Next Article in Journal
Fuzzy Fault Tree Analysis and Safety Countermeasures for Coal Mine Ground Gas Transportation System
Next Article in Special Issue
Solubility of Methane in Ionic Liquids for Gas Removal Processes Using a Single Multilayer Perceptron Model
Previous Article in Journal
Pulsed Electric Field Technology for Recovery of Proteins from Waste Plant Resources and Deformed Mushrooms: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Use of Thermodynamically Consistent Phase Equilibrium Data to Obtain a Generalized Padé-Type Model for the Henry’s Constants of Gases in Ionic Liquids

by
Claudio A. Faúndez
1,*,
Luis A. Forero
2 and
José O. Valderrama
3
1
Departamento de Física, Universidad de Concepción, Casilla 160-C, Concepción 3349001, Chile
2
Facultad de Ingeniería Química, Universidad Pontificia Bolivariana, Medellín 56006, Colombia
3
Centro de Información Tecnológica, Monseñor Subercaseaux 667, La Serena 1700037, Chile
*
Author to whom correspondence should be addressed.
Processes 2024, 12(2), 343; https://doi.org/10.3390/pr12020343
Submission received: 26 September 2023 / Revised: 28 November 2023 / Accepted: 6 December 2023 / Published: 6 February 2024
(This article belongs to the Special Issue Ionic Liquids: Solvent Properties and Organic Reactivity)

Abstract

:
A generalized Padé-type expression is proposed for Henry’s constant of gases in ionic liquids. The constants are determined using an equation of state, and generalized expressions for the Henry’s constants of the gases in the ionic liquids are proposed. The solute gases included in the study were oxygen, hydrogen, and carbon monoxide in three solvent ionic liquids ([MDEA][Cl], [Bmim][PF6], and [Hmim][TF2N]). The Valderrama–Patel–Teja equation of state with the mixing rules of Kwak and Mansoori are employed to correlate the solubility data, to examine the thermodynamic consistency of the experimental data, and to determine the fugacity (fi) for each concentration (xi) of the solute gas in the liquid phase. From these data, the fugacity coefficients (fiL/xi) are determined to obtain Henry´s constant as Hi = lim(fiL/xi) when xi→0. The calculated Henry’s constants are correlated in terms of the temperature and acentric factor of the gases to finally obtain a generalized expression for Henry´s constant, Hi.

1. Introduction

Previous studies have demonstrated that ionic liquids are good candidates for capturing different types of gases and possess several advantages over traditional industrial solvents [1,2]. The number of papers that have reported solubility data for gases in ionic liquids has increased in recent years, and several correlating and predicting methods have been proposed [3,4,5,6,7,8,9]. The values of the solubility of gases in ionic liquids, experimental or calculated, are needed for designing solvents, developing new gas–liquid processes, developing computer simulation methods [2,10], and obtaining better knowledge on the mechanisms involved in the absorption of gases in ionic liquids [4,11].
During the last ten years, research on the different aspects and applications of ionic liquids has notably grown; studies have appeared every day in the scientific literature, including studies on the synthesis of new ionic liquids, their environmental impact and toxicity, their technical and economic analysis, and experimental and theoretical research to determine their physical and physicochemical properties [12,13]. The phase behavior of mixtures involving ionic liquids has also received increasing attention, and several studies have been conducted on aspects such as liquid–liquid equilibrium, liquid–gas equilibrium, the thermodynamic consistency of data, solubility parameters, experimental difficulties, miscibility, and the effects of impurities [11,14,15,16,17,18]. General methods to correlate and predict the phase behavior of mixtures containing ILs are scarce, although the necessary thermodynamic tools are readily available. The main interest in recent years has been in carbon dioxide capture technology by ILs, and potential industrial applications have been analyzed. A good review of these studies and their applications was presented by Zhang et al. [19].
The present work considers the data of three gases (O2, H2, and CO) mixed in three ionic liquids ([Bmim][PF6], [Hmim][TF2N], and [DMEA][CI]), for which experimental data are available in the open literature. Some studies presented in the literature have discussed the difficulties in obtaining accurate data on the solubility of gases in ionic liquids, especially with some gases and in low concentration ranges. Lei et al. [11] discussed this subject and mentioned the following problems: (i) discrepancies in the variation in the solubility of CO as the temperature changes, which are caused by possible large errors in the experiments since the solubility of CO is low; and (ii) for oxygen, it is unclear how the variation in solubility occurs as the temperature increases. Most researchers agree that O2 solubility decreases with increasing temperature, although some researchers argue that there is no influence of temperature; (iii) for hydrogen solubility, there is no agreement, and some authors stated that the solubility of H2 in some ionic liquids first increases with increasing temperature and then decreases as the temperature increases further [20]. These factors impose an additional challenge to modeling solubility data, so a flexible model that can absorb this peculiar behavior is needed. Additionally, a consistency test could provide insight into these discrepancies.
Several different theoretical models, correlations, and equations of state have been used to correlate and predict the solubility of gases in ionic liquids. The concept of Henry´s law is applied in several approaches, including approaches that use the classical equations of state (EoSs) or incorporate activity coefficients. There are also applications of group contribution methods, quantum chemistry calculations, and statistical mechanics-based molecular approaches [21,22,23].
Anthony et al. [24] reported the solubility of nine gases (CO2, C2H4, C2H6, CH4, Ar, 02, C0, H2, and N2) in the ionic liquid [Bmim][PF6] at T = 10 °C, 25 °C, and 50 °C, and they reported the associated Henry´s constants. Kumelan et al. [25] determined the solubilities of CO in the ionic liquid [Bmim][PF6] for temperatures ranging from 293 to 373 K and at pressures up to 9.1 MPa. The solubility–pressure data were correlated through the extended Henry´s law. Kumelan et al. [5,26] reported experimental data on the solubility of H2, CO, and O2 in the ionic liquid [Hmim][TF2N] up to pressures of approximately 10 MPa and at temperatures ranging from 293 to 413 K, and the researchers correlated the data using the extended Henry´s law. Kumelan et al. [27] reported experimental data on the solubility of H2 in the ionic liquid [Bmim][PF6] for temperatures ranging from 313 K to 373 K and at pressures up to approximately 9 MPa.
The solubilities of gases (CO, CO2, N2, H2, O2) in [MDEA][Cl] have been reported by Zhao et al. [7], and the solubility data were correlated using the Krichevsky–Kasarnovsky equation. Raeissi et al. [2] experimentally investigated the solubility of CO in the ionic liquid [Bmim][TF2N] at several temperatures and pressures. Liu et al. [9] studied the solubility of eight common gases in [Emim][Tf2N] using molecular dynamics simulation, and Ramdin et al. [28] used Monte Carlo simulations. Akbari et al. [29] applied a perturbed hard trimer chain (PHTC) equation of state (EoS) to model the solubilities of CO2, H2S, and R143a gases in several ionic liquids for temperatures ranging from 282 K to 353.15 K and at pressures up to 5 MPa. The proposed EoS was applied to estimate Henry’s constant and Gibbs energy of solvation of CO2 in [C8mim][PF6]. The results for Henry’s constant on the mole fraction scale were compared with the experimental data. Eichenlaub et al. [30] studied the solubility of CO2 in 62 ILs using three quantitative structure–property relationship models based on six descriptors. More recently, Jalili et al. [31] reported experimental data on the solubility of CO2 and H2S in the ionic liquid [Bzmim][TF2N] at temperatures ranging from 303 to 343 K and pressures up to approximately 3.1 and 1.6 MPa, respectively. The experimental data were correlated using the extended Henry´s law combined with a simplified Pitzer´s model.
In other developments, equations of state have received some attention. For example, Carvalho et al. [32,33] used the PR EoS with the Wong–Sandler mixing rules, including the UNIQUAC model, to calculate Henry´s constants of CO2 in two ionic liquids. Shiffet and Yokozeki [34,35,36] published several works in which they measured the solubilities of different gases in different types of ILs and treated the data using the simple Redlich–Kwong EoS. Faúndez et al. [18,37,38] used the PR EoS with Kwak–Mansoori mixing rules to predict the solubility of NH3, refrigerants, H2S, and SO2 in ILs. The SAFT equation of state and liquid phase models, such as Wilson’s equation, NRTL, UNIQUAC, and UNIFAC, have been used to describe the liquid phase in gas–ionic liquid mixtures [2,11,17,39]. The regular solution theory has also been considered for modeling gas solubilities in five ILs at low pressures and low mole fractions [14] and for correlating CO2 solubility in imidazolium-type ionic liquids [40].
Some situations in industrial processes involve mixture phase equilibrium in which one or more substances in the mixture (but not all of them) are present at low concentrations in the liquid phase (e.g., below 20% on a molar basis). These situations are considered “solubility” problems, and in general, they are treated as special cases of phase equilibrium. In problems related to gases dissolved in liquids, the fundamental phase equilibrium relation, which is the equality of fugacity of the solute (component at low concentration) in both phases, is certainly valid:
f ¯ i V = f ¯ i L
The fugacity of a component in the vapor phase is usually expressed through the fugacity coefficient φ, while the fugacity of a component in the liquid phase is expressed through the activity coefficient (γ) or through the fugacity coefficient (φ). The fugacity is related to T, P, and the concentration through standard thermodynamic relations [21]. If φ is used for both phases, the method used to solve the phase equilibrium problem is known as the “equation of state method”. If γ is used for the liquid phase and φ is used for the gas phase, the procedure is known as the “gamma-phi method”.
As mentioned above, a simplified procedure is commonly used in solubility cases, although this procedure is also based on the fundamental Equation (1). At low concentrations, the fugacity of a component in the liquid phase follows a linear relation with the solute concentration [22], which can be written as follows:
f ¯ i L = H i x i
In this equation, i represents the solute, and Hi is a parameter known as Henry´s constant, defined as follows:
H i = l i m x i 0 f ¯ i L x i  at constant T
As observed, Equation (3) yields Henry’s constant at the vapor pressure of the solvent, which is corrected for pressure effects, as explained in what follows.
Based on the definition given by Equation (3), Henry’s constant depends on temperature and pressure, considering that the fugacity of the solute in the liquid phase depends on these variables. The dependency of Hi on T and P is described in the literature, and details on the form in which Henry’s law can be applied to higher pressures have been reported [22]. Furthermore, Henry’s law is applicable at low concentrations of solute in liquid solvent (a vague value that depends on the type of solvent and the type of solute). Additionally, if Henry’s law is applied to higher concentrations, a correction term must be added to account for the higher solute concentration. To account for both factors, the Krichevsky–Kasarnovsky–Ilinskaya equation is used, as commonly described in the literature [21].
l n H 1 f ¯ 1 L x 1 = l n H 1 P 2 s + v 1 R T P P 2 s + A R T x 1 2 1
The second term on the right-hand side of Equation (4) was proposed by Krichevsky and Kasarnovsky, and the third term was proposed by Krichevsky and Ilinskaya [21]. In Equation (4), v1 is the partial molar volume at infinite dilution of the gas solute in the liquid phase, and P2S is the vapor pressure of the solvent (ionic liquid) at the system temperature. The term A is a parameter of the Margules equation used by Krichevsky and Ilinskaya to include the activity coefficient of the solute in the liquid solvent.
Of the many EoSs available in the literature, the so-called cubic equations of state derived from the proposal of van der Waals interactions are widely used to treat these types of systems [41]. Among the cubic equations, the Patel–Teja–Valderrama equation (VPT) proposed by one of the authors [42] is employed, and the parametrization developed by Kwak and Mansoori [43] is used. This modified expression of the VPT equation fulfills the basic requirement of the van der Waals model. Therefore, the parameters of the EoS are constant and not functions of the temperature.
The interesting and novel points of this paper include the combination of thermodynamic concepts such as thermodynamic consistency of solubility data, the use of an accurate cubic equation of state, the incorporation of flexible mixing rules, the determination of Henry´s constants of gases in ILs, and their generalization in terms of the temperature and of the acentric factor of the gases.

2. The Equation of State

The cubic EoS derived from the van der Waals proposal can be expressed in the following general form:
P = R T V b a c α ( T ) V V + d + c V d
Most of the cubic equations used in research and academia are obtained by introducing restrictions to the parameters of the general expression Equation (5). For the van der Waals equation, the parameters c and d are zero; for the Soave–Redlich–Kwong equation, c is zero and d = b; for the Peng–Robinson equation, d = c = b; and for the Patel–Teja–Valderrama equation, d equals b. Therefore, the VPT equation is:
P = R T V b a c α ( T ) V V + b + c v b with a c = Ω a R 2 T C 2 P C ;   Ω b = b R T c P c ;   Ω c = c R T c P c α T R = 1 + F 1 T R 0.5 2 Ω a = 0.6612 0.7616 Z c Ω c = 0.5777 1.8718 Z c Ω b = 0.0221 0.2087 Z c F = 0.4628 + 3.5823 ω Z c    8.1942 ( ω Z c ) 2
with
In the modifications presented by Kwak and Mansoori [43], the VPT equation is written as follows:
P = RT V b a + R T d 2 a d R T V ( V   + b ) + c ( V b )
and for mixtures:
P = RT V b m a m + R T d m 2 a m d m R T V ( V + b m ) + c m ( V b m )
The combination rules for the mixture parameters (am, bm, cm, and dm) are as follows:
a m = i j y i y j a i j
a i j = a i a j 1 k i j
a i = a ( T C i ) ( 1 + F i ) 2
b m = i j y i y j b i j
b i j = b i 1 / 3 + b j 1 / 3 2 3 1 l i j
b i = 0.07780 R T c i P c i
c m = i j y i y j d i j
c i j = c i 1 / 3 + c j 1 / 3 2 3 1 m i j
d m = i j y i y j d i j
d i j = d i 1 / 3 + d j 1 / 3 2 3 1 n i j
d i = a ( T c i ) F i 2 R T c i
The parameters included in Equation (8), that is, kij, lij, mij, and nij, are determined using experimental data obtained for the phase equilibrium (PTx data). The initial values of the parameters are estimated, and the phase equilibrium equation is applied to determine the bubble pressure. The deviations between the experimental and calculated pressures are determined. If deviations are above established limits, the parameters are adjusted, and calculations are performed again until deviations are within acceptable defined values. The Levenberg–Mardquart method was applied [44] to optimize the searching of the parameters. The method uses the average absolute deviation between the calculated and the experimental bubble pressure as an objective function. The optimum parameters are those values of kij, lij, mij, and nij that generate the lowest average deviation in bubble pressure and do not generate deviations higher than 10%.
The average absolute deviation between the calculated pressure Pcal and the experimental pressure Pexp is defined as follows:
% Δ P = 100 N i = 1 N P c a l P e x p i P i e x p
Once the optimum parameters are determined, the values are automatically exported to an Excel sheet where a linear correlation between fiL/xi and xi is performed, and Henry´s constant is determined as the position coefficient of the linear model, following Equation (3).

3. Data and Equations Used

In this study, three gases (O2, H2, and CO) in ionic liquids ([MDEA][Cl], [Bmim][PF6], and [Hmim][TF2N]) were analyzed. The ranges of concentration and pressure of the data for each isotherm are presented in Table 1. The critical properties needed to use the VPT equation of state for gases and to calculate Henry’s constants are obtained from the literature [45]. For the ionic liquids, the critical properties were determined by the method of Valderrama and Rojas [46]. A total of 286 data points with approximately 6 to 9 points for each of the forty isotherms were used in the analysis. Table 2 lists the values of these properties.
Within the pressure ranges of the mixtures studied (from 0.8 to 10 MPa), the second term on the right-hand side of Equation (4) can be important. Additionally, for the solubility ranges examined in this study (from 0.008 to 0.233 in mole fraction), the third term on the right-hand side of Equation (4) could reach values that affect the Henry’s constant, depending on the value of Margules parameter A, that was included in the third term of Equation (4). Preliminary calculations generated small, negligible values for (A/RT). These calculations were performed by fitting Henry’s constant at a fixed temperature in terms of pressure and mole fraction in the form l n H 1 = A + B ( P P 2 s ) + C ( x 1 2 1 ) . Data from the literature [7] for similar systems indicate that A = ln ( H o ) ≈ 5 to 6, B = v 1 / R T < 0.01, and C = (A/RT) < 0.001. Therefore, to achieve simplicity without losing accuracy, the third term is not further considered in this study. Equation (4) reduces to the following:
l n f ¯ 1 L x 1 = l n H 1 P 2 s + v 1 R T P P 2 s
If Henry’s constant is calculated from isothermal data by extrapolating ( f ¯ i L / x i ) as
xi → 0, the value obtained corresponds to Henry’s constant at the system temperature and at the pressure of the system when xi = 0, that is, the saturation pressure of the ionic liquid (P2S). In the present study, the temperature ranged from 293 to 413 K, as the vapor pressure of the ionic liquid was very low, approximately between 10−4 and 10−2 bars. Thus, the effect of pressure estimated using Equation (4) to correct Henry’s constant from P2S to P = 1 is negligible. Finally, ln H (P2S) ≈ ln H (P → 0) and Henry’s law are as follows:
l n f ¯ 1 L x 1 = l n H 1 P 0 + v 1 R T P P 2 s
An equation of state is used to determine the fugacity of the solute in the liquid phase that fulfills the fundamental equation of phase equilibrium. To apply an equation of state to mixtures, mixing rules must be used to describe the dependency of the EoS parameters on concentration and of combining rules to describe the interactions between molecules in the mixture.

4. Results

The data used in this work were first tested for thermodynamic consistency, following a method used in the past by the authors. Details of the consistency test can be found elsewhere [18,38,47,48,49] but are briefly summarized in Table 3 to clarify the work presented in this paper.
As detailed in these several papers, the consistency test for isothermal data is reduced to check the fulfillment of Equation (29) in Table 3:
1 P x 1 d P = 1 ( Z 1 ) φ 1 d φ 1 + ( 1 x 1 ) x 1 ( Z 1 ) φ 2 d   φ 2
The left-hand side is calculated from purely experimental data (pressure P and solubility x1), while the right-hand side is determined by calculating the compressibility factor Z and the fugacity coefficients (φ1 and φ2) from an equation of state. If a set of data is consistent, the left-hand side should be equal to the right-hand side of Equation (23), within acceptable defined deviations [47].
Table 4 shows the results of the consistency test applied to all the mixtures considered in this work (40 sets of data). As shown in the table, twenty-five systems were thermodynamically consistent (TC), ten sets of data were not fully consistent (NFC), and only five sets were thermodynamically inconsistent (TI). These sets of data were not considered further in this study for determining and generalizing Henry´s constants. This consideration is in agreement with the meaning of the test based on the Gibbs–Duhem equation; if a set of data does not pass the test, the data are thermodynamically inconsistent and probably incorrect.
After the thermodynamic consistency of the data was checked, Henry’s constants were determined as explained above. This was achieved by using the definition of Henry´s constant (Equation (3)).
To evaluate this limit, the two data points with the lowest solubility (xi→0) for each isotherm were considered. In the (fiL/xi)-vs.-xi plot, a straight line passing through the two points is fitted, so the position coefficient generates Henry´s constant H o at the saturation pressure ( P 2 s ) of the ionic liquid for the given constant temperature.
After H o = H 1 ( P 2 s ) is found, the effect of pressure at a constant given temperature is calculated using the Krichevsky–Kasarnovsky relation:
l n f ¯ 1 L x 1 = l n H ( P ) = l n H o + v 1 R T P P 2 s
In this equation, H o is known, so v1 can be directly calculated from Equation (32). This calculation is simply performed by fixing the position coefficient equal to l n H o when P = ( P 2 s ) and calculating v1 from the slope of the line.
The results for all cases studied are listed in Table 5. In the table, NC indicates that values were not calculated because the available data were found to be thermodynamically inconsistent.
Henry´s constant for the mixtures considered in this study has been reported by other authors, although the ranges of temperature are not the same in all the cases. Therefore, for some of the mixtures, only a general non-quantitative comparison can be carried out. Table 6 presents the values reported by these authors and the results obtained from the generalized correlations proposed in this work.
For the three gases in [MDEA][Cl], Henry’s constants were reported by Zhao et al. [7]. For the three gases in the ionic liquid [Bmim][PF6], the values are provided by Jacquemin et al. [50] and Sharma et al. [51]. For the three gases in [Hmim][TF2N], the values of Henry’s constant are reported by Shi et al. [52], Costa-Gomez [53], Raessi et al. [54], and Kumelan et al. [26].
Comparisons are difficult due to the differences in the temperature range of the data used by each author. For the gases in the ionic liquids [MDEA][Cl], the values of Henry’s constants are determined using data in the same temperature ranges for the literature values and those calculated in this work. The reported values in both cases have deviations lower than 5%. For the gases in [Bmim][PF6], the values show a slightly greater discrepancy but are still acceptable considering that the temperature ranges are different. Not much can be said of the results for the three gases in [Hmim][TF2N] because the temperature ranges, between literature and calculated values, are very different.

5. Generalized Correlations

Henry´s constants determined in this work, using only thermodynamically consistent data, were modeled as a function of the temperature for the three gases. The acentric factor is the property used to distinguish between the gases. The model has the form of a Padé expression with two variables, for each ionic liquid:
l n H o = a + b ω T c + d ω
Table 7 shows the expression for each of the three ionic liquids considered in this study. This equation allows estimating the Henry’s constants for oxygen, hydrogen, and carbon monoxide in the ionic liquids [Bmim][PF6], [MDEA][Cl], and [Hmim][TF2N].
The Henry’s constants for the twelve mixtures calculated from the equation of state ( H o EOS) and those estimated from the Padé equations presented in Table 7 ( H o Padé) are plotted in Figure 1. In this figure, dots are experimental data and lines are calculated values.
As observed in Figure 1, some slight dispersion is found, but these deviations may be acceptable for primary estimations. More importantly, the feasibility of generalizing properties such as Henry´s constants opens great possibilities for future research in the area of fluid property estimation for ionic liquids and mixtures containing ionic liquids.

6. Conclusions

From the results presented and discussed in this paper, the following main conclusions can be drawn: (i) the VPT equation of state accurately correlates with the solubility of gases in ionic liquids (as indicated by the low deviations in bubble pressure values presented in Table 4); (ii) the Kwak–Mansoori method has the required flexibility to converge in all cases studied (as observed by the wide range of values of the binary interaction parameters k12, l12, m12, and n12, presented in Table 4, which range from −3.48 to 2.16); and (iii) the general modeling of Henry´s constants (one equation for any of the gases) opens great possibilities for future research in this area of mixture properties, demonstrating that some generalization is possible.

Author Contributions

Conception and design of study, C.A.F., L.A.F. and J.O.V.; acquisition of data, J.O.V. and L.A.F.; analysis and interpretation of data: C.A.F., L.A.F. and J.O.V.; drafting the manuscript, C.A.F., L.A.F. and J.O.V.; revising the manuscript critically for important intellectual content, C.A.F., L.A.F. and J.O.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Concepción through the research grant VRID N°219.011.062-INV.

Data Availability Statement

All data supporting this research are available in the References section.

Acknowledgments

The authors thank their respective institutions for permanent support. LAF thanks the Faculty of Chemical Engineering of the University Pontificia Bolivariana, Medellin-Colombia. CAF thanks the Direction of Research of the University of Concepción for permanent support.

Conflicts of Interest

The authors declare no conflicts of interest.

Glossary

Symbols
am; bm, cm, dmequation of state parameters for mixtures
ffugacity
HiHenry´s constant
kij, lij, mij, nijbinary interaction parameters
Mmolecular weight
Nnumber of experimental data
Ppressure
P2Svapor pressure of the solvent (ionic liquid) at the system temperature
Pccritical pressure
P C i critical pressure of component i
Rideal gas constant
Ttemperature
Tccritical temperature
T C i critical temperature of component i
Inumber of different components in the system
Vvolume
v1partial molar volume at infinite dilution of the gas solute in the liquid phase
x1experimental solubility of gas in the ionic liquid (mole fraction)
Zcompressibility factor
Zccritical compressibility factor
Abbreviations
EoSequation of state
NCvalues not calculated (in Table 5)
NFCnot fully consistent
TCthermodynamically consistent
TIthermodynamically inconsistent
VPTValderrama–Patel–Teja
Greek Letters
φfugacity coefficient
ωacentric factor
percent deviation
%ΔAarea deviation
%ΔPpressure deviation
Super/subscripts
ccritical
calcalculated
expexperimental
i, jcomponents i and j in a mixture

References

  1. Yokozeki, A.; Shiflett, M.B.; Junk, C.P.; Grieco, L.M.; Foo, T. Physical and Chemical Absorptions of Carbon Dioxide in Room-Temperature Ionic Liquids. J. Phys. Chem. B 2008, 112, 16654–16663. [Google Scholar] [CrossRef]
  2. Raeissi, S.; Florusse, L.J.; Peters, C.J. Purification of flue gas by ionic liquids: Carbon monoxide capture in [bmim][TF2N]. AIChE J. 2013, 59, 3886–3891. [Google Scholar] [CrossRef]
  3. Shariati, A.; Peters, C.J. High-pressure phase behavior of systems with ionic liquids: Measurements and modeling of the binary system fluoroform+1-ethyl-3-methylimidazolium hexafluorophosphate. J. Supercrit. Fluids 2003, 25, 109–117. [Google Scholar] [CrossRef]
  4. Kumełan, J.; Kamps, I.P.-S.; Urukova, I.; Tuma, D.; Maurer, G. Solubility of oxygen in the ionic liquid [bmim][PF6]: Experimental and molecular simulation results. J. Chem. Thermodyn. 2005, 37, 595–602. [Google Scholar] [CrossRef]
  5. Kumelan, J.; Kamps, A.P.-S.; Tuma, D.; Maurer, G. Solubility of H2 in the ionic liquid [hmim][Tf2N]. J. Chem. Eng. Data 2006, 51, 1364–1367. [Google Scholar] [CrossRef]
  6. Yokozeki, A.; Shiflett, M.B. Ammonia Solubilities in Room-Temperature Ionic Liquids. Ind. Eng. Chem. Res. 2007, 46, 1605–1610. [Google Scholar] [CrossRef]
  7. Zhao, Y.; Zhang, X.; Dong, H.; Zhen, Y.; Li, G.; Zeng, S.; Zhang, S. Solubilities of gases in novel alcamines ionic liquid 2-[2-hydroxyethyl (methyl) amino] ethanol chloride. Fluid Phase Equilibria 2011, 302, 60–64. [Google Scholar] [CrossRef]
  8. Mortazavi-Manesh, S.; Satyro, M.; Marriott, R.A. Modelling carbon dioxide solubility in ionic liquids. Can. J. Chem. Eng. 2013, 91, 783–789. [Google Scholar] [CrossRef]
  9. Liu, H.; Dai, S.; Jiang, D.-E. Solubility of Gases in a Common Ionic Liquid from Molecular Dynamics Based Free Energy Calculations. J. Phys. Chem. B 2014, 118, 2719–2725. [Google Scholar] [CrossRef] [PubMed]
  10. Jacquemin, J.; Gomes, M.F.C.; Husson, P.; Majer, V. Solubility of carbon dioxide, ethane, methane, oxygen, nitrogen, hydrogen, argon, and carbon monoxide in 1-buthyl-3-methylimidazolium tetrafluoroborate between temperatures 283 K and 343 K and at pressures close to atmospheric. J. Chem. Thermodyn. 2006, 38, 490–502. [Google Scholar] [CrossRef]
  11. Lei, Z.; Dai, C.; Chen, B. Gas Solubility in Ionic Liquids. Chem. Rev. 2014, 114, 1289–1326. [Google Scholar] [CrossRef] [PubMed]
  12. Kokorin, A. (Ed.) Ionic Liquids: Applications and Perspectives; InTech: Rijeza, Croatia, 2010. [Google Scholar]
  13. Plechkova, N.V.; Seddon, K.R. (Eds.) Ionic Liquids Completely UnCOILed: Critical Expert Overviews; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015. [Google Scholar]
  14. Scovazzo, P.; Camper, D.; Kieft, J.; Poshusta, J.; Koval, C.; Noble, R. Regular Solution Theory and CO2 Gas Solubility in Room-Temperature Ionic Liquids. Ind. Eng. Chem. Res. 2004, 43, 6855–6860. [Google Scholar] [CrossRef]
  15. Arce, A.; Earle, M.J.; Katdare, S.P.; Rodríguez, H.; Seddon, K.R. Phase equilibria of mixtures of mutually immiscible ionic liquids. Fluid Phase Equilibria 2007, 261, 427–433. [Google Scholar] [CrossRef]
  16. Santiago, R.S.; Santos, G.R.; Aznar, M. UNIQUAC correlation of liquid–liquid equilibrium in systems involving ionic liquids: The DFT–PCM approach. Fluid Phase Equilibria 2009, 278, 54–61. [Google Scholar] [CrossRef]
  17. Vega, L.F.; Vilaseca, O.; Llovell, F.; Andreu, J.S. Modeling ionic liquids and the solubility of gases in them: Recent advances and perspectives. Fluid Phase Equilibria 2010, 294, 15–30. [Google Scholar] [CrossRef]
  18. Faúndez, C.A.; Díaz-Valdés, J.F.; Valderrama, J.O. Consistency test of solubility data of ammonia in ionic liquids using the modified Peng–Robinson equation of Kwak and Mansoori. Fluid Phase Equilibria 2013, 348, 33–38. [Google Scholar] [CrossRef]
  19. Zhang, X.; Zhang, X.; Dong, H.; Zhao, Z.; Zhang, S.; Huang, Y. Carbon capture with ionic liquids: Overview and progress. Energy Environ. Sci. 2012, 5, 6668–6681. [Google Scholar] [CrossRef]
  20. Jacquemin, J.; Husson, P.; Majer, V.; Gomes, M.F.C. Influence of the Cation on the Solubility of CO2 and H2 in Ionic Liquids Based on the Bis(trifluoromethylsulfonyl)imide Anion. J. Solut. Chem. 2007, 36, 967–979. [Google Scholar] [CrossRef]
  21. Walas, S.M. Phase Equilibria in Chemical Engineering; Butterworth Pub.: Storeham, MA, USA, 1985. [Google Scholar]
  22. JPrausnitz, M.; Lichtenthaler, R.N.; de Azevedo, E.G. Molecular Thermodynamics of Fluid-Phase Equilibria; Prentice Hall International Series: Woodland Park, NJ, USA, 1999. [Google Scholar]
  23. Smith, J.M.; Van Ness, H.C.; Abbott, M.M. Introduction to Chemical Engineering Thermodynamics, 7th ed.; Mc Graw Hill Co.: New York, NY, USA, 2001. [Google Scholar]
  24. Anthony, J.L.; Maginn, E.J.; Brennecke, J.F. Solubilities and Thermodynamic Properties of Gases in the Ionic Liquid 1-n-Buthyl-3-methyllimidazolium Hexafluorophosphate. J. Phys. Chem. B 2002, 106, 7315–7320. [Google Scholar] [CrossRef]
  25. Kumełan, J.; Kamps, D.P.-S.; Tuma, D.; Maurer, G. Solubility of CO in the ionic liquid [bmim][PF6]. Fluid Phase Equilibria 2005, 228–229, 207–211. [Google Scholar] [CrossRef]
  26. Kumelan, J.; Kamps, A.P.-S.; Tuma, D.; Maurer, G. Solubility of the single gases carbon monoxide and oxygen in the ionic liquid [hmim][Tf2N]. J. Chem. Eng. Data 2009, 54, 966–971. [Google Scholar] [CrossRef]
  27. Kumelan, J.; Kamps, A.P.-S.; Tuma, D.; Maurer, G. Solubility of H2 in the ionic liquid [bmim][PF6]. J. Chem. Eng. Data 2006, 51, 11–14. [Google Scholar] [CrossRef]
  28. Ramdin, M.; Balaji, S.P.; Vicent-Luna, J.M.; Gutiérrez-Sevillano, J.J.; Calero, S.; de Loos, T.W.; Vlugt, T.J.H. Solubility of the Precombustion Gases CO2, CH4, CO, H2, N2, and H2S in the Ionic Liquid [bmim][Tf2N] from Monte Carlo Simulations. J. Phys. Chem. C 2014, 118, 23599–23604. [Google Scholar] [CrossRef]
  29. Akbari, F.; Almutairi, F.D.; Alavianmehr, M.M. Solubility of gases in ionic liquids using PHTC equation of state. J. Mol. Liq. 2019, 276, 553–561. [Google Scholar] [CrossRef]
  30. Eichenlaub, J.; Rakowska, P.W.; Kloskowski, A. User-assisted methodology targeted for building structure interpretable QSPR models for boosting CO2 capture with ionic liquids. J. Mol. Liq. 2022, 350, 118511. [Google Scholar] [CrossRef]
  31. Jalili, A.H.; Mehdizadeh, A.; Ahmadi, A.N.; Zoghi, A.T.; Shokouhi, M. Solubility behavior of CO2 and H2S in 1-benzyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ionic liquid. J. Chem. Thermodyn. 2022, 167, 106721. [Google Scholar] [CrossRef]
  32. Carvalho, P.J.; Álvarez, V.H.; Machado, J.J.; Pauly, J.; Daridon, J.-L.; Marrucho, I.M.; Aznar, M.; Coutinho, J.A. High pressure phase behavior of carbon dioxide in 1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ionic liquids. J. Supercrit. Fluids 2009, 48, 99–107. [Google Scholar] [CrossRef]
  33. Carvalho, P.J.; Álvarez, V.H.; Schröder, B.; Gil, A.M.; Marrucho, I.M.; Aznar, M.; Santos, L.M.N.B.F.; Coutinho, A.P. Specific Solvation Interactions of CO2 on Acetate and Trifluoroacetate Imidazolium Based Ionic Liquids at High Pressures. J. Phys. Chem. B 2009, 113, 6803–6812. [Google Scholar] [CrossRef] [PubMed]
  34. Shiflett, M.B.; Elliott, B.A.; Niehaus, A.M.S.; Yokozeki, A. Separation of N2O and CO2 using Room-Temperature Ionic liquid [bmim][Ac]. Sep. Sci. Technol. 2012, 47, 411–421. [Google Scholar] [CrossRef]
  35. Yokozeki, A.; Shiflett, M.B. Hydrogen purification using room-temperature ionic liquids. Appl. Energy 2007, 84, 351–361. [Google Scholar] [CrossRef]
  36. Yokozeki, A.; Shiflett, M.B. Separation of Carbon Dioxide and Sulfur Dioxide Gases Using Room-Temperature Ionic Liquid [hmim][Tf2N]. Energy Fuels 2009, 23, 4701–4708. [Google Scholar] [CrossRef]
  37. Faúndez, C.A.; Barrientos, L.A.; Valderrama, J.O. Modeling and thermodynamic consistency of solubility data of refrigerants in ionic liquids. Int. J. Refrig. 2013, 36, 2242–2250. [Google Scholar] [CrossRef]
  38. Faúndez, C.A.; Díaz-Valdés, J.F.; Valderrama, J.O. Testing solubility data of H2S and SO2 in ionic liquids for sulfur-removal processes. Fluid Phase Equilibria 2014, 375, 152–160. [Google Scholar] [CrossRef]
  39. Domańska, U.; Marciniak, A. Solubility of 1-Alkyl-3-methylimidazolium Hexafluorophosphate in Hydrocarbons. J. Chem. Eng. Data 2003, 48, 451–456. [Google Scholar] [CrossRef]
  40. Dębski, B.; Hänel, A.; Aranowski, R.; Stolte, S.; Markiewicz, M.; Veltzke, T.; Cichowska-Kopczyńska, I. Thermodynamic interpretation and prediction of CO2 solubility in imidazolium ionic liquids based on regular solution theory. J. Mol. Liq. 2019, 291, 110477. [Google Scholar] [CrossRef]
  41. Valderrama, J.O. The State of the Cubic Equations of State. Ind. Eng. Chem. Res. 2003, 42, 1603–1618. [Google Scholar] [CrossRef]
  42. Valderrama, J.O. A generalized Patel-Teja equation of state for polar and non-polar fluids and mixtures. J. Chem. Eng. Jpn. 1990, 23, 87–91. [Google Scholar] [CrossRef]
  43. Kwak, T.I.; Mansoori, G.A. van der Waals mixing rules for cubic equations of state. Applications for supercritical fluid extraction modeling. Chem. Eng. Sci. 1986, 41, 1303–1309. [Google Scholar] [CrossRef]
  44. Reilly, M. Computer Programs for Chemical Engineering Education; Sterling Swift: Austin, TX, USA, 1972; Volume 2. [Google Scholar]
  45. Reid, R.C.; Sherwood, T.K.; Street, R.E. The Properties of Gases and Liquid, 3rd ed.; McGraw Hill Book Company: New York, NY, USA, 1977. [Google Scholar]
  46. Valderrama, J.O.; Rojas, R.E. Critical Properties of Ionic Liquids. Revisited. Ind. Eng. Chem. Res. 2009, 48, 6890–6900. [Google Scholar] [CrossRef]
  47. Valderrama, J.O.; Álvarez, V.H. A versatile thermodynamic consistency test for in- complete phase equilibrium data of high-pressure gas–liquid mixtures. Fluid Phase Equilibria 2004, 226, 149–159. [Google Scholar] [CrossRef]
  48. Valderrama, J.O.; Reátegui, A.; Sanga, W.W. Thermodynamic Consistency Test of Vapor−Liquid Equilibrium Data for Mixtures Containing Ionic Liquids. Ind. Eng. Chem. Res. 2008, 47, 8416–8422. [Google Scholar] [CrossRef]
  49. Mohammadi, A.H.; Eslamimanesh, A.; Richon, D. Wax Solubility in Gaseous System: Thermodynamic Consistency Test of Experimental Data. Ind. Eng. Chem. Res. 2011, 50, 4731–4740. [Google Scholar] [CrossRef]
  50. Jacquemin, J.; Husson, P.; Majer, V.; Gomes, M.F.C. Low-pressure solubilities and thermodynamics of solvation of eight gases in 1-butyl-3-methylimidazolium hexafluorophosphate. Fluid Phase Equilibria 2006, 240, 87–95. [Google Scholar] [CrossRef]
  51. Sharma, A.; Julcour, C.; Kelkar, A.A.; Deshpande, R.M.; Delmas, H. Mass Transfer and Solubility of CO and H2 in Ionic Liquid. Case of [Bmim][PF6] with Gas-Inducing Stirrer Reactor. Int. Eng. Chem. Res. 2009, 48, 4075–4082. [Google Scholar] [CrossRef]
  52. Shi, W.; Maginn, E.J. Molecular Simulation and Regular Solution Theory Modeling of Pure and Mixed Gas Absorption in the Ionic Liquid 1-n-Hexyl-3-methylimidazolium Bis(Trifluoromethylsulfonyl)amide ([hmim][Tf2N]). J. Phys. Chem. B 2008, 112, 16710–16720. [Google Scholar] [CrossRef]
  53. Gomes, M.F.C. Low-Pressure Solubility and Thermodynamics of Solvation of Carbon Dioxide, Ethane, and Hydrogen in 1-Hexyl-3-methylimidazolium Bis(trifluoromethylsulfonyl)amide between Temperatures of 283 K and 343 K. J. Chem. Eng. Data 2007, 52, 472–475. [Google Scholar] [CrossRef]
  54. Raeissi, S.; Florusse, L.J.; Peters, C.J. Hydrogen Solubilities in the IUPAC Ionic Liquid 1-Hexyl-3-methylimidazolium Bis(Trifluoromethylsulfonyl)Imide. J. Chem. Eng. Data 2011, 56, 1105–1107. [Google Scholar] [CrossRef]
Figure 1. Henry’s constants calculated with the general Padé expression and with the accurate equation of state for gases in: (a) [MDEA][Cl]; (b) [Bmim][PF6]; and (c) [Hmim][TF2N].
Figure 1. Henry’s constants calculated with the general Padé expression and with the accurate equation of state for gases in: (a) [MDEA][Cl]; (b) [Bmim][PF6]; and (c) [Hmim][TF2N].
Processes 12 00343 g001aProcesses 12 00343 g001b
Table 1. Ranges of temperature, solubility, and pressure for the gas + ionic liquid systems considered in this study.
Table 1. Ranges of temperature, solubility, and pressure for the gas + ionic liquid systems considered in this study.
SystemsT (K)Δx (Gas in the IL)ΔP (Bar)References
O2 + [MDEA][Cl]3130.099–0.33420.01–73.41[7]
3180.099–0.31721.20–74.11
3230.094–0.31619.12–74.10
3280.092–0.29120.23–69.28
3330.089–0.29120.52–76.23
H2 + [MDEA][Cl]3130.032–0.13112.93–54.44
3180.030–0.12512.82–58.33
3230.030–0.11015.82–62.02
3280.030–0.10516.32–58.99
3330.029–0.10016.37–58.70
CO + [MDEA][Cl]3130.078–0.23326.13–82.68
3180.075–0.21526.99–83.62
3230.066–0.20525.03–81.90
3280.060–0.19025.83–85.71
3330.057–0.17625.95–83.88
O2 + [Bmim][PF6]2930.008–0.04315.44–90.11[4]
3130.006–0.04311.60–90.58
3330.007–0.04312.27–90.95
3530.007–0.04413.49–89.48
3730.008–0.04514.44–89.11
H2 + [Bmim][PF6]3130.004–0.02216.72–90.99[27]
3330.003–0.02412.35–90.24
3530.004–0.02614.68–91.19
3730.003–0.02910.85–90.10
CO + [Bmim][PF6]2930.006–0.04112.55–91.07[25]
3130.008–0.03916.75–89.63
3340.004–0.0398.93–86.20
3540.008–0.04014.76–89.38
3730.004–0.0427.69–90.17
O2 + [Hmim][TF2N]2930.028–0.11516.08–75.92[26]
3330.024–0.12114.40–85.93
3730.022–0.11914.27–87.42
H2 + [Hmim][TF2N]2930.008–0.04914.65–97.84
3330.014–0.05622.61–93.07[5]
3730.012–0.06215.94–90.78
4130.014–0.07416.27–94.99
CO + [Hmim][TF2N]2930.020–0.10914.29–89.90
3330.020–0.10415.65–91.98[26]
3730.018–0.10514.97–91.91
4130.025–0.11019.57–98.11
Table 2. Properties of the solutes and solvents considered in this study. The data for gases are from Reid et al. [45], and the data for the ILs are from Valderrama and Rojas [46].
Table 2. Properties of the solutes and solvents considered in this study. The data for gases are from Reid et al. [45], and the data for the ILs are from Valderrama and Rojas [46].
ComponentsMTc/KPc/MPaVc (m3/kmol)ω
O231.99154.605.040.073400.0250
H22.0233.191.310.064150.2160
CO28.01132.903.500.093200.0660
[Bmim][PF6]284.2719.41.730.762500.7917
[Hmim][TF2N]447.41298.82.391.104400.3893
[MDEA][Cl]154.60720.703.610.451601.1247
Table 3. Main expressions of the thermodynamic consistency test used in this work.
Table 3. Main expressions of the thermodynamic consistency test used in this work.
AGibbs–Duhem equation in terms of fugacity coefficients
i = 1 I x i d G R R T = H R R T 2 d T + v R R T d P (24)
BAt constant temperature:
( Z 1 ) P d P = x 1 d ( L n ϕ 1 ) + x 2 d ( L n ϕ 2 ) (25)
And arranging terms:
1 P d P = x 1 ( Z 1 ) ϕ 1 d ϕ 1 + ( 1 x 1 ) ( Z 1 ) ϕ 2 d ϕ 2 (26)
1 P x 1 d P = 1 ( Z 1 ) 1 d 1 + ( 1 x 1   ) x 1 ( Z 1 ) 2 d 2 (27)
CExperimental and estimated areas:
A P = 1 P x 1 d P (28)
A ϕ = 1 ( Z 1 ) ϕ 1 d ϕ 1 + ( 1 x 1 ) x 1 ( Z 1 ) ϕ 2 d ϕ 2 (29)
DIndividual percent area deviation in the range [−20% to +20%]
A i = 100 A ϕ A P A P i (30)
Individual deviation in the system pressure in the range [−10% to +10%]
P i = 100 P i c a l P i e x p P i e x p i (31)
Table 4. Optimum values of binary interaction parameters k12, l12, m12, n12, and deviations in the correlated bubble pressure and in the area test. In the last column are the results of the area test.
Table 4. Optimum values of binary interaction parameters k12, l12, m12, n12, and deviations in the correlated bubble pressure and in the area test. In the last column are the results of the area test.
SystemsT (K)k12l12m12n12∣P%∣Pmax%∣A%∣Results
O2 + [MDEA][Cl]3130.0935−0.28310.9772−0.22841.3−3.39.7TC
318−1.9379−0.1677−0.7994−1.08912.84.615.2NFC
323−1.4627−0.1710−0.3144−1.36020.8−1.98.1NFC
328−1.3057−0.3855−0.18850.82911.3−1.68.0TC
333−0.9675−0.21310.1288−1.06060.50.95.5TC
H2 + [MDEA][Cl]313−0.8914−0.41101.00740.65830.7−2.18.0TC
318−1.0617−0.30600.9274−0.84991.4−4.312.7NFC
323−1.3801−0.42790.78540.98633.26.915.1NFC
328−1.2594−0.28080.8802−1.57491.22.722.5TI
333−0.6554−0.31941.0469−0.72021.03.013.5TI
CO + [MDEA][Cl]313−0.1329−0.24890.8007−0.05840.4−0.95.8TC
318−0.3208−0.27940.63230.34380.2−0.42.6TC
3230.6315−0.09581.3610−1.50040.4−0.77.5TC
3280.7043−0.26121.38640.52230.40.75.4TC
3330.0512−0.28310.90720.42852.0−3.57.9TC
O2 + [Bmim][PF6]2930.93700.8163−1.4508−1.34710.6−1.55.1TC
3130.11870.3718−1.3672−0.77250.20.42.0TC
333−0.37210.0417−1.0753−0.71780.30.52.8TC
3530.49830.5920−1.4202−0.33410.40.82.1TC
373−0.38220.0141−0.9040−0.73610.30.52.2TC
H2 + [Bmim][PF6]3130.40470.0482−1.49580.98220.4−1.54.2TC
3330.27820.0433−1.34950.77200.2−0.42.4TC
353−1.1105−0.1328−0.8160−0.96120.40.92.9TC
373−0.94770.0350−1.0469−0.95070.20.42.3TC
CO + [Bmim][PF6]2930.4407−0.65691.2148−0.56280.8−1.64.6TC
313−0.2501−1.29531.37160.85820.71.47.3TC
3340.6404−0.33890.8296−1.00460.30.72.9TC
3540.80750.4370−0.6824−0.37931.02.44.3TC
373−0.37400.0885−0.9212−0.62120.4−1.53.6TC
O2 + [Hmim][TF2N]293−0.0655−0.67000.4426−0.18422.0−3.428.7NFC
333−0.5649−0.0699−0.5849−3.47823.8−4.38.7TC
373−0.0822−0.71060.49300.08862.8−3.46.1TC
H2 + [Hmim][TF2N]293−0.7862−0.59610.5086−1.81600.5−1.236.4TI
333−0.0850−1.08451.49511.56270.20.634.3TI
373−0.8501−1.03561.00291.37140.2−0.531.8NFC
413−1.4295−0.77720.4772−0.33540.20.58.5NFC
CO + [Hmim][TF2N]293−0.6299−0.2578−0.5352−2.11542.0−3.079.6TI
333−1.2862−0.5464−1.65841.42980.7−1.138.6NFC
373−0.9619−0.5485−0.5671−0.05951.1−2.514.7NFC
413−1.3618−0.6958−1.26752.16380.5−1.912.0NFC
Table 5. Values of the calculated Henry´s constants H o (bar) obtained for all mixtures studied.
Table 5. Values of the calculated Henry´s constants H o (bar) obtained for all mixtures studied.
SystemsT (K) H o (Bar)v1 (m3/kmol)
O2 + [MDEA][Cl]313194.410.0313
318206.370.0325
323193.020.0597
328212.310.0343
333219.590.0550
H2 + [MDEA][Cl]313392.750.0446
318417.450.0664
323518.490.0530
328NCNC
333NCNC
CO + [MDEA][Cl]313327.390.0270
318342.960.0397
323362.710.0346
328413.780.0312
333444.850.0262
O2 + [Bmim][PF6]2931765.300.0312
3131786.100.0341
3331791.900.0356
3531791.900.0350
3731767.200.0381
H2 + [Bmim][PF6]3134133.200.0210
3333711.990.0264
3533352.030.0255
3733061.100.0183
CO + [Bmim][PF6]2931964.600.0288
3141903.100.0456
3341926.990.0490
3541883.670.0601
3731971.350.0384
O2 + [Hmim][TF2N]293558.430.0364
333588.480.0534
373636.340.0470
H2 + [Hmim][TF2N]293NCNC
333NCNC
3731351.440.0446
4131338.80.0456
CO + [Hmim][TF2N]293NCNC
333739.580.0552
373797.930.0394
413757.310.0702
Table 6. Comparison of the calculated Henry´s constants for all the systems considered in this study.
Table 6. Comparison of the calculated Henry´s constants for all the systems considered in this study.
CaseSystemsRange of T (K) H o (Bar)
1O2 + [MDEA][Cl]313–333194–220This work
313–333190–220Zhao et al. [7]
2H2 + [MDEA][Cl]313–323393–519This work
313–333440–560Zhao et al. [7]
3CO + [MDEA][Cl]313–333327–445This work
313–333320–440Zhao et al. [7]
4O2 + [Bmim][PF6]293–3731765–1792This work
293–3431024–1610Jacquemin et al. [50]
5H2 + [Bmim][PF6]373–4133061–4133This work
293–3733890–4350Sharma et al. [51]
6CO + [Bmim][PF6]293–3731883–1971This work
283–3431201–1356Jacquemin et al. [50]
7O2 + [Hmim][TF2N]293–373558–636This work
313–373623–790Shi et al. [52]
8H2 + [Hmim][TF2N]373– 4131339–1351This work
293–3431258–1984Costa-Gomez [53]
293– 3531501–1981Raessi et al. [54]
9CO + [Hmim][TF2N]333–413740–798This work
293–413319–347Kumelan et al. [26]
Table 7. Expressions for the Henry’s constants of the gases in three ionic liquids.
Table 7. Expressions for the Henry’s constants of the gases in three ionic liquids.
Ionic liquid ln   H o for any of the gases
[Bmim][PF6] ln   H o = (18.1 − 0.040⋅ωT)/(2.44 − 2.68⋅ω)
[MDEA][Cl] ln   H o = (−2.69 + 12.81⋅ωT)/(2.08 + 652.6⋅ω)
[Hmim][TF2N] ln   H o = (597.8 + 0.7934⋅ωT)/(93.66 + 5.91⋅ω)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Faúndez, C.A.; Forero, L.A.; Valderrama, J.O. Use of Thermodynamically Consistent Phase Equilibrium Data to Obtain a Generalized Padé-Type Model for the Henry’s Constants of Gases in Ionic Liquids. Processes 2024, 12, 343. https://doi.org/10.3390/pr12020343

AMA Style

Faúndez CA, Forero LA, Valderrama JO. Use of Thermodynamically Consistent Phase Equilibrium Data to Obtain a Generalized Padé-Type Model for the Henry’s Constants of Gases in Ionic Liquids. Processes. 2024; 12(2):343. https://doi.org/10.3390/pr12020343

Chicago/Turabian Style

Faúndez, Claudio A., Luis A. Forero, and José O. Valderrama. 2024. "Use of Thermodynamically Consistent Phase Equilibrium Data to Obtain a Generalized Padé-Type Model for the Henry’s Constants of Gases in Ionic Liquids" Processes 12, no. 2: 343. https://doi.org/10.3390/pr12020343

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop