] T. Makino, et al., Physical and CO2-absorption properties of imidazolium ionic liq-
ionic liquids with the tPC-PSAFT equation of state, J. Phys. Chem. B 110 (18) uids with tetracyanoborate and bis (trifluoromethanesulfonyl) amide anions, J.
(2006) 9262–9269. Solut. Chem. 43 (9–10) (2014) 1601–1613.
[26] E.K. Karakatsani, I.G. Economou, Perturbed chain-statistical associating fluid theory [58] P.D. Ting, et al., Application of the PC-SAFT equation of state to asphaltene phase be-
extended to dipolar and quadrupolar molecular fluids, J. Phys. Chem. B 110 (18) havior, Asphaltenes, heavy oils, and petroleomics, Springer 2007, pp. 301–327.
(2006) 9252–9261. [59] Y. Kim, et al., Solubility measurement and prediction of carbon dioxide in ionic liq-
[27] M.B. Shiflett, A. Yokozeki, Solubility of CO2 in room temperature ionic liquid [hmim] uids, Fluid Phase Equilib. 228 (2005) 439–445.
[Tf2N], J. Phys. Chem. B 111 (8) (2007) 2070–2074. [60] X. Zhang, et al., Absorption of CO2 in the ionic liquid 1-n-hexyl-3-
[28] G. Soave, Equilibrium constants from a modified Redlich-Kwong equation of state, methylimidazolium tris (pentafluoroethyl) trifluorophosphate ([hmim][FEP]): a
Chem. Eng. Sci. 27 (6) (1972) 1197–1203. molecular view by computer simulations, J. Phys. Chem. B 113 (21) (2009)
[29] E.-K. Shin, B.-C. Lee, High-pressure phase behavior of carbon dioxide with ionic liq- 7591–7598.
uids: 1-alkyl-3-methylimidazolium trifluoromethanesulfonate, J. Chem. Eng. Data [61] X. Zhang, Z. Liu, W. Wang, Screening of ionic liquids to capture CO2 by COSMO-RS
53 (12) (2008) 2728–2734. and experiments, AICHE J. 54 (10) (2008) 2717–2728.
[30] P.J. Carvalho, et al., High pressure phase behavior of carbon dioxide in 1-alkyl-3- [62] M.J. Muldoon, et al., Improving carbon dioxide solubility in ionic liquids, J. Phys.
methylimidazolium bis (trifluoromethylsulfonyl) imide ionic liquids, J. Supercrit. Chem. B 111 (30) (2007) 9001–9009.
Fluids 48 (2) (2009) 99–107. [63] M.G. Freire, et al., Mutual solubilities of water and the [C n mim][Tf2N] hydrophobic
[31] P.J. Carvalho, et al., High carbon dioxide solubilities in ionic liquids, J. Phys. Chem. B 112 (6) (2008) 1604–1610.
trihexyltetradecylphosphonium-based ionic liquids, J. Supercrit. Fluids 52 (3) [64] P. Navia, J. Troncoso, L. Romaní, Viscosities for ionic liquid binary mixtures with a
(2010) 258–265. common ion, J. Solut. Chem. 37 (5) (2008) 677–688.
[32] D. Fu, et al., Effect of water content on the solubility of CO2 in the ionic liquid [bmim] [65] P. Katti, M. Chaudhri, Viscosities of binary mixtures of benzyl acetate with dioxane,
[PF6], J. Chem. Eng. Data 51 (2) (2006) 371–375. aniline, and m-cresol, J. Chem. Eng. Data 9 (3) (1964) 442–443.
[33] D.-Y. Peng, D.B. Robinson, A new two-constant equation of state, Ind. Eng. Chem. [66] L. Grunberg, A.H. Nissan, Mixture law for viscosity, Nature 164 (4175) (1949) 799.
Fundam. 15 (1) (1976) 59–64. [67] K.R. Seddon, A. Stark, M.-J. Torres, Influence of chloride, water, and organic solvents
[34] A. Firoozabadi, Thermodynamics of Hydrocarbon Reservoirs, McGraw-Hill, New on the physical properties of ionic liquids, Pure Appl. Chem. 72 (12) (2000)
York, 1999. 2275–2287.
[35] D.S.H. Wong, S.I. Sandler, A theoretically correct mixing rule for cubic equations of [68] J. Wang, et al., A volumetric and viscosity study for the mixtures of 1-n-butyl-3-
state, AICHE J. 38 (5) (1992) 671–680. methylimidazolium tetrafluoroborate ionic liquid with acetonitrile, dichlorometh-
[36] V.H. Álvarez, M. Aznar, Thermodynamic modeling of vapor–liquid equilibrium of bi- ane, 2-butanone and N, N–dimethylformamide, Green Chem. 5 (5) (2003) 618–622.
nary systems ionic liquid+ supercritical {CO2 or CHF3} and ionic liquid+ hydrocar- [69] L.F. Zubeir, et al., Solubility and diffusivity of CO2 in the ionic liquid 1-butyl-3-
bons using Peng–Robinson equation of state, J. Chin. Inst. Chem. Eng. 39 (4) (2008) methylimidazolium tricyanomethanide within a large pressure range (0.01 MPa
353–360. to 10 MPa), J. Chem. Eng. Data 60 (6) (2015) 1544–1562.
[37] J. Gross, G. Sadowski, Perturbed-chain SAFT: an equation of state based on a pertur- [70] E.N. Lay, V. Taghikhani, C. Ghotbi, Measurement and correlation of CO2 solubility in
bation theory for chain molecules, Ind. Eng. Chem. Res. 40 (4) (2001) 1244–1260. the systems of CO2+ toluene, CO2+ benzene, and CO2+ n-hexane at near-critical and
[38] M. Wertheim, Fluids with highly directional attractive forces. I. Statistical thermody- supercritical conditions, J. Chem. Eng. Data 51 (6) (2006) 2197–2200.
namics, J. Stat. Phys. 35 (1–2) (1984) 19–34. [71] R.I. Canales, J.F. Brennecke, Liquid-liquid phase split in ionic liquid+ toluene mix-
[39] S.H. Huang, M. Radosz, Equation of state for small, large, polydisperse, and associat- tures induced by CO2, AICHE J. 61 (9) (2015) 2968–2976.
ing molecules, Ind. Eng. Chem. Res. 29 (11) (1990) 2284–2294. [72] N.I. Diamantonis, I.G. Economou, Evaluation of statistical associating fluid theory
[40] W.G. Chapman, et al., New reference equation of state for associating liquids, Ind. (SAFT) and perturbed chain-SAFT equations of state for the calculation of thermo-
Eng. Chem. Res. 29 (8) (1990) 1709–1721. dynamic derivative properties of fluids related to carbon capture and sequestration,
Energy Fuel 25 (7) (2011) 3334–3343.
M. Aghaie et al. / Journal of Molecular Liquids 275 (2019) 323–337 337

[73] M. Rahmati-Rostami, et al., Solubility of H 2 S in ionic liquids [hmim][PF 6],[hmim] [74] J. Kumełan, et al., Solubility of H2 in the ionic liquid [hmim][Tf2N], J. Chem. Eng. Data
[BF 4], and [hmim][Tf 2 N], J. Chem. Thermodyn. 41 (9) (2009) 1052–1055. 51 (4) (2006) 1364–1367.

Вам также может понравиться