Abstract

Novel aqueous (aq) blends of N-methyldiethanolamine (MDEA), sulfolane (TMSO2), and 1-butyl-3-methyl-imidazolium acetate ([bmim][Ac]) with amine activator 2-methylpiperazine (2-MPZ) are analyzed through conductor-like screening model for real solvents (COSMO-RS) for possible application in the chemisorption of CO2. The molecules associated are analyzed for their ground-state energy, σ potential, and σ surface. Thermodynamic and physicochemical properties have been assessed and paralleled with the experimental data. Vapor pressure of the blended systems and pure component density and viscosity have been compared successfully with the experimental data. Important binary interaction parameters for the aqueous blends over a wide temperature, pressure, and concentration range have been estimated for NRTL, WILSON, and UNIQUAC 4 models. The COSMO-RS theory is further applied in calculating the expected CO2 solubility over a pressure range of 1.0–3.0 bar and temperature range of 303.15–323.15 K. Henry’s constant and free energy of solvation to realize the physical absorption through intermolecular interaction offered by the proposed solvents. Perceptive molecular learning from the behavior of chemical constituents involved indicated that the best suitable solvent is aq (MDEA + 2-MPZ).
1. Introduction
The quest to reduce CO2 emissions via different routes has been a major concern over the past few decades. The process intensification of the existing CO2 capture techniques and introduction of novel solvents for achieving the same through chemisorption or physisorption has been proposed by many researchers.1,2 An extensive lab-scale development of vapor–liquid equilibria,3 kinetic studies,4 thermophysical properties,5,6 calculation of binary interaction parameters,7 improvement in the existing modeling techniques,8 proposing new correlational analysis,9 optimizing the reaction or process parameters,10 defining the structural property relationships,11 heat of absorption,12 etc. are an integral part of the development of new solvents for CO2 or other acid gas absorption applications. However, most of the experimental investigations at the pilot scale tend to be very expensive, and therefore, an efficient solvent screening through various analyses of the proposed solvents is the need of the hour to arrive at a conclusion of their possible applicability at the plant scale. Conclusively, researchers have shifted the research a step back at the quantum-molecular level to understand the basic phenomena of the novel solvents than to lab- or pilot-scale studies for achieving the anticipated CO2 separation.13−17
The conductor-like screening model for real solvents (COSMO-RS) is a method of quantum chemical calculations grouped with statistical thermodynamics. The same has been widely applied for accurate prediction of thermodynamic or essential behavioral properties such as Gibb’s free energy, activity coefficients, partition coefficients, etc. Calculating the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) gaps to define the polarities or electronegativities associated with a specific molecule yields a useful insight for behavioral analysis of any molecule.18 This significantly reduces the amount of time, energy, and cost associated with the detailed study required for a set of selected solvents. The variations of the predictive capabilities of COSMO-RS can be deduced from a brief literature survey. COSMO-RS applications in the membrane separation processes through COSMOmic simulations have been widely studied in the recent past, indicating the efficiency of estimation of partition coefficients with fewer deviations from experimental studies in comparison to molecular dynamics simulation evaluations.19−23 The self-assembly of the surfactants Triton X-114 and Triton X-100 in water solutions at different concentrations and temperatures has been reported,24 where the partition behavior of neutral solutes and micellar structures was studied. The research findings are of great value for extraction and purification through membrane processes using surfactants. The partition behavior of various amino acids at distinct ionization states was predicted successfully using COSMO-RS indicated25 for the extraction of biomolecules using surfactants. A detailed octanol/water partition coefficient prediction through the conceptual application of molecular dynamics and the subsequent input to COSMOmic results in efficient assemblies of complex multiphase fluids are also reported. However, the prediction accuracy depends highly on the i/p molecular structures acquired from molecular dynamics simulations.26 Similar studies can also be performed with respect to the CO2 capture process of adsorption through membranes.
Further with respect to CO2 absorption or adsorption, the important properties of understating are vapor pressure, excess enthalpy, excess Gibb’s energy, infinite dilution activity coefficient (IDAC), activity coefficients, chemical potentials, physiochemical properties such as density, viscosity, refractive index, toxicity, biodegradability, thermal stability, corrosion behavior, and CO2 solubility at a specified temperature and pressure, etc. Thorough knowledge of these properties plays a significant role in selecting a solvent. This is owing to the fact that each of these properties signifies the overall effectiveness of the CO2 capture process. For instance, if the selected solvent exhibits a high viscosity at the absorption temperature and pressure, the overall pumping costs from the regenerator and absorber columns are expected to be very high. On the other hand, if the chemical potentials of the selected solvents are negative in the selected range of temperature, the solvents will not react with CO2 chemically at all, leading to merely physisorption in the capture process.
Assessing the possibilities of COSMO-RS for the important parameters of prediction and analysis, the current study is proposed for CO2 capture through absorption. The nonlinearity exhibited in the absorption of CO2 in previously studied solvents27,28 suggests that using a quantum calculation method may yield an accurate estimation of various involved thermodynamic properties. The selection of proposed blends is based on an extensive literature survey. The analysis of amine activator 2-methylpiperazine (2-MPZ)4,29,30 for enhancement of CO2 capture is carried out in tertiary amine N-methyldiethanolamine (MDEA),31,32 physical solvent sulfolane (TMSO2),33,34 and imidazolium-based ionic liquid 1-butyl-3-methyl-imidazolium acetate ([bmim][Ac]).35,36 MDEA is reported to exhibit a high CO2 loading, although it has a low reaction rate. The latter is owing to the fact that MDEA acts only as a weak base to release free OH– that further interacts with CO2. Further, the reaction of MDEA with CO2 is less exothermic when compared to primary amines. MDEA additionally offers various advantages such as high thermal and chemical degradation opposition and low solution vapor pressure in comparison to monoethanolamine (MEA) and diethanolamine (DEA).37,38 TMSO2 is also considered for its qualities of high physical absorption capacity, high thermal stability, low heat of absorption, and lower corrosion characteristics.33,34 1-Butyl-3-methylimidazolium acetate, being a room-temperature ionic liquid, offers an insignificantly low vapor pressure and is recognized for its thermal stability and CO2 capture capacity.
The solvents are chosen in such a manner as to provide molecular and thermodynamic insights into each category of solvents. The base solvents, i.e., MDEA, TMSO2, and [bmim][Ac], are proposed in the concentration range of 2.5–3.5 mol·kg–1, whereas the concentration range of the activator is varied from 0.5 to 1.5 mol·kg–1. The selected temperature range is 303.15–323.15 K in view of application in absorption phenomena. The variables studied in the respective range of solvents are based on the recommendations in the literature.29−39 Various important properties such as σ potential, σ profile, vapor pressure, pure component density and viscosity, infinite dilution activity coefficients, activity coefficients, Gibb’s free energy, chemical potential, CO2 solubility, Henry’s law coefficient, etc. have been analyzed through the COSMO-RS theory using COSMOtherm to evaluate the applicability of the proposed solvents.
2. Computational Methods and Theory
The molecules under study, i.e., N-methyldiethanolamine (MDEA), sulfolane (TMSO2), 1-butyl-3-methylimidazole, cation ([bmim]), acetate anion ([Ac]), H2O, CO2, and 2-methylpiperazine (2-MPZ), were selected within COSMOtherm (COSMOlogic GmbH, Leverkusen, Germany). Single conformers with the least ground-state energy were selected for each of the molecular compounds obtained with BP-TZVPD-FINE-level COSMO calculations that incorporate a full geometry optimization by density functional theory (DFT) using the Becke and Perdew (BP) functional with the triple-ζ valence polarized (TZVP) basic set. The detailed specifications of the chemicals for the experimental work are presented in Table 1. The analysis of σ surface, σ potential, and vapor–liquid equilibrium and estimation of vapor pressure, pure component density and viscosity, infinite dilution activity coefficients, CO2 solubility, Henry’s law coefficient, and pKa values are carried out over a wide range of temperatures using the respective modules within COSMOtherm. For pKa calculations, the respective protonated structures of the molecules were developed using TURBOMOLE. All of the estimated values have been presented up to three significant digits after the decimal.
Table 1. Specifications of the Chemicals Used in the Present Study.
3. Results and Discussion
The detailed mathematical relationships of various studied parameters with chemical potential are presented in Table 2.40 The conforming significance of the simulated properties has been conversed alongside the analysis results.
Table 2. Detailed Description of the Mathematical Expressions for the Current Work.
| sr. no. | property of the system | mathematical equation | description | |
|---|---|---|---|---|
| 1 | σ profile of the whole system ps(σ) | ![]() |
•“xi” is the mole fraction of component “i” in the mixture | |
| •“pXi(σ)” is the σ profile of any molecule X | ||||
| 2 | σ profile of molecule X | •“ni(σ)” is the number of distributed segments that has surface charge density σ | ||
| •“Ai(σ)” is the segment surface area that has charge density σ | ||||
| •“Ai” is the area of the whole surface cavity rooted in the medium | ||||
| 3 | chemical potential of a surface segment with screening charge density | •“μs(σ)” is the chemical potential of a surface segment | ||
| •“σ” is the polarity of the surface under study | ||||
| 4 | vapor pressure | •“pSi” is the vapor pressure of the component under study | ||
| •“pref” is the vapor pressure of the reference component if any considered | ||||
| •“μgasi” is the chemical potential of the component in the gas phase | ||||
| •“μSi” is the chemical potential of the pure component in mixture S | ||||
| •“R” is the universal gas constant | ||||
| •“T” is the temperature at which the vapor pressure is to be estimated | ||||
| 5 | pure component density | ![]() |
•“MWi” is the molecular weight of the molecule | |
| •“NA” is Avogadro’s number | ||||
| •“V̅i” is the corrected molar liquid volume | ||||
| 6 | corrected molar liquid volume | V̅i = (cHMF × HiMF) + (cHHB × Hi) + (cvcosmo × Vicosmo) + (cM2 × Mi) + (cNring × NiRing) + (∑kcAk × Aik) | •“HiMF” is the pure component misfit interaction enthalpy | |
| •“HiHB” is the hydrogen-bonding enthalpy | ||||
| •“Vicosmo” is the COSMO volume | ||||
| •“Mi2” is the second σ-moment | ||||
| •“NiRing” is the number of ring atoms | ||||
| •“Aik” is the surface area associated with the molecule | ||||
| 7 | pure component viscosity (based on a QSPR) | ln(ηi) = (cArea × Ai) + (cM2 × Mi2) + (cM2 × Mi) + (cNring × NiRing) + (∑kcAk × Aik) | •“Ai” is the surface area | |
| •“Mi2” is the second σ-moment of the compound | ||||
| •“NiRing” is the number of ring atoms in the compound | ||||
| •“TSi” is the pure component entropy at a specific temperature | ||||
| •cArea, cM2, cNRing, cTS, and co are the generic parameters for the QSPR approach for liquid viscosity, with energy values in kcal·mol–1 and areas in Å2 | ||||
| 8 | activity coefficient at infinite dilution | •“γ∞” is the activity coefficient of compound “i” at infinite dilution | ||
| •“μi∞” is the pseudo-chemical potential of “i” at infinite dilution | ||||
| •“μi0” is the chemical potential of “i” in its pure liquid state | ||||
| 9 | COSMO model for VLE | •“Ptot” is the total vapor pressure of the mixture | ||
| •“xi” is the mole fraction of the compounds in the liquid phase | ||||
| •“yi” is the mole fraction of compounds in the gas phase | ||||
| 10 | nonrandom two-liquid (NRTL) model | •“τij” and “τji” are the binary interaction parameters of the associated constituents | ||
| Gij = exp(−αijτij) | •“α = αi j = αji” is the nonrandomness factor that signifies the molecule–molecule or molecule–electrolyte (if any) (default value of 0.3) | |||
| 11 | WILSON model | •“λij,” “λji,” “Vi,” “Vj,” “aji,” and “aij” are binary interaction parameters of the system under study | ||
| 12 | UNIQUAC 4 model | ln(γi) = ln(γiC) + ln(γiR) | •“ln(γiC)” is the combinatorial contribution signifying the entropic size and shape transformations of the compounds | |
![]() | ||||
| the parametric equations are as follows | ||||
![]() |
•“Φi” and “Θi” are the normalized volume and surface area fraction of species “i” in the blended solvent | |||
| •“xi” is the mole fraction | ||||
| the enthalpy interactions among various constituents in the UNIQUAC 4 model are quantified by the residual contribution in the calculation of the activity coefficient; the mentioned term is described as | •“ri” is the volume | |||
| •“qi” is the surface area for each individual species (τji ≠ τij) | ||||
| enthalpy, being closely related to temperature, the residual contribution of binary interaction major parameters “τij” is further taken as an inverse logarithm function of temperature and is given as | ||||
| the compound-specific UNIQUAC volume and surface area parameters are presented as | ||||
| 13 | gas solubility | pj = pj0 × xj × γj | •“pj0” is the vapor pressure of the pure compound | |
| •“γj” is the activity coefficient | ||||
| •“xj” is the mole fraction | ||||
| •“pj” is the partial pressure of compound “j” | ||||
| 14 | Henry’s law coefficient | •“Hjs” is Henry’s law coefficient of compound “j” in solvent “s” | ||
| •“μjs,∞” is the ideal gas-phase chemical potential | ||||
| •“μjgas” is the infinite dilution state chemical potential | ||||
| •“γjs,∞” is the activity coefficient at infinite dilution | ||||
| •“pjP” is the pure compound vapor pressure |
3.1. σ Profile and σ Potential Analysis
The intermolecular interactions of the selected solvents among all of the constituents along with CO2 contribute largely toward CO2 solubility. This further hinges on the associated shape, size, polarity, and type of molecules. Chemical potential “μ” of any species in a solution is evaluated using screening charge density “σ” on the surface of molecules through the COSMO-RS theory within three major norms: (i) the liquid state is incompressible, (ii) all fragments of molecular surfaces can be in interaction with one another, and (iii) individual pairwise interactions of molecular surface areas are permitted. This screening charge density also helps in understanding electrostatic interaction, hydrogen-bonding energy, dispersion, etc. The σ-profile of any molecule is obtained through the weighted sum of the profiles of all of its included components. σ profiles also signify the spreading of screening charge density (σ) of molecules. This distribution is carried out in three categories: (a) nonpolar region (−0.0084 e·Å–2 < σ < 0.0084 e·Å–2), hydrogen-bond donor (HBD) region (σ < −0.0084 e·Å–2), and hydrogen-bond acceptor (HBA) region (σ > 0.0084 e·Å–2).
The σ surfaces of H2O, [bmim] cation, [Ac] anion, CO2, 2-MPZ, MDEA, and TMSO2 are shown in Figure 1 along with the energy associated with each molecule. The zones in the figure can be explained as follows. Red: hydrogen-bond acceptor; blue: hydrogen-bond donor; and green: nonpolar section of the molecule. Single conformers with the least ground-state energy associated with the molecule were considered for the current work.
Figure 1.
σ surface of (a) CO2, (b) H2O, (c) [bmim] cation, (d) [Ac] acetate anion, (e) 2-MPZ, (f) MDEA, and (g) TMSO2.
The σ profile and the corresponding σ potential for the molecules functional to majorly chemical potential are shown in Figure 2. These properties define the attraction of selected solvents with the desired solute, thereby determining the extent of possible separation.
Figure 2.

COSMOtherm generated (a) σ profile and (b) σ potential of MDEA, TMSO2, [bmim][Ac] cation, acetate [Ac] anion, H2O, CO2, and 2-MPZ.
The negative polarities of any molecule are indicated by positive screening charge density in a σ-scale and vice versa.41 The least σ surface was obtained for H2O in the extensive range of −0.02 e·Å–2, and +0.02 e·Å–2 specifies the positive and negative polarities of the associated atoms in the H2O molecule. Successively, it can also be seen from Figure 2a that key portions of σ charge densities of the [bmim] cation, TMSO2, MDEA, and 2-MPZ are negative and those for CO2 and the [Ac] anion are positive in nature. Further, the peak intensities of MDEA and 2-MPZ are very competitive with each other, indicating the possible high CO2 solubility offered by both. Also, if the peaks of TMSO2 are analyzed, it can be perceived to have a positive and negative σ charge density with two sharp peaks. The higher peak is, however, present on the negative side. It can thus be concluded that as CO2 and selected solvents present different charge densities, the selected solvents could provide good CO2 absorption. This conclusion is in agreement with the fact that the negative surface pieces of [bmim], 2-MPZ, TMSO2, and MDEA can react essentially with the positive surface pieces of CO2.42 Although the [bmim] cation is seen to provide the highest negative surface, on combining it with the associated [Ac] cations’ positive surface, it results in overall less polarity available in comparison to MDEA and 2-MPZ.
The intermolecular interaction of a solvent toward the molecular surface that it comes in contact with polar or nonpolar behavior can be qualitatively discussed in terms of the σ potential.43 The positive σ potential of CO2 over the studied charge density of −0.03 to +0.03 e·Å–2 indicates its capability as a H-bond acceptor (Figure 2b). The σ potential behavior of CO2 is almost symmetrical and concave in nature over the entire charge density. On the contrary, the parameter is asymmetrical for [bmim], [Ac], 2-MPZ, MDEA, and TMSO2, leaning more to the negative side of charge density. Additionally, only the [bmim] cation is associated with a positive σ potential in the negative surface charge density region. The molecules signifying a −ve σ potential act as H-bond donors, whereas the +ve σ potential suggests that CO2 is a H-bond acceptor. However, understanding the [bmim] cation alone does not provide any technical application since, in the present study, it is associated with the [Ac] anion. Combining the chemical potentials of both the [bmim] cation and the [Ac] anion leads to the overall negative charge density, proposing it to be a H-bond donor. The formation or loss of a H bond is usually at the S, N, or O atoms of any molecules. At a molecular level, the possible interaction of CO2 with any chosen solvent depends highly on the H-bond acceptor or donor capacity. The interaction strength of CO2 can hence be determined with the order as MDEA > 2-MPZ > [bmim][Ac] > TMSO2 > H2O. At a lab or plant scale, the same concept is understood by the reaction mechanism of zwitterions, proton exchange reactions, and formation or dissolution of bicarbonates, unstable/stable bicarbamates, and dicarbamates.44,45
3.2. Vapor Pressure Analysis
The potential applications of any solvent in diverse fields of chemical engineering depend on many important characteristics such as thermal and mechanical stability, low degradation and toxicity levels, the extent of biodegradation offered, recyclability, vapor pressures, etc. Among these many essential features, vapor pressure is considered to be very important for the CO2 capture process. This is due to the fact that any solvent exhibiting high vapor pressure will lead to huge losses during regeneration. Also, if the vapor pressure is too high, the solubility of CO2 or other acid solute gases decreases at high temperatures. The latter is because, at high temperatures, the diffusivity is expected to increase considerably. On the other hand, if the vapor pressure is too low, e.g., in the case of pure ionic liquids, the diffusion is also too less at low temperatures. Hence, an optimum vapor pressure is desired to achieve both absorption and regeneration cost-effectively. Considering that the initial screening of solvents for any application requires major experimental investigations, the cost-effectiveness can be reduced, provided that the proficient prediction methods for such properties are available. In the same line, many researchers have proposed vapor pressure estimation through traditional or modified thermodynamic equations such as PR-EoS, UNIFAC, UNIFAC-Lei, etc.46−50 The efficacy of any such model depends on the number of assumptions made during calculations, in addition to the number of thermodynamic parameters calculated. For the current work, quantum calculations through COSMO-RS are carried out including the combinatorial and residual parametric contributions of the molecules under study. The vapor pressures are estimated using the boiling points of individual pure constituents of the systems under study as the reference point. The variance in the correlated and experimental values of the parameters under study is obtained through calculation of deviation using the following equation
| 1 |
where N, Yiexp, and Yi indicate the number of data points, experimental value, and modeled or COSMOtherm estimated value of any variable, respectively.
The vapor pressures of aqueous (aq) (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) have been measured using the validated experimental methodology of our previous work.51 The measurements have been carried out at 303.15, 313.15, and 323.15 K and a total solvent concentration of 4.0 mol·kg–1. The activator 2-MPZ concentration is varied from 0.5 to 1.5 mol·kg–1 in the studied solvents. Further, estimation of Antoine equation coefficients is carried out in the temperature range of 298.15–333.15 K. The obtained outcomes are presented in Figure 3 and Tables 3 and 4. The deviations among the experimental and predicted values of vapor pressures for aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) are 16.255, 15.777, and 13.407%, respectively. The obtained deviation through the COSMO-RS theory is quite less when compared to the analytical expressions used for similar estimations.52,53 Nonetheless, the obtained deviations can be attributed to the uncertainties associated with the experimental procedure inclusive of variations in temperature, pressure, and compositions, which have been retained constant throughout the experimentation. Also, while performing quantum calculations, such macroscopic deviations are not considered.
Figure 3.

Residual plot of vapor pressure experimental and COSMO predicted for aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) systems.
Table 3. Comparison of Experimental and COSMO Predicted Vapor Pressure of Aq (MDEA + 2-MPZ), Aq (TMSO2 + 2-MPZ), and Aq ([bmim][Ac] + 2-MPZ) Systems.
| T (K) | 303.15 | 303.15 | 313.15 | 313.15 | 323.15 | 323.15 | ||
|---|---|---|---|---|---|---|---|---|
| system | concentration (mol·kg–1) | experimental | predicted | experimental | predicted | experimental | predicted | % AAD |
| aq (MDEA + 2-MPZ) | (3.509 + 0.509) | 48.9 | 38.412 | 62.1 | 68.191 | 109.6 | 115.879 | 16.255 |
| (3.017 + 1.008) | 47.6 | 38.628 | 61.4 | 68.545 | 98.6 | 116.437 | ||
| (2.502 + 1.509) | 33.1 | 38.856 | 60.7 | 68.917 | 90.3 | 117.023 | ||
| aq (TMSO2 + 2-MPZ) | (3.501 + 0.509) | 56.5 | 38.624 | 65.5 | 68.393 | 111.7 | 115.958 | 15.777 |
| (3.012 + 1.008) | 55.2 | 38.709 | 65.5 | 68.552 | 102.0 | 116.243 | ||
| (2.500 + 1.509) | 53.8 | 38.829 | 62.1 | 68.767 | 101.4 | 116.614 | ||
| aq ([bmim][Ac] + 2-MPZ) | (3.507 + 0.509) | 45.5 | 37.839 | 60.7 | 67.039 | 100.7 | 113.719 | 13.407 |
| (3.002 + 1.008) | 44.1 | 38.074 | 60.7 | 67.454 | 101.4 | 114.423 | ||
| (2.510 + 1.509) | 44.1 | 38.306 | 58.6 | 67.863 | 101.4 | 115.116 |
Table 4. COSMO Predicted Antoine Equation Coefficients in Aq (MDEA + 2-MPZ), Aq (TMSO2 + 2-MPZ), and Aq ([bmim][Ac] + 2-MPZ) Systemsa.
| system | concentration (mol·kg–1) | A | B | C |
|---|---|---|---|---|
| aq (MDEA + 2-MPZ) | (3.509 + 0.509) | 18.101 | 3495.889 | –61.252 |
| (3.017 + 1.008) | 18.106 | 3498.779 | –61.057 | |
| (2.502 + 1.509) | 18.112 | 3501.691 | –60.853 | |
| aq (TMSO2 + 2-MPZ) | (3.501 + 0.509) | 18.072 | 3495.181 | –60.728 |
| (3.012 + 1.008) | 18.086 | 3500.407 | –60.568 | |
| (2.500 + 1.509) | 18.099 | 3504.717 | –60.426 | |
| aq ([bmim][Ac] + 2-MPZ) | (3.507 + 0.509) | 18.065 | 3498.714 | –60.718 |
| (3.002 + 1.008) | 18.083 | 3504.428 | –60.511 | |
| (2.510 + 1.509) | 18.097 | 3508.562 | –60.358 |
(P is in millibar and T is in
kelvin).
3.3. Estimation of Pure Component Density and Viscosity
The experimental analysis of pure component density and viscosity has been reported by many researchers for different purposes in carbon capture systems, mainly for estimation of kinetic parameters and pumping costs,54,55 concise designing of absorption–stripper columns,56 understanding the nonideal behavior through analysis of viscosity deviation or excess molar properties,57 etc. In the present work, pure component density and viscosity are estimated through the quantitative structure–property relationship (QSPR) approach, which consists of many inherent properties of the involved molecules (Table 2). The pure component density and viscosity measurements have been carried out using a density and sound velocity meter (DSA 5000 M, Anton Paar, Austria) and an Anton Paar AMVn rolling ball viscometer with the standard uncertainties of u(T) = 0.01 K, u(P) = 0.2 kPa, u(ρ) = 0.5 kg·m–3, and u(η) = 0.07 mPa·s. The adopted detailed methodology can be referred to from our previous work.51
A comparison of the experimental and COSMO estimated density and viscosity of the involved chemical species is presented in Tables 5 and 6. For the case of 2-MPZ, since it is in crystalline form, the experimental measurement was difficult and hence viscosity was estimated using Aspen plus58 and COSMOtherm and a comparison of the two has been presented. The least deviations observed are with respect to the viscosity of TMSO2 and density of [bmim][Ac], i.e., 0.707 × 102 and 1.067%, respectively. A major reason for this huge deviation, especially with respect to viscosity, is owing to the fact that the structural relationship utilized for the prediction of pure component properties, i.e., QSPR, is assumed to be independent of temperature, which is usually not due to non-Newtonian and nonideal liquid systems. However, similar observations are also reported elsewhere.59 As a matter of fact, it can be concluded from the present comparison of experimental and COSMO-computed values of density and viscosity that the estimation is not accurate for both the properties. Further, the prediction of pure component density and viscosity does not get reflected in other evaluations for the reason that, as indicated in Table 2, the other properties are calculated as a function of chemical potential. Hence, effective prediction of chemical potential results in accurate predictions of the properties under consideration such as vapor pressure, except pure component density and viscosity.
Table 5. Comparison of Experimental and COSMO Predicted Density (ρ, kg·m–3) of Pure MDEA, TMSO2, 2-MPZ, and [bmim][Ac].
| system |
||||||||
|---|---|---|---|---|---|---|---|---|
| MDEA |
2-MPZ |
TMSO2 |
[bmim][Ac] |
|||||
| data | ρexp | ρpred | ρexp | ρpred | ρexp | ρpred | ρexp | ρpred |
| T (K) | ||||||||
| 298.15 | 1036.8 | 984.816 | 875.5 | 995.703 | 1013.3 | 1346.294 | 1052.1 | 1067.123 |
| 303.15 | 1032.9 | 979.811 | 872.7 | 991.028 | 1008.1 | 1340.947 | 1049.1 | 1063.032 |
| 308.15 | 1029.2 | 974.833 | 869.8 | 986.363 | 1004.3 | 1335.579 | 1045.9 | 1058.923 |
| 313.15 | 1025.4 | 969.882 | 866.9 | 981.711 | 1000.4 | 1330.191 | 1043.0 | 1054.796 |
| 318.15 | 1020.9 | 964.957 | 863.9 | 977.069 | 996 | 1324.783 | 1040.0 | 1050.652 |
| 323.15 | 1017.7 | 960.059 | 860.9 | 972.441 | 992.8 | 1319.356 | 1037.0 | 1046.493 |
| 328.15 | 1012.9 | 955.187 | 857.8 | 967.824 | 988.4 | 1313.911 | 1034.1 | 1042.318 |
| 333.15 | 1009.1 | 950.341 | 854.7 | 963.220 | 985.1 | 1308.450 | 1031.1 | 1038.129 |
| % AAD | 5.439 | 13.189 | 32.936 | 1.067 | ||||
Table 6. Comparison of Experimental and COSMO Predicted Viscosity (η, mPa·s) of Pure MDEA, 2-MPZ, and TMSO2.
| system |
||||||
|---|---|---|---|---|---|---|
| MDEA |
2-MPZ |
TMSO2 |
||||
| data | ηexp | ηpred | ηpred,aspen | ηpred,COSMO | ηexp | ηpred |
| T (K) | ||||||
| 298.15 | 77.75 | 3.839 | 0.47 | 4.462 | 10.28 | 2.917 |
| 303.15 | 56.26 | 3.459 | 0.46 | 4.005 | 10.22 | 2.648 |
| 308.15 | 46.47 | 3.127 | 0.45 | 3.607 | 9.06 | 2.411 |
| 313.15 | 34.66 | 2.836 | 0.43 | 3.259 | 7.84 | 2.202 |
| 318.15 | 28.67 | 2.580 | 0.42 | 2.954 | 6.58 | 2.016 |
| 323.15 | 23.29 | 2.354 | 0.41 | 2.686 | 6.18 | 1.851 |
| 328.15 | 17.42 | 2.154 | 0.40 | 2.449 | 5.19 | 1.705 |
| 333.15 | 14.55 | 1.976 | 0.39 | 2.239 | 4.88 | 1.573 |
| % 10–2 × AAD | 0.911 | 6.405 | 0.707 | |||
3.4. Infinite Dilution Activity Coefficients
The magnitude of nonideality in liquid solutions is often in terms of infinite dilution activity coefficients (IDACs). Since the solvents associated with absorption of CO2 and consequently the same solvent for desorption while releasing CO2 from it are expected to show nonideality behavior, such nonidealities are also variously expressed in terms of property deviations such as density and viscosity deviations from ideality. A few important thermodynamic properties, for instance, partition coefficient, separation factors, and Henry’s law coefficient, are also dictated by activity coefficients. However, the experimental calculation of IDACs is difficult and expensive. Hence, a preliminary analysis through the apposite prediction method is of huge interest.60−62 The estimation of IDACs of aqueous, nonaqueous, organic, or ionic liquid systems, through COSMO-RS, is reported in the literature to be very efficient owing to the reason the same being evaluated in the absenteeism of mean-field approximation.63 The nonpolarity or active polarity of a compound is decided by the higher or lower value of IDACs in the system (Table 2). Table 7 presents the COSMO predicted infinite dilution activity coefficients of MDEA, 2-MPZ, and TMSO2 in H2O as a function of temperature in the range of 298.15–333.15 K. With an increase in the temperature, IDACs of MDEA, TMSO2, and 2-MPZ in water were found to increase. Although the IDACs values for MDEA and 2-MPZ are very low in comparison to the TMSO2 values, for the currently studied compounds, all of the IDACs are greater or less than unity, indicating the highly nonideal behavior of such systems.43
Table 7. COSMO Predicted Activity Coefficients of MDEA, TMSO2, and 2-MPZ at Infinite Dilution in Water.
| T (K) | 298.15 | 303.15 | 308.15 | 313.15 | 318.15 | 323.15 | 328.15 | 333.15 | |
|---|---|---|---|---|---|---|---|---|---|
| γ∞ | MDEA | 0.671 | 0.823 | 0.997 | 1.193 | 1.411 | 1.650 | 1.909 | 2.186 |
| TMSO2 | 4.255 | 4.493 | 4.719 | 4.931 | 5.126 | 5.304 | 5.462 | 5.601 | |
| 2-MPZ | 0.246 | 0.305 | 0.375 | 0.455 | 0.549 | 0.655 | 0.774 | 0.908 |
3.5. Vapor–Liquid Equilibrium Relationship of Aq (MDEA + 2-MPZ), Aq (TMSO2 + 2-MPZ), and Aq ([bmim][Ac] + 2-MPZ)
At a specific temperature and pressure, the extent of reaction resulting in formation of products or scattering of various constituents is dependent on the chemical potential and subsequently on the Gibbs’ free energy. Further, the maximum gas solubility at equilibrium is also dependent on the values of the activity coefficients of the solvent. The chemical potential is also considered a sum total of various energies as well as factors affecting the energies such as internal, density, temperature, enthalpy, etc., of any molecule. This indicates the dependency of acid–gas separations at a molecular level on the thermodynamic properties such as chemical potential, enthalpy, Gibb’s free energy, etc.
Henceforth, the vapor–liquid equilibrium of the ternary mixtures of aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) related to the nonlinear behavior of the various constituents involved is determined using the COSMO-RS theory. The VLE estimated through COSMO is based on the vapor pressure and activity coefficients of individual pure constituents in the mixture (Table 2). The VLE is additionally modeled using NRTL, WILSON, and UNIQUAC 4 models. The obtained VLE data is presented in the form of activity coefficients, excess Gibbs free energy, excess enthalpy, chemical potential, and individual partial pressures associated with each system at 303.15, 313.15, and 323.15 K. The total pressures of the aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) systems are taken to be 0–124 mbar. The difference between the COSMO and NRTL/WILSON/UNIQUAC 4 model predicted activity coefficients is calculated in terms of the root-mean-square deviation (RMSD) using the following equation
| 2 |
where yCOSMO is the COSMO predicted property and yi is the NRTL, WILSON, or UNIQUAC predicted property value.
The activity coefficients and binary interaction parameters for NRTL, WILSON, and UNIQUAC 4 models of the systems under study were estimated and are presented in Tables 8, 9, and 10, respectively. The majority of the binary interaction parameters of the studied systems are found to be different because the systems are asymmetric, i.e., τji ≠ τij.64,65
Table 8. COSMO Predicted NRTL Parameters for the Activity Coefficients in (H2O (1) + MDEA (2) + 2-MPZ (3)), (H2O (1) + TMSO2 (2) + 2-MPZ (3)), and (H2O (1) + [bmim][Ac] (2) + 2-MPZ (3)) Systems.
| system | aq (MDEA + 2-MPZ) |
aq (TMSO2 + 2-MPZ) |
aq ([bmim][Ac] + 2-MPZ) |
||||||
|---|---|---|---|---|---|---|---|---|---|
| T (K) | 303.15 | 313.15 | 323.15 | 303.15 | 313.15 | 323.15 | 303.15 | 313.15 | 323.15 |
| A | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 |
| τ12 | –0.952 | –0.863 | –0.753 | 0.376 | 0.456 | 1.336 | –1.918 | –1.839 | –1.760 |
| τ13 | 0.803 | 0.975 | 1.228 | 0.920 | 1.207 | 5.479 | 0.055 | 0.099 | 0.151 |
| τ21 | 1.999 | 1.900 | 1.761 | 1.580 | 1.388 | 0.533 | –3.167 | –3.053 | –2.943 |
| τ23 | 1.231 | 1.150 | 1.052 | 0.740 | 0.643 | 0.663 | 1.484 | 1.381 | 1.289 |
| τ31 | –2.167 | –2.108 | –2.077 | –2.267 | –2.229 | –2.727 | –2.052 | –1.895 | –1.746 |
| τ32 | 0.016 | –0.018 | –0.029 | 1.686 | 1.600 | 1.287 | 1.416 | 1.393 | 1.375 |
| RMSD | 0.329 | 0.309 | 0.291 | 0.394 | 0.371 | 0.331 | 0.531 | 0.493 | 0.459 |
Table 9. COSMO Predicted WILSON Parameters for the Activity Coefficients in (H2O (1) + MDEA (2) + 2-MPZ (3)), (H2O (1) + TMSO2 (2) + 2-MPZ (3)), and (H2O (1) + [bmim][Ac] (2) + 2-MPZ (3)) Systems.
| system | aq (MDEA + 2-MPZ) |
aq (TMSO2 + 2-MPZ) |
aq ([bmim][Ac] + 2-MPZ) |
||||||
|---|---|---|---|---|---|---|---|---|---|
| T (K) | 303.15 | 313.15 | 323.15 | 303.15 | 313.15 | 323.15 | 303.15 | 313.15 | 323.15 |
| λ12 | 0.223 | 0.221 | 0.224 | 0.210 | 0.232 | 0.258 | 11.673 | 10.800 | 10.011 |
| λ13 | 4.132 | 3.878 | 3.658 | 4.199 | 3.944 | 3.719 | 4.721 | 4.251 | 3.838 |
| λ21 | 1.817 | 1.731 | 1.641 | 0.518 | 0.525 | 0.528 | 5.144 | 4.904 | 4.675 |
| λ23 | 0.865 | 0.918 | 0.964 | 0.223 | 0.250 | 0.276 | 0.376 | 0.381 | 0.384 |
| λ31 | 1.209 | 1.103 | 0.994 | 1.259 | 1.126 | 0.994 | 1.247 | 1.173 | 1.098 |
| λ32 | 0.315 | 0.329 | 0.347 | 0.275 | 0.319 | 0.361 | 0.087 | 0.099 | 0.112 |
| a12 | –0.199 | –0.201 | –0.215 | –0.061 | –0.125 | –0.196 | –2.487 | –2.521 | –2.553 |
| a13 | –1.878 | –1.901 | –1.924 | –1.888 | –1.911 | –1.934 | –1.958 | –1.958 | –1.955 |
| a21 | 0.744 | 0.798 | 0.858 | 1.397 | 1.434 | 1.476 | 0.020 | 0.051 | 0.083 |
| a23 | 0.167 | 0.136 | 0.109 | 0.882 | 0.838 | 0.801 | 0.573 | 0.584 | 0.596 |
| a31 | 0.909 | 0.996 | 1.095 | 0.885 | 0.983 | 1.095 | 0.890 | 0.958 | 1.031 |
| a32 | 0.617 | 0.608 | 0.595 | 0.802 | 0.735 | 0.679 | 1.487 | 1.456 | 1.424 |
| v1 | 25.855 | 25.855 | 25.855 | 25.855 | 25.855 | 25.855 | 25.855 | 25.855 | 25.855 |
| v2 | 161.363 | 161.363 | 161.363 | 136.050 | 136.050 | 136.050 | 137.507 | 137.507 | 137.507 |
| v3 | 141.349 | 141.349 | 141.349 | 141.350 | 141.349 | 141.349 | 141.349 | 141.349 | 141.349 |
| RMSD | 0.346 | 0.320 | 0.299 | 0.404 | 0.378 | 0.355 | 0.749 | 0.687 | 0.630 |
Table 10. COSMO Predicted UNIQUAC 4 Parameters for the Activity Coefficients in (H2O (1) + MDEA (2) + 2-MPZ (3)), (H2O (1) + TMSO2 (2) + 2-MPZ (3)), and (H2O (1) + [bmim][Ac] (2) + 2-MPZ (3)) Systems.
| systems | aq (MDEA + 2-MPZ) |
aq (TMSO2 + 2-MPZ) |
aq ([bmim][Ac] + 2-MPZ) |
||||||
|---|---|---|---|---|---|---|---|---|---|
| T (K) | 303.15 | 313.15 | 323.15 | 303.15 | 313.15 | 323.15 | 303.15 | 313.15 | 323.15 |
| q1 | 1.653 | 1.622 | 1.579 | 2.223 | 2.014 | 1.838 | 7.058 | 6.572 | 6.086 |
| q2 | 1.692 | 1.705 | 1.707 | 0.996 | 0.983 | 0.964 | 19.599 | 18.544 | 17.453 |
| q3 | 9.946 | 9.665 | 9.349 | 11.529 | 10.621 | 9.920 | 30.774 | 28.501 | 26.338 |
| r1 | 1.086 | 1.066 | 1.038 | 1.461 | 1.324 | 1.208 | 5.639 | 5.251 | 4.863 |
| r2 | 2.232 | 2.249 | 2.253 | 1.279 | 1.262 | 1.237 | 24.320 | 23.012 | 21.658 |
| r3 | 12.801 | 12.439 | 12.034 | 14.839 | 13.671 | 12.768 | 39.608 | 36.684 | 33.899 |
| τ12 | 0.108 | 0.101 | 0.100 | 0.117 | 0.127 | 0.136 | 2.240 | 2.278 | 2.322 |
| τ13 | 2.241 | 2.132 | 2.026 | 1.729 | 1.689 | 1.624 | 1.447 | 1.443 | 1.438 |
| τ21 | 4.717 | 4.540 | 4.341 | 2.823 | 2.724 | 2.625 | 1.012 | 0.984 | 0.955 |
| τ23 | 1.454 | 1.447 | 1.439 | 1.087 | 1.084 | 1.072 | 1.466 | 1.449 | 1.433 |
| τ31 | 0.849 | 0.849 | 0.845 | 1.394 | 1.327 | 1.271 | 1.183 | 1.166 | 1.148 |
| τ32 | 0.245 | 0.261 | 0.275 | 0.025 | 0.030 | 0.034 | 0.549 | 0.557 | 0.565 |
| u12 | 1.341 | 1.426 | 1.478 | 1.295 | 1.286 | 1.282 | –0.486 | –0.512 | –0.541 |
| u13 | –0.486 | –0.471 | –0.453 | –0.329 | –0.326 | –0.311 | –0.223 | –0.228 | –0.233 |
| u21 | –0.935 | –0.942 | –0.943 | –0.625 | –0.624 | –0.619 | –0.007 | 0.010 | 0.029 |
| u23 | –0.225 | –0.230 | –0.234 | –0.050 | –0.050 | –0.045 | –0.230 | –0.231 | –0.231 |
| u31 | 0.098 | 0.102 | 0.109 | –0.200 | –0.176 | –0.154 | –0.101 | –0.096 | –0.089 |
| u32 | 0.848 | 0.836 | 0.829 | 2.213 | 2.181 | 2.175 | 0.362 | 0.364 | 0.367 |
| a12 | 675.017 | 717.561 | 743.797 | 651.785 | 646.952 | 645.321 | –244.517 | –257.853 | –272.239 |
| a13 | –244.652 | –237.117 | –228.163 | –166.034 | –164.243 | –156.605 | –112.064 | –114.933 | –117.345 |
| a21 | –470.235 | –473.772 | –474.405 | –314.656 | –313.838 | –311.885 | –3.667 | 5.039 | 14.787 |
| a23 | –113.390 | –115.781 | –117.811 | –25.154 | –25.191 | –22.549 | –115.886 | –116.072 | –116.211 |
| a31 | 49.453 | 51.136 | 54.603 | –100.711 | –88.555 | –77.544 | –50.919 | –48.067 | –44.636 |
| a32 | 426.718 | 420.757 | 416.955 | 1113.699 | 1097.438 | 1094.228 | 182.039 | 183.115 | 184.581 |
| RMSD | 0.279 | 0.262 | 0.247 | 0.325 | 0.309 | 0.295 | 0.375 | 0.353 | 0.333 |
The obtained results are presented graphically for aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) in Figures 4, 5, and 6, respectively. The mole fraction of water is 0.7 for the presented data. The activity coefficients of 2-MPZ are found to be very less when compared to MDEA and H2O (Figure 4a). The temperature dependency of the parameter is observed not to be high. Also, as a function of the MDEA concentration, the activity coefficients of 2-MPZ were found to show an inverse relationship. However, for both MDEA and H2O, it does not change much. Insignificant deviations of HE (excess enthalpy) and GE (excess Gibbs free energy) with respect to the temperature change from 303.15 to 323.15 K are observed (Figure 4b). Nevertheless, both the properties were found to have negative values that increase as a function of the MDEA concentration. The −ve GE indicates the spontaneous mixing because of the thermodynamic driving forces between the involved components and −ve HE indicates an exothermic reaction/mixing in both the systems that are proven by the values of heat of absorption.66,67 The chemical potential, on the other hand, is observed to be more temperature-dependent. The values of μH2O remain almost unchanged. Decreasing the concentration of 2-MPZ results in high chemical potential of MDEA and vice versa (Figure 4c).
Figure 4.

COSMO predicted (a) activity coefficient, (b) excess enthalpy and Gibb’s free energy, and (c) chemical potential as a function of mole fraction of MDEA for the aq (MDEA + 2-MPZ) system at xH2O = 0.7.
Figure 5.

COSMO predicted (a) activity coefficient, (b) excess enthalpy and Gibb’s free energy, and (c) chemical potential as a function of mole fraction of TMSO2 for the aq (TMSO2 + 2-MPZ) system at xH2O = 0.7.
Figure 6.

COSMO predicted (a) activity coefficient, (b) excess enthalpy and Gibb’s free energy, and (c) chemical potential as a function of mole fraction of [bmim][Ac] for the aq ([bmim][Ac] + 2-MPZ) system at xH2O = 0.7.
Similarly, with respect to the TMSO2 concentration, the activity coefficients of TMSO2 and H2O are observed to decrease and increase slightly simultaneously as a function of TMSO2 (Figure 5a). The behavior of 2-MPZ is similar to that found for the aq (MDEA + 2-MPZ) system. The behavior of all three systems with respect to chemical potential is found to be very similar (Figures 5c and 6c). The activity coefficients for [bmim][Ac] are very less when compared to 2-MPZ and H2O at any given concentration (Figure 6a). This is also expected due to the lesser chemisorption but greater physisorption behavior conferred by the ionic liquid to the acid gas. Also, though the HE and GE values obtained in the aq ([bmim][Ac] + 2-MPZ) system are negative but are a function of [bmim][Ac], the values are found to further decrease unlike those for the other two systems (Figure 6b).
3.6. CO2 Solubility in Aq (MDEA + 2-MPZ), Aq (TMSO2 + 2-MPZ), and Aq ([bmim][Ac] + 2-MPZ)
The experimental measurement at the lab scale or estimation of CO2 solubility in suitable solvents has been a key research area for application.29,34 Considering the lower partial pressures of CO2 in the flue gas stream, the CO2 solubility has been predicted using the activity coefficients estimated through COSMO-RS over the temperature and pressure range of 303.15–323.15 K and 1.0–3.0 bar for aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) systems at varying compositions. The solubilities are presented in terms of the mole fraction of CO2 in the liquid phase (xCO2). An inverse relationship between the CO2 solubility and temperature for all of the studied solvent blends was observed (Figure 7). Further, although increasing the CO2 pressure as well as the activator 2-MPZ concentration in all solvents resulted in an increase in CO2 solubility (Table 11), however, in the case of the aq (MDEA + 2-MPZ) system, CO2 solubility is observed to be almost similar over the chosen compositional range. This may be attributed to the fact that MDEA is a tertiary amine when compared with TMSO2 and [bmim][Ac], which have higher CO2 solubility. Hence, with the increasing concentration of 2-MPZ in the aqueous blends of (MDEA + 2-MPZ), the concentration of MDEA is also simultaneously decreased. Hence, the CO2 solubility at high MDEA concentration can be understood to be compensated by a decrease in the 2-MPZ concentration. Decisively, the highest CO2 solubility is observed for the aq (MDEA + 2-MPZ) concentration at 303.15 K. However, it should be also considered that there may be a deviation when the same solvents are studied experimentally for CO2 absorption. This discrepancy may be attributed to variables affecting the process, nonideality associated with gas and liquid phases, vapor pressures, temperature, maintenance of the partial pressure in the system, etc.
Figure 7.

COSMO predicted CO2 solubility of aq (3.509 m MDEA + 0.509 2-MPZ), aq (3.501 m TMSO2 + 0.509 m 2-MPZ), and aq (3.507 m [bmim][Ac] + 0.509 m 2-MPZ) as a function of temperature and pressure.
Table 11. COSMO Predicted CO2 Solubility (100 × x) from 1.0 to 3.0 Bar Pressure in the Temperature Range of 303.15–323.15 Ka.
|
T (K) |
T (K) |
T (K) |
||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| (MDEA + 2-MPZ) (mol·kg–1) | P (bar) | 303.15 | 313.15 | 323.15 | (TMSO2 + 2-MPZ) (mol·kg–1) | 303.15 | 313.15 | 323.15 | ([bmim][Ac] + 2-MPZ) (mol·kg–1) | 303.15 | 313.15 | 323.15 |
| (3.509 + 0.509) | 1.0 | 0.069 | 0.058 | 0.049 | (3.501 + 0.509) | 0.054 | 0.045 | 0.039 | (3.507 + 0.509) | 0.054 | 0.045 | 0.039 |
| 1.5 | 0.103 | 0.086 | 0.074 | 0.081 | 0.068 | 0.058 | 0.081 | 0.068 | 0.058 | |||
| 2.0 | 0.137 | 0.115 | 0.099 | 0.108 | 0.091 | 0.078 | 0.108 | 0.091 | 0.078 | |||
| 2.5 | 0.171 | 0.144 | 0.123 | 0.135 | 0.113 | 0.097 | 0.135 | 0.113 | 0.097 | |||
| 3.0 | 0.205 | 0.172 | 0.148 | 0.162 | 0.136 | 0.116 | 0.161 | 0.136 | 0.116 | |||
| (3.017 + 1.008) | 1.0 | 0.069 | 0.058 | 0.049 | (3.012 + 1.008) | 0.056 | 0.047 | 0.040 | (3.002 + 1.008) | 0.056 | 0.047 | 0.040 |
| 1.5 | 0.103 | 0.087 | 0.074 | 0.084 | 0.071 | 0.061 | 0.084 | 0.070 | 0.060 | |||
| 2.0 | 0.137 | 0.115 | 0.099 | 0.112 | 0.094 | 0.081 | 0.111 | 0.094 | 0.080 | |||
| 2.5 | 0.171 | 0.144 | 0.123 | 0.141 | 0.118 | 0.101 | 0.139 | 0.117 | 0.100 | |||
| 3.0 | 0.205 | 0.173 | 0.148 | 0.169 | 0.142 | 0.121 | 0.167 | 0.140 | 0.120 | |||
| (2.502 + 1.509) | 1.0 | 0.069 | 0.058 | 0.049 | (2.500 + 1.509) | 0.058 | 0.049 | 0.042 | (2.510 + 1.509) | 0.058 | 0.049 | 0.042 |
| 1.5 | 0.103 | 0.086 | 0.074 | 0.088 | 0.074 | 0.063 | 0.087 | 0.073 | 0.063 | |||
| 2.0 | 0.137 | 0.115 | 0.099 | 0.117 | 0.098 | 0.084 | 0.116 | 0.097 | 0.083 | |||
| 2.5 | 0.171 | 0.144 | 0.123 | 0.146 | 0.122 | 0.105 | 0.144 | 0.121 | 0.104 | |||
| 3.0 | 0.205 | 0.173 | 0.148 | 0.175 | 0.147 | 0.126 | 0.173 | 0.146 | 0.125 | |||
x is the mole fraction of CO2 in the loaded solvent.
3.7. Estimation of Henry’s Constant in CO2 and N2O and Free Energy of Solvation
Henry’s constant signifies the physical solubility conferred by any solvent selectively to a gas that can either be measured experimentally or through the mathematical expressions available in the literature.68HCO2 or HN2O is a contributive property through misfit in the intermolecular interactions, hydrogen bonding, and van der Waals forces of attraction. The estimation of Henry’s law coefficient involves the estimation of solvation free energies, which are further related to the chemical potential associated with the gas and solvent at a specified temperature and pressure (Table 2). The calculated results are reported in Figure 8 and Table 12 for all of the systems under study. A low value of Henry’s law coefficient indicates a higher CO2 solubility. It can thus be inferred from the obtained results that increasing the concentration of 2-MPZ from ≈0.5 to ≈1.5 mol·kg–1, for all base solvents of TMSO2 and [bmim][Ac], yields an increase of physical solubility, whereas the maximum physical solubility in the case of aq (MDEA + 2-MPZ) is obtained at a 1.008 mol·kg–1 concentration of 2-MPZ. Increasing the 2-MPZ concentration further in the aq (MDEA + 2-MPZ) system results in a decrease in physical solubility that may be owing to the fact that the number of amino groups in the overall blend has increased, which majorly contributes to chemical solubility. Further, the solubilities are found to be much higher at low temperatures for all systems. Henry’s law coefficients for CO2 and N2O are obtained in the order: aq (MDEA + 2-MPZ) > aq (TMSO2 + 2-MPZ) > aq ([bmim][Ac] + 2-MPZ) systems.
Figure 8.

COSMO predicted Henry’s constant of CO2 and N2O in (a) aq (MDEA + 2-MPZ), (b) aq (TMSO2 + 2-MPZ), and (c) aq ([bmim][Ac] + 2-MPZ) systems as a function of composition and temperature.
Table 12. COSMO Predicted Henry’s Constant (H, bar) and Gibbs’ Free Energy of Solvation (ΔGs, kcal·mol–1) in the Temperature Range of 303.15–323.15 Ka.
|
T (K) |
T (K) |
T (K) |
|||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| aq (MDEA + 2-MPZ) (mol·kg–1) | 303.15 | 313.15 | 323.15 | aq (TMSO2 + 2-MPZ) (mol·kg–1) | 303.15 | 313.15 | 323.15 | aq ([bmim][Ac] + 2-MPZ) (mol·kg–1) | 303.15 | 313.15 | 323.15 | ||
| CO2 | H | (3.509 + 0.509) | 1100.644 | 1238.922 | 1371.935 | (3.501 + 0.509) | 1396.815 | 1572.858 | 1741.921 | (3.507 + 0.509) | 1400.345 | 1575.528 | 1744.370 |
| ΔGs | 4.219 | 4.432 | 4.639 | 4.363 | 4.580 | 4.792 | 4.364 | 4.582 | 4.793 | ||||
| N2O | H | 925.420 | 1057.576 | 1186.703 | 1137.946 | 1303.575 | 1465.748 | 1257.580 | 1432.083 | 1602.166 | |||
| ΔGs | 4.115 | 4.333 | 4.546 | 4.239 | 4.464 | 4.681 | 4.299 | 4.522 | 4.739 | ||||
| CO2 | H | (3.017 + 1.008) | 1097.466 | 1235.339 | 1368.044 | (3.012 + 1.008) | 1341.653 | 1510.724 | 1673.258 | (3.002 + 1.008) | 1353.436 | 1522.609 | 1685.678 |
| ΔGs | 4.217 | 4.430 | 4.637 | 4.338 | 4.555 | 4.766 | 4.344 | 4.560 | 4.771 | ||||
| N2O | H | 918.970 | 1050.473 | 1179.077 | 1093.273 | 1252.391 | 1408.311 | 1197.891 | 1365.014 | 1528.088 | |||
| ΔGs | 4.110 | 4.329 | 4.542 | 4.215 | 4.439 | 4.656 | 4.270 | 4.492 | 4.708 | ||||
| CO2 | H | (2.502 + 1.509) | 1099.129 | 1237.204 | 1370.175 | (2.500 + 1.509) | 1293.996 | 1457.038 | 1613.922 | (2.510 + 1.509) | 1304.265 | 1467.245 | 1624.386 |
| ΔGs | 4.218 | 4.431 | 4.638 | 4.317 | 4.533 | 4.743 | 4.321 | 4.537 | 4.747 | ||||
| N2O | H | 916.253 | 1047.660 | 1176.281 | 1054.550 | 1208.024 | 1358.521 | 1138.477 | 1298.193 | 1454.226 | |||
| ΔGs | 4.109 | 4.328 | 4.540 | 4.193 | 4.416 | 4.633 | 4.240 | 4.461 | 4.676 | ||||
1 bar of gas per 1 mol of solvent is taken as the reference state for the H and ΔG calculation.
3.8. Estimation of the Dissociation Constant (pKa) of MDEA and 2-MPZ
Proton exchange is considered to be one of the major reactions that occur during the interaction between CO2 and amines.2,69 This proton exchange reaction rate constant is often described using the dissociation constant (pKa) of the reaction. The −ve log of pKa of conjugate acid reveals the extent of basicity offered by the chosen solvent. The above property can be either measured experimentally using the acid–base titration method or can also be predicted using quantum methods. The dissociation constants of 2-MPZ and MDEA have been calculated using COSMO. Initially, single conformers with ground-state energy were selected. The selected geometry was then edited to create a cationic structure of the same. Further, this cationic structure was optimized to calculate the energy using Turbomole software. The generated cationic structures along with the optimized energy are presented in Figure 9. Based on the free-energy change in any molecule and corresponding cationic structure, the pKa values were estimated in water, acetonitrile, and tetrahydrofuran solvents. The predicted pKa values of MDEA and 2-MPZ are given in Table 13 at 25 °C. Simulated results show that the pKa values of 2-MPZ are relatively higher compared with MDEA in all solvents, demonstrating it to have the possibility of enhanced CO2 solubility.
Figure 9.

Turbomole optimized generated (a) N-methyldiethanolamine cation and (b) 2-methylpiperazine cation.
Table 13. pKa Values of MDEA and 2-MPZ at 298.15 K in Various Solvents.
| solvent | H2O | acetonitrile | tetrahydrofuran | |
|---|---|---|---|---|
| pKa | MDEA | 3.758 | 12.176 | 8.836 |
| 2-MPZ | 5.885 | 14.333 | 10.387 |
4. Conclusions
A comprehensive thermodynamic analysis through the COSMO-RS theory has been carried out for the proposed enhanced CO2 solubility by 2-methylpiperazine in aqueous solvents of N-methyldiethanolamine, sulfolane, and 1-butyl-2-methyl-imidazolium acetate. The appositeness of chosen solvents, i.e., MDEA, TMSO2, and [bmim][Ac], and their blends with 2-MPZ is determined using σ potential and σ surface analyses. The results indicate the suitability of 2-MPZ for a variety of solvents under study. Conclusively, the aqueous blend of MDEA and 2-MPZ is preferred over other solvents. Vapor pressures of the solvents, pure component density, and viscosity are also simulated and compared with experimental data, indicating a featured calculation through the involved method. The minimum deviations for estimation of vapor pressure and density for [bmim][Ac] with % AAD are 13.407 and 1.067, respectively. Additionally, activity coefficient-based thermodynamic models, viz., NRTL, WILSON, and UNIQUAC 4 model parameters, are assessed for aq (MDEA + 2-MPZ), aq (TMSO2 + 2-MPZ), and aq ([bmim][Ac] + 2-MPZ) systems. The nonideal behavior of the systems is indicated by the simulated values of infinite dilution activity coefficients, activity coefficient, excess enthalpy, excess Gibbs’ free energy, and chemical potential. CO2 solubility in all of the solvents is predicted at 303.15–323.15 K and 1.0–3.0 bar pressure. Moreover, COSMO anticipated pKa values of MDEA and 2-MPZ indicate a higher pKa value for 2-MPZ than MDEA. This further confirms the expected higher CO2 solubility in 2-MPZ when compared with MDEA.
Acknowledgments
The authors thank Pandit Deendayal Petroleum University, Gandhinagar, India, for providing the necessary computational and experimentation facilities.
The authors declare no competing financial interest.
References
- Shirazizadeh H. A.; Haghtalab A. Measurement and modeling of CO2 solubility in binary aqueous DMSO and MDEA and their ternary mixtures at different temperatures and compositions. Fluid Phase Equilib. 2020, 112845 10.1016/j.fluid.2020.112845. [DOI] [Google Scholar]
- Afkhamipour M.; Mofarahi M.; Rezaei A.; Mahmoodi R.; Lee C. H. Experimental and theoretical investigation of equilibrium absorption performance of CO2 using a mixed 1-dimethylamino-2-propanol (1DMA2P) and monoethanolamine (MEA) solution. Fuel 2019, 256, 115877 10.1016/j.fuel.2019.115877. [DOI] [Google Scholar]
- Zong L.; Chen C. C. Thermodynamic modeling of CO2 and H2S solubilities in aqueous DIPA solution, aqueous sulfolane-DIPA solution, and aqueous sulfolane-MDEA solution with electrolyte NRTL model. Fluid Phase Equilib. 2011, 306, 190–203. 10.1016/j.fluid.2011.04.007. [DOI] [Google Scholar]
- Chen X.; Rochelle G. T. Modeling of CO2 absorption kinetics in aqueous 2-methylpiperazine. Ind. Eng. Chem. Res. 2013, 52, 4239–4248. 10.1021/ie3023737. [DOI] [Google Scholar]
- Khan S. N.; Hailegiorgis S. M.; Man Z.; Shariff A. M.; Garg S. Thermophysical properties of aqueous 1-butyl-3-methylimidazolium acetate [bmim] [Ac] + monoethanolamine (MEA) hybrid as a solvent for CO2 capture. Procedia Eng. 2016, 148, 1326–1331. 10.1016/j.proeng.2016.06.553. [DOI] [Google Scholar]
- Taboada M. E.; Véliz D. M.; Galleguillos H. R.; Graber T. A. Solubility, density, viscosity, electrical conductivity, and refractive index of saturated solutions of lithium hydroxide in water + ethanol. J. Chem. Eng. Data 2005, 50, 187–190. 10.1021/je0497449. [DOI] [Google Scholar]
- Sanku M. G.; Svensson H. Modelling the precipitating non-aqueous CO2 capture system AMP-NMP, using the unsymmetric electrolyte NRTL. Int. J. Greenhouse Gas Control 2019, 89, 20–32. 10.1016/j.ijggc.2019.07.006. [DOI] [Google Scholar]
- Mota Martinez M. T.; Kroon M. C.; Peters C. J. Modeling CO2 solubility in an ionic liquid: a comparison between a cubic and a group contribution EoS. J. Supercrit. Fluids 2015, 101, 54–62. 10.1016/j.supflu.2015.02.024. [DOI] [Google Scholar]
- Liu S.; Ling H.; Gao H.; Tontiwachwuthikul P.; Liang Z.; Zhang H. Kinetics and new bronsted correlations study of CO2 absorption into primary and secondary alkaloamines with and without steric-hindrance. Sep. Purif. Technol. 2020, 233, 115998 10.1016/j.seppur.2019.115998. [DOI] [Google Scholar]
- Papadopoulos A. I.; Shavalieva G.; Papadokonstantakis S.; Seferlis P.; Perdomo F. A.; Galindo A.; Jackson G.; Adjiman C. S. An approach for simultaneous computer-aided molecular design with holistic sustainability assessment: application to phase-change CO2 capture solvents. Comput. Chem. Eng. 2020, 135, 106769 10.1016/j.compchemeng.2020.106769. [DOI] [Google Scholar]
- Gonzalez-Miquel M.; Talreja M.; Ethier A. L.; Flack K.; Switzer J. R.; Biddinger E. J.; Pollet P.; Palomar J.; Rodriguez F.; Eckert C. A.; Liotta C. L. COSMO-RS studies: structure-property relationships for CO2 capture by reversible ionic liquids. Ind. Eng. Chem. Res. 2012, 51, 16066–16073. 10.1021/ie302449c. [DOI] [Google Scholar]
- Kim I.; Hoff K. A.; Hessen E. T.; Warberg T. H.; Svendsen H. F. Enthalpy of absorption of CO2 with alkanolamine solutions predicted from reaction equilibrium constants. Chem. Eng. Sci. 2009, 64, 2027–2038. 10.1016/j.ces.2008.12.037. [DOI] [Google Scholar]
- Dehghani M.; Asghari M.; Mohammadi A. H.; Mokhtari M. Molecular simulation and Monte Carlo study of structural-transport-properties of PEBA-MFI zeolite mixed matric membranes for CO2, CH4 and N2 separation. Comput. Chem. Eng. 2017, 103, 12–22. 10.1016/j.compchemeng.2017.03.002. [DOI] [Google Scholar]
- Jayarathna S. A.; Lie B.; Melaaen M. C. Dynamic modelling of the absorber of a post-combustion CO2 capture plant: modeling and simulations. Comput. Chem. Eng. 2013, 53, 178–189. 10.1016/j.compchemeng.2013.03.002. [DOI] [Google Scholar]
- Haron N.; Sairi N. A.; Lee V. S. Microstructures, interactions and dynamics properties studies of N-methyldiethanolamine + guanidinium triflate ionic liquid + water tertiary system at the standard temperature. Mol. Simul. 2016, 42, 655–666. 10.1080/08927022.2015.1068942. [DOI] [Google Scholar]
- Mohan O.; Trinh Q. T.; Banerjee A.; Mushrif S. H. Predicting CO2 adsorption and reactivity on transition metal surfaces using popular density functional theory methods. Mol. Simul. 2019, 45, 1163–1172. 10.1080/08927022.2019.1632448. [DOI] [Google Scholar]
- Houndonougbo Y.; Kuczera K.; Subramaniam B.; Laird B. B. Prediction of phase equilibria and transport properties in carbon-dioxide expanded solvents by molecular simulation. Mol. Simul. 2007, 33, 861–869. 10.1080/08927020701310923. [DOI] [Google Scholar]
- Whangbo M.-H.; Gordon E. E.; Xiang H.; Koo H. J.; Lee C. Prediction of spin orientations in terms of HOMO-LUMO interactions using spin-orbit coupling as perturbation. Acc. Chem. Res. 2015, 48, 3080–3087. 10.1021/acs.accounts.5b00408. [DOI] [PubMed] [Google Scholar]
- Klamt A. COSMO-RS for aqueous solvation and interfaces. Fluid Phase Equilib. 2016, 407, 152–158. 10.1016/j.fluid.2015.05.027. [DOI] [Google Scholar]
- Ossman T.; Fabre G.; Trouillas P. Interaction of wine anthocyanin derivatives with lipid bilayer membranes. Comput. Theor. Chem. 2016, 1077, 80–86. 10.1016/j.comptc.2015.10.034. [DOI] [Google Scholar]
- Mehling T.; Ingram T.; Storm S.; Bobe U.; Liu F.; Michel M.; Smirnova I. Estimation of LPC/water partition coefficients using molecular modeling and micellar liquid chromatography. Colloids Surf., A 2013, 431, 105–113. 10.1016/j.colsurfa.2013.04.028. [DOI] [Google Scholar]
- Ritter E.; Racheva R.; Jakobtorweihen S.; Smirnova I. Influence of D-glucose as additive on thermodynamics and physical properties of aqueous surfactants two-phase systems for the continuous micellar extraction. Chem. Eng. Res. Des. 2017, 121, 149–162. 10.1016/j.cherd.2017.02.032. [DOI] [Google Scholar]
- Dołżonek J.; Cho C. W.; Stepnowski O.; Markiewicz M.; Thoming J.; Stolte S. Membrane partitioning of ionic liquid cations, anions and ion pairs- estimating the bioconcentration potential of organic ions. Environ. Pollut. 2017, 228, 378–389. 10.1016/j.envpol.2017.04.079. [DOI] [PubMed] [Google Scholar]
- Yordanova D.; Smirnova I.; Jakobtorweihen S. Molecular modeling of Triton X Micelles: force field parameters self-assembly, and partition equilibria. J. Chem. Theory Comput. 2015, 11, 2329–2340. 10.1021/acs.jctc.5b00026. [DOI] [PubMed] [Google Scholar]
- Storm S.; Aschenbrenner D.; Smirnova I. Reverse miscellar extraction of amino acids and complex enzyme mixtures. Sep. Purif. Technol. 2014, 123, 23–24. 10.1016/j.seppur.2013.11.035. [DOI] [Google Scholar]
- Turchi M.; Cai Q.; Lian G. An evaluation of in-silico methods for predicting solute partition in multiphase complex fluids- a case study of octanol/water partition coefficient. Chem. Eng. Sci. 2019, 197, 150–158. 10.1016/j.ces.2018.12.003. [DOI] [Google Scholar]
- Liang X.; Li Y.; Wu X.; Shen J.; Lee K. Y. Nonlinearity analysis and multi-model modeling of an MEA-based post-combustion CO2 capture process for advanced control design. Appl. Sci. 2018, 8, 1053 10.3390/app8071053. [DOI] [Google Scholar]
- Karlsson H.; Svensson H. Rate of absorption for CO2 absorption systems using a wetted wall column. Energy Procedia 2017, 114, 2009–2023. 10.1016/j.egypro.2017.03.1335. [DOI] [Google Scholar]
- Kim Y. E.; Choi J. H.; Nam S. C.; Yoon Y. I. CO2 absorption capacity using aqueous potassium carbonate with 2-methylpiperazine and piperazine. J. Ind. Eng. Chem. 2012, 18, 105–110. 10.1016/j.jiec.2011.11.078. [DOI] [Google Scholar]
- Yuan Y.; Sherman B.; Rochelle G. T. Effects of viscosity on CO2 absorption in aqueous piperazine / 2-methylpiperazine. Energy Procedia 2017, 114, 2103–2120. 10.1016/j.egypro.2017.03.1345. [DOI] [Google Scholar]
- Hafizi A.; Mokari M. H.; Khalifeh R.; Farsi M.; Rahimpour M. R. Improving the CO2 solubility in aqueous mixture of MDEA and different polyamine promoters: The effects of primary and secondary functional groups. J. Mol. Liq. 2019, 111803 10.1016/j.molliq.2019.111803. [DOI] [Google Scholar]
- Zong L.; Chen C. C. Thermodynamic modeling of CO2 and H2S solubilities in aqueous DIPA solution, aqueous sulfolane-DIPA solution, and aqueous sulfolane-MDEA solution with electrolyte NRTL model. Fluid Phase Equilib. 2011, 306, 190–203. 10.1016/j.fluid.2011.04.007. [DOI] [Google Scholar]
- Shokouhi M.; Jalili A. H.; Zoghi A. T.; Ahari J. S. Carbon dioxide solubility in aqueous sulfolane solution. J. Chem. Thermodyn. 2019, 132, 62–72. 10.1016/j.jct.2018.12.004. [DOI] [Google Scholar]
- Jalili A. H.; Shokouhi M.; Samani F.; Jenab M. J. Measuring the solubility of CO2 and H2S in sulfolane and the density and viscosity of saturated liquid binary mixtures of (sulfolane + CO2) and (sulfolane + H2S). J. Chem. Thermodyn. 2015, 85, 13–25. 10.1016/j.jct.2015.01.001. [DOI] [Google Scholar]
- Safarov J.; Geppert-Rybczynska M.; Kul I.; Hassel E. Thermophysical properties of 1-butyl-3-methylimidazolium acetate over a wide range of temperatures and pressures. Fluid Phase Equilib. 2014, 383, 144–155. 10.1016/j.fluid.2014.10.015. [DOI] [Google Scholar]
- Osman K.; Ramjugernath D.; Coquelet C. CO2 solubility in hybrid solvents containing 1-butyl-3-methylimidazolium tetrafluoroborate and mixtures of alkanolamine. J. Chem. Eng. Data 2015, 60, 2380–2391. 10.1021/acs.jced.5b00273. [DOI] [Google Scholar]
- Chakravarty T.; Phukan U.; Weilund R. Reaction of acid gases with mixtures of amines. Chem. Eng. Prog. 1985, 81, 32–36. [Google Scholar]
- Kohl A. L.; Nielsen R.. Gas Purification; Gulf Professional Publishing, 1997. [Google Scholar]
- Aghehrochaboki R.; Chaboki Y. A.; Maleknia S. A.; Irani V. Polyethyleneimine functionalized grapheme oxide/methyldiethanolamine nanofluid: preparation, characterization, and investigation of CO2 absorption. J. Environ. Chem. Eng. 2019, 7, 103285 10.1016/j.jece.2019.103285. [DOI] [Google Scholar]
- Eckert F.; Klamt A.. COSMOtherm, release 19.0.1; COSMOlogic GmbH & Co. KG: Leverkusen, 2013.
- Sumon K. Z.; Henni A. Ionic liquids for CO2 capture using COSMO-RS: effect of structure, properties and molecular interactions on solubility and selectivity. Fluid Phase Equilib. 2011, 310, 39–55. 10.1016/j.fluid.2011.06.038. [DOI] [Google Scholar]
- Klamt A.; Eckert F. COSMO-RS: a novel and efficient method for the priori prediction of thermophysical data of liquids. Fluid Phase Equilib. 2000, 172, 43–72. 10.1016/S0378-3812(00)00357-5. [DOI] [Google Scholar]
- Matheswaran P.; Wilfred C. D.; Kurnia K. A.; Ramli A. Overview of activity coefficient of thiophene at infinite dilution in ionic liquids and their modeling using COSMO-RS. Ind. Eng. Chem. Res. 2016, 55, 788–797. 10.1021/acs.iecr.5b04152. [DOI] [Google Scholar]
- Sun C.; Dutta P. K. Infrared Spectroscopic Study of Reaction of Carbon Dioxide with Aqueous Monoethanolamine Solutions. Ind. Eng. Chem. Res. 2016, 55, 6276–6283. 10.1021/acs.iecr.6b00017. [DOI] [Google Scholar]
- Richner G.; Puxty G. Assessing the Chemical Speciation during CO2 absorption by Aqueous Amines using in Situ FTIR. Ind. Eng. Chem. Res. 2012, 51, 14317–14324. 10.1021/ie302056f. [DOI] [Google Scholar]
- Valderrama J. O.; Forero L. A. An analytical expression for the vapor pressure of ionic liquids based on an equation of state. Fluid Phase Equilib. 2012, 317, 77–83. 10.1016/j.fluid.2011.12.021. [DOI] [Google Scholar]
- Peng D. Y.; Robinson D. B. A New Two-Constant Equation of State. Ind. Eng. Chem. Fundam. 1976, 15, 59–64. 10.1021/i160057a011. [DOI] [Google Scholar]
- Katritzky A. R.; Slavov S. H.; Dobchev D. A.; Karelson M. Rapid QSPR model development technique for prediction of vapor pressure of organic compounds. Comput. Chem. Eng. 2007, 31, 1123–1130. 10.1016/j.compchemeng.2006.10.001. [DOI] [Google Scholar]
- Lei Z.; Yu G.; Su Y.; Dai C. Vapor pressure measurements and predictions for the binary and ternary systems containing ionic liquid [EMIM] [Tf2N]. J. Mol. Liq. 2017, 231, 272–280. 10.1016/j.molliq.2017.01.110. [DOI] [Google Scholar]
- Hekayati J.; Roosta A.; Javanmardi J. On the prediction of the vapor pressure of ionic liquids based on the principle of corresponding states. J. Mol. Liq. 2017, 225, 118–126. 10.1016/j.molliq.2016.11.031. [DOI] [Google Scholar]
- Balchandani S.; Mandal B. P.; Dharaskar S. Measurements and modeling of vapor liquid equilibrium of CO2 in amine activated imidazolium ionic liquid solvents. Fluid Phase Equilib. 2020, 521, 112643 10.1016/j.fluid.2020.112643. [DOI] [Google Scholar]
- Joshipura M. H.; Dabke S. P.; Subrahmanyam N. Development and comparison of cohesion function relationship for PR equation of state. Int. J. Chem. Eng. Res. 2009, 1, 123–134. [Google Scholar]
- Valderrama J. O.; Forero L. A. An analytical expression for the vapor pressure of ionic liquids based on an equation of state. Fluid Phase Equilib. 2012, 317, 77–83. 10.1016/j.fluid.2011.12.021. [DOI] [Google Scholar]
- Kumar S.; Cho J. H.; Moon I. Ionic liquid-amine blends and CO2 BOLs: prospective solvents for natural gas sweetening and CO2 capture technology-a review. Int. J. Greenhouse Gas Control 2014, 20, 87–116. 10.1016/j.ijggc.2013.10.019. [DOI] [Google Scholar]
- Wappel D.; Gronald G.; Kalb R.; Draxler J. Ionic liquids for post-combustion CO2 absorption. Int. J. Greenhouse Gas Control 2010, 4, 486–494. 10.1016/j.ijggc.2009.11.012. [DOI] [Google Scholar]
- Paul S.; Ghoshal A. K.; Mandal B. P. Kinetics of absorption of carbon dioxide into aqueous blends of 2-(1-piperazinyl)-ethylamine and N-methyldiethanolamine. Chem. Eng. Sci. 2009, 64, 1618–1622. 10.1016/j.ces.2008.12.034. [DOI] [Google Scholar]
- Balchandani S.; Mandal B. P.; Dharaskar S.; Kumar A.; Bandyopadhyay S. S. Thermally induced characterization and modeling of physiochemical, acoustic rheological, and thermodynamic properties of novel blends of (HEF + AEP) and (HEF + AMP) for CO2/H2S absorption. Environ. Sci. Pollut. Res. 2019, 26, 32209–32223. 10.1007/s11356-019-06305-5. [DOI] [PubMed] [Google Scholar]
- Aspen Technology Inc. Aspen Physical Property System, version 7.1; Aspen Technology Inc.: Cambridge, 2008.
- Bara J. E.; Moon J. D.; Reclusado K. R.; Whitley J. W. COSMOtherm as a tool for estimating the thermophysical properties of alkylimidazoles as solvents for CO2 separations. Ind. Eng. Chem. Res. 2013, 52, 5498–5506. 10.1021/ie400094h. [DOI] [Google Scholar]
- Diedenhofen M.; Eckert F.; Klamt A. Prediction of infinite dilution activity coefficients of organic compounds in ionic liquids using COSMO-RS. J. Chem. Eng. Data 2003, 48, 475–479. 10.1021/je025626e. [DOI] [Google Scholar]
- Putnam R.; Taylor R.; Klamt A.; Eckert F.; Schiller M. Prediction of infinite dilution activity coefficients using COSMO-RS. Ind. Eng. Chem. Res. 2003, 42, 3635–3641. 10.1021/ie020974v. [DOI] [Google Scholar]
- Paduszyński K. An overview of performance of COSMO-RS approach in predicting activity coefficients of molecular solutes in ionic liquids and derived properties at infinite dilution. Phys. Chem. Chem. Phys. 2017, 19, 11835–11850. 10.1039/C7CP00226B. [DOI] [PubMed] [Google Scholar]
- Voutsas E. C.; Tassios D. P. Prediction of Infinite-Dilution Activity Coefficients in Binary Mixtures with UNIFAC. A Critical Evaluation. Ind. Eng. Chem. Res. 1996, 35, 1438–1445. 10.1021/ie9503555. [DOI] [Google Scholar]
- Dash S. K.; Samanta A. N.; Bandyopadhyay S. S. Solubility of carbon dioxide in aqueous solution of 2-amino-2-methyl-1-propanol and piperazine. Fluid Phase Equilib. 2011, 307, 166–174. 10.1016/j.fluid.2011.05.009. [DOI] [Google Scholar]
- Renon H.; Prausnitz J. M. Local compositions in thermodynamic excess functions for liquid mixtures. AIChE J. 1968, 14, 135–144. 10.1002/aic.690140124. [DOI] [Google Scholar]
- Liang Y.; Liu H.; Rongwong W.; Liang Z.; Idem R.; Tontowachwuthikul P. Solubility, absorption heat and mass transfer studies of CO2 absorption into aqueous solution of 1-dimethylamino-2-propanol. Fuel 2015, 144, 121–129. 10.1016/j.fuel.2014.11.098. [DOI] [Google Scholar]
- Ling H.; Liu S.; Gao H.; Zhang H.; Liang Z. Solubility of N2O, equilibrium solubility, mass transfer study and modeling of CO2 absorption into aqueous monoethanolamine (MEA)/1-dimethylamino-2-propanol (1DMA2P) solution for post-combustion CO2 capture. Sep. Purif. Technol. 2020, 232, 115957 10.1016/j.seppur.2019.115957. [DOI] [Google Scholar]
- Edwards T. J.; Maurer G.; Newman J.; Prausnitz J. M. Vapor-liquid equilibria in multicomponent aqueous solutions of volatile weal electrolytes. AIChE J. 1978, 24, 966–976. 10.1002/aic.690240605. [DOI] [Google Scholar]
- Benamor A.; Aroua M. K. Modeling of CO2 solubility and carbamate concentration in DEA, MDEA and their mixtures using Deshmukh-Mather model. Fluid Phase Equilib. 2005, 231, 150–162. 10.1016/j.fluid.2005.02.005. [DOI] [Google Scholar]






