1,891
Views
42
CrossRef citations to date
0
Altmetric
Review Series

Part 5c: Solvent chemistry: solubility of CO2 in reactive solvents for post-combustion CO2

, , , , &
Pages 467-484 | Published online: 10 Apr 2014

Abstract

Concerns about climate change have propelled research efforts to develop affordable and environmentally benign technologies to capture CO2 from large emission sources, which can subsequently be used either for enhanced oil recovery or stored in other geological storage sites. The study of the solubility of CO2 in solvents is therefore of great interest, both from the theoretical and practical points of view. To screen solvents or to design CO2 capture processes, knowledge of the equilibrium solubility of CO2 in the solvents is necessary. A large body of solubility data of CO2 in aqueous and non-aqueous solutions of prominent industrial amines are available in literature. We present such data along with a description of the experimental techniques and thermodynamic models used. Innovations made to obtain an optimum solubility of CO2 and to minimize the energy cost of a desired CO2 capture system by adopting different kinds of solvents are also reviewed.

Figure 1.  High-pressure solubility data for CO2 in high concentrated chemical solvents at 40°C.

Red: primary amines; turquoise: sterically hindered amines; gray: secondary amines; blue: diamines.

1-MPZ: 1-methyl piperazine; AEEA: 2-((2-aminoethyl)amino)ethanol; AEPD: 2-amino-2-ethyl-1,3-propanediol; AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; AMPD: 2-amino-1-methyl-1,3-propanediol; DEA: Diethanolamine; DGA: Diglycolamine; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; MIPA: Monoisopropanolamine; PE: 2-piperidineethanol; PZ: Piperazine; TIPA: Triisopropanolamine.

Data taken from Citation[17,19,29–42].

Figure 1.  High-pressure solubility data for CO2 in high concentrated chemical solvents at 40°C.Red: primary amines; turquoise: sterically hindered amines; gray: secondary amines; blue: diamines.1-MPZ: 1-methyl piperazine; AEEA: 2-((2-aminoethyl)amino)ethanol; AEPD: 2-amino-2-ethyl-1,3-propanediol; AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; AMPD: 2-amino-1-methyl-1,3-propanediol; DEA: Diethanolamine; DGA: Diglycolamine; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; MIPA: Monoisopropanolamine; PE: 2-piperidineethanol; PZ: Piperazine; TIPA: Triisopropanolamine.Data taken from Citation[17,19,29–42].
Figure 2.  Low-pressure solubility data for CO2 in chemical solvents (low concentrated) at 40°C.

Red: primary amines; turquoise: sterically hindered amines; gray: secondary amines; blue: diamines; black: tertiary amines.

3AP: 3-aminopyridine-2-carboxaldehyde thiosemicarbazone; 4DEA2B: 4-(diethylamino)-2-butanol; AEEA: 2-((2-aminoethyl)amino)ethanol; AEPD: 2-amino-2-ethyl-1,3-propanediol; AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; DEA: Diethanolamine; DGA: Diglycolamine; DIPA: Diisopropanolamine; HXDD: 1,6-hexanediamine; MAE: 2-methylamino ethanol; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; MO: Morpholine; PZ: Piperazine; TEA: Triethanolamine; TIPA: Triisopropanolamine.

Data taken from Citation[19,33,41–54].

Figure 2.  Low-pressure solubility data for CO2 in chemical solvents (low concentrated) at 40°C.Red: primary amines; turquoise: sterically hindered amines; gray: secondary amines; blue: diamines; black: tertiary amines.3AP: 3-aminopyridine-2-carboxaldehyde thiosemicarbazone; 4DEA2B: 4-(diethylamino)-2-butanol; AEEA: 2-((2-aminoethyl)amino)ethanol; AEPD: 2-amino-2-ethyl-1,3-propanediol; AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; DEA: Diethanolamine; DGA: Diglycolamine; DIPA: Diisopropanolamine; HXDD: 1,6-hexanediamine; MAE: 2-methylamino ethanol; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; MO: Morpholine; PZ: Piperazine; TEA: Triethanolamine; TIPA: Triisopropanolamine.Data taken from Citation[19,33,41–54].
Figure 3.  Change in the solubility of CO2 in aqueous methyldiethanolamine after addition of other amines at 40°C.

AMP: 2-amino-2-methyl-1-propanol; DEA: Diethanolamine; DIPA: Diisopropanolamine; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; PZ: Piperazine.

Data taken from Citation[46,55–58].

Figure 3.  Change in the solubility of CO2 in aqueous methyldiethanolamine after addition of other amines at 40°C.AMP: 2-amino-2-methyl-1-propanol; DEA: Diethanolamine; DIPA: Diisopropanolamine; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; PZ: Piperazine.Data taken from Citation[46,55–58].
Figure 4.  Change of solubility of CO2 in piperazine after the addition of other amines at 40°C.

AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; DEA: Diethanolamine; DIPA: Diisopropanolamine; MDEA: Methyldiethanolamine; PZ: Piperazine; TEA: Triethanolamine; TIPA: Triisopropanolamine.

Data taken from Citation[22,55,58–63].

Figure 4.  Change of solubility of CO2 in piperazine after the addition of other amines at 40°C.AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; DEA: Diethanolamine; DIPA: Diisopropanolamine; MDEA: Methyldiethanolamine; PZ: Piperazine; TEA: Triethanolamine; TIPA: Triisopropanolamine.Data taken from Citation[22,55,58–63].
Figure 5.  Change in the solubility of CO2 in aqueous diglycolamine by addition of morpholine at 25°C.

DGA: Diglycolamine; MO: Morpholine.

Data taken from Citation[52].

Figure 5.  Change in the solubility of CO2 in aqueous diglycolamine by addition of morpholine at 25°C.DGA: Diglycolamine; MO: Morpholine.Data taken from Citation[52].
Figure 6.  Change in the solubility of CO2 at 100 kPa.

3AP: 3-aminopyridine-2-carboxaldehyde thiosemicarbazone; AEEA: 2-((2-aminoethyl)amino)ethanol; AEPD: 2-amino-2-ethyl-1,3-propanediol; AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; AMPD: 2-amino-1-methyl-1,3-propanediol; DEA: Diethanolamine; DGA: Diglycolamine; DIPA: Diisopropanolamine; MAE: 2-methylamino ethanol; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; PZ: Piperazine; TEA: Triethanolamine; TIPA: Triisopropanolamine.

Data taken from Citation[17,30,31,33–36,41,42,53,54,56,60,61,64].

Figure 6.  Change in the solubility of CO2 at 100 kPa.3AP: 3-aminopyridine-2-carboxaldehyde thiosemicarbazone; AEEA: 2-((2-aminoethyl)amino)ethanol; AEPD: 2-amino-2-ethyl-1,3-propanediol; AHPD: 2-amino-2-(hydroxymethyl)-1, 3-propanediol; AMP: 2-amino-2-methyl-1-propanol; AMPD: 2-amino-1-methyl-1,3-propanediol; DEA: Diethanolamine; DGA: Diglycolamine; DIPA: Diisopropanolamine; MAE: 2-methylamino ethanol; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; PZ: Piperazine; TEA: Triethanolamine; TIPA: Triisopropanolamine.Data taken from Citation[17,30,31,33–36,41,42,53,54,56,60,61,64].
Figure 7.  Comparison of the solubility of CO2 in different mixtures of ionic liquids and amines.

[BMIM][BF4]: 1-butyl-3-methylimidazolium tetraflluoroborate; BHEAA: Bis(2-hydroxyethyl)ammonium acetate; MDEA: Methyldiethanolamine; MEA: Monoethanolamine.

Data taken from Citation[65–67].

Figure 7.  Comparison of the solubility of CO2 in different mixtures of ionic liquids and amines.[BMIM][BF4]: 1-butyl-3-methylimidazolium tetraflluoroborate; BHEAA: Bis(2-hydroxyethyl)ammonium acetate; MDEA: Methyldiethanolamine; MEA: Monoethanolamine.Data taken from Citation[65–67].
Figure 8.  CO2 capture capacity of aqueous ionic liquids, amines and their mixtures.

[BMIM][BF4]: 1-butyl-3-methylimidazolium tetraflluoroborate; IL: Ionic liquid; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; PZ: Piperazine; TEA: Triethanolamine.

Data taken from Citation[68].

Figure 8.  CO2 capture capacity of aqueous ionic liquids, amines and their mixtures.[BMIM][BF4]: 1-butyl-3-methylimidazolium tetraflluoroborate; IL: Ionic liquid; MDEA: Methyldiethanolamine; MEA: Monoethanolamine; PZ: Piperazine; TEA: Triethanolamine.Data taken from Citation[68].
Figure 9.  Comparison of the experimental data for CO2 partial pressure of the methyldiethanolamine–H2O–CO2 system and the model results.

Methyldiethanolamine concentration is ∼8 M.

e-NRTL: Electrolyte nonrandom two-liquid; MDEA: Methyldiethanolamine; T: Temperature.

Reproduced with permission from Citation[69] © American Chemical Society (2011).

Figure 9.  Comparison of the experimental data for CO2 partial pressure of the methyldiethanolamine–H2O–CO2 system and the model results.Methyldiethanolamine concentration is ∼8 M.e-NRTL: Electrolyte nonrandom two-liquid; MDEA: Methyldiethanolamine; T: Temperature.Reproduced with permission from Citation[69] © American Chemical Society (2011).

The reduction and control of atmospheric CO2 has received considerable attention due to its contribution to climate change. Post-combustion capture of CO2 from large stationary sources such as power plants and subsequent storage could reduce anthropogenic CO2 emissions considerably. The process of separating CO2 from flue gas through chemical absorption with a suitable reactive solvent is considered to be the most viable technology in the short and medium terms. The selection of a solvent for commercial post-combustion CO2 capture is a complex process and requires laboratory-scale and pilot plant data, as well as the use of accurate process simulators. Preliminary screening of such solvents has been carried out in many laboratories all over the world. One of the key components of solvent screening is the determination of the equilibrium solubility of CO2 as a function of temperature and pressure. Such information is used to determine the solvent circulation rate to achieve a certain degree of CO2 removal. Recent screening solubility studies were published by Singh et al.Citation[1,2], Puxity et al.Citation[3], Chowdhury et al.Citation[4,5] and Porcheron et al.Citation[6,7]. In this present article, we report on the available experimental techniques and data for the solubility of CO2 in the most common and promising solvents of interest to CO2 capture operations. The latest reviews on the solubility of CO2 in amine solutions were published in 2000 by Rochelle et al.Citation[8] and in 2007 by Anufrikov et al.Citation[9].

Vapor–liquid equilibrium or equilibrium solubility

Data gathered in this review provides a comprehensive database of vapor–liquid equilibrium (VLE) data, which can be used for testing and developing theoretical models and correlations, and are of interest to process design software developers and researchers working on solvent screening studies. The overall cost of the CO2 capture plant is determined partially by the thermophysical properties of the solvent. Incorrect estimates of the absorber size can lead to large columns, and failure to meet product specifications are quite often the result of inaccurate VLE data Citation[10]. Although VLE data are of great importance, the reported literature data for the solubility of CO2 in aqueous solvents are scattered Citation[11,12]. Therefore, there is a need for obtaining good quality and internally consistent VLE data in order to prove the extent and the accuracy of existing thermodynamic databases for solvent–CO2 systems Citation[13].

Experimental methods

The selection of the experimental method for VLE depends not only on the thermodynamic conditions, but also on the type of system, properties of the individual components, the required amount of material, the analytical amenities of measuring the composition of the equilibrium phases, the demand for accuracy of the measurement and many other factors. In CO2 capture research, the existing experimental set-ups can be classified in several ways Citation[14,15]. They can be classified as closed- or open-circuit methods or as analytical (or direct sampling method) and synthetic (indirect methods). In the analytical method, both liquid and vapor phases are sampled and analyzed. For example, the vapor phase composition is determined by chromatography, whereas the liquid phase is analyzed by titration (or chromatography). In the synthetic method, a mass balance is used for calculating the amount of acid gas absorbed by the solvent. The acid gas quantity introduced in the equilibrium cell is determined based on the knowledge of the pressure–volume–temperature conditions Citation[16]. Three different types of VLE cells have been used extensively to obtain equilibrium solubility data. These cells are used to bring the vapor and liquid phases into thermal and composition equilibrium at known temperatures and pressures. The cells differ in the way they bring together (mix) the two phases. The compositions of the equilibrium phases are then determined by withdrawing samples and analyzing them.

▪ Static analytic method

Static analytical methods involve the establishment of a thermodynamic equilibrium in the cell followed by the analysis of solution samples. The constraints of the static analytical methods are those of the applied analysis technique such as spectroscopy or chromatography. It is important to use a reliable sampling technique, which allows withdrawing samples that are small enough not to perturb the established equilibrium. A known amount of solvent is placed in the cell and a required amount of CO2 is introduced. The system is kept until equilibrium is established and then the pressure and mole fractions of components in both vapor and liquid phases are measured. A typical example of static analytical method, shown in Supplementary Figure 1, was used by Ma’mun et al.Citation[16].

Noncirculation method

Rocking design was used by Lawson and Garst (Supplementary Figure 2)Citation[17]. Severe corrosion to the point of failure in the weld of the sample port was the major disadvantage observed for this kind of equipment. An internal/mechanical stirrer (Supplementary Figure 3) was used by Tong et al.Citation[13]. Reliable and repeatable results were obtained by this method for a wide range of temperatures. A variable-volume static cell was used by Addicks and Owren (Supplementary Figure 4)Citation[18]. A major disadvantage of this method is that the mass transport over the small vapor–liquid surface in the equilibrium cell is so slow that without stirring, pressure, volume and temperature can become constant without equilibrium being reached Citation[18].

Circulation method

In the circulation method, a known amount of solvent is placed in the equilibrium cell, nitrogen is introduced into the cell for CO2 batching after the temperature is equilibrated, and the equilibrium pressure due to the amine and water vapor pressures and the nitrogen pressure are recorded. CO2 is then introduced into the cell and the change in the pressure of the batching vessel is recorded. Equilibrium is considered to be attained if the total pressure in the cell remains constant for a long interval of time. The equilibrium partial pressure of CO2 is the difference between the total pressure and the nitrogen equilibrium pressure recorded before. Both liquid and vapor phases were sampled for analysis. A circulation method (Supplementary Figure 5)Citation[18] was used by Kadiwala et al.Citation[19]. Although this kind of apparatus is used for a wide range of pressures and temperatures, their accuracy at low pressures is less compared with noncirculation static synthetic methods.

Static synthetic method

Static synthetic methods do not require sampling devices, but visualize the VLE process of a known composition. The equipment is simple, but the preparation of the solution to be studied has to be done adequately. The solubility is measured at a fixed temperature and a complete data set for multisolute solutions is calculated by linking with an analytical technique. Sidi-Boumedine et al. used computer-operated static apparatus (Supplementary Figure 6)Citation[12]. The obtained data were high in accuracy and reproducibility with the data obtained by the analytical method.

▪ Flow method

Partial pressure of CO2 in the gas flow is fixed in the course of an experiment and the composition of the liquid solution through which the gas flow is passed changes until the solution comes to equilibrium with the gas. The quantities set in the experiment are the temperature, number of moles of water and amine in the solution, and the partial pressure of CO2. A typical example of flow system is shown in Supplementary Figure 7Citation[20]. A series of equilibrium samples were necessary to check the reproducibility and accuracy. Detailed discussion of these experimental techniques was published by Anufrikov et al.Citation[9]. There are two types of methods available to approach equilibrium.

The static method (closed circuit) is where gas mixtures are bubbled and recirculated through the solvent. After equilibrium is reached, circulation is stopped and samples are withdrawn for analysis. A closed circuit cell used by Ma’mun et al. is shown in Supplementary Figure 8Citation[21].

The dynamic method (open circuit) is where gas mixtures are passed continuously through an absorber and the solubility is continuously measured. This method allows the gas to be analyzed without disturbing the equilibrium, but it is more difficult to operate. An absorption cell used by Daneshvar et al. is shown in Supplementary Figure 9Citation[22]. CO2 content in the liquid is usually analyzed with the following methods:

▪ Titration method: BaCO3 precipitation and titration with an acid Citation[11];

▪ Knorr method: evolving CO2 from the solution by addition of acid and absorbing it in ascarite Citation[23];

▪ Chittck apparatus Citation[24];

▪ Calibrated gas chromatography Citation[11];

▪ Infrared CO2 analyzer Citation[16];

▪ Fourier infrared spectroscopy Citation[25];

▪ 13C nuclear magnetic resonance spectroscopy: used to evaluate carbamates stability constant, distribution of CO2 species and individual ionic concentrations Citation[26];

▪ Raman spectroscopy: to evaluate carbamates stability constant, distribution of CO2 species and individual ionic concentrations Citation[27];

▪ Material balance Citation[28].

Supplementary Table 1 summarizes the experiments conducted during the past eight decades (1936–2012) with the developments made in phase analysis.

Analysis of the published solubility data

Solubility of CO2 in aqueous monoethanolamine (MEA), diethanolamine (DEA), methyldiethanolamine (MDEA), 2-amino-2-methyl-1-propanol (AMP), piperazine (PZ) and amine blends have been reported by many researchers at different temperatures, pressures and amine concentrations (Supplementary Tables 2–7). In order to analyze the solubility data available for different solvents, the change in the solubility (loading, α = molCO2/molamine) with respect to pressures and temperatures were studied in the same concentration ranges. Flue gases resulting from combustion processes are available at low-pressure conditions. The concentration of CO2 in the flue gas stream is rather low and can vary in general from 5 to 14% by volume, depending on the type of fuel used. The low concentration of CO2 combined with the low pressure of the gas stream reduces the partial pressure of CO2. Since the absorption process is driven by partial pressure gradient, the absorption of CO2 by chemical solvents is favored under high-pressure conditions. The challenge is to find a solvent that would be effective in absorbing CO2 under low partial pressure conditions. The partial pressure of CO2 varies from 100 kPa (blast furnace) to 4 kPa in the case of turbine exhaust gas (14 kPa in coal-fired power plant). The partial pressure of CO2 increases in the stripper at 120°C Citation[29]. As solubility data are, in general, not available at very low pressures for a wide range of solvents, a range of partial pressures from 10 kPa to 500 kPa was used to compare different solvents in terms of CO2 capacity and demonstrate that the choice of ‘best’ solvent depends on the partial pressure of CO2.

▪ Change in the solubility of CO2 in solvents by changing pressures

represents the high-pressure solubility of CO2 available in the literature in highly concentrated chemical solvents (10–60 wt%) at 40°C Citation[17,19,29–42]. As well known, when the pressure increases the solubility of CO2 in the solvent increases. At 500 kPa, a maximum solubility (1.2 molCO2/molamine) occurs for an aqueous sterically hindered amine diamine, 2-amino-2-(hydroxymethyl)-1, 3-propanediol (AHPD, 10 wt%) followed by an aqueous cyclic diamine, PZ (10 wt%), and low solubility (0.6 molCO2/molamine) occurs for the primary amine, diglycolamine (DGA, 60 wt%). From the structure of these chemicals, it can be concluded that the increase in hindrance to the amino groups increases the loading capacity; the solubility of CO2 in amines follows, in general, an order such that: primary amines < hindered amines < secondary amines < tertiary amines < diamines. represents the low-pressure solubility data from the literature for chemical solvents at low concentration (<30 wt%) at 40°C Citation[19,33,41–54]. At 100 kPa, triethanolamine (30 wt%) has the lowest solubility (0.4 molCO2/molamine) and 1,6-hexadiamine (10 wt%) has the highest solubility (1.4 molCO2/molamine) at 40°C.

represents the change in solubility of a tertiary amine (MDEA) when mixed with other amines Citation[46,55–58]. Adding a hindered amine with a tertiary amine could make a better solvent blend to achieve a maximum loading, as well as an increase in the reaction rate with CO2. shows the change in the solubility of CO2 in aqueous PZ solutions by adding other amines Citation[22,55,58–63]. It can be concluded that the addition of sterically hindered amine with aqueous PZ (promoter) will yield a better mixed solvent with high absorption capacity and absorption rate for CO2. is a good representation of the change in the solubility of CO2 in aqueous DGA (a primary amine) with the addition of morpholine (a cyclic amine) Citation[52]. It can be observed that the addition of morpholine enhances the solubility of CO2 in aqueous DGA (0.23–0.26). At low pressures, the enhancement is low (≈0.001) but is higher at higher pressures (≈0.03).

▪ Changes in the solubility of CO2 in solvents by changing temperatures

The solubility of CO2 in different solvents with respect to temperatures is very important to estimate the amount of heat required to regenerate the solvent for the absorption/regeneration cycle in CO2 capture operations, and in the energy balances involved in the design calculations. As the temperature increases, the solubility of CO2 in the solvent decreases and the slope of this linear change indicates the amount of heat at which the solvent can be regenerated from undesirable products such as the carbamates. The solubility of CO2 in the solvent has to be more at lower temperature (40°C) and less at a higher temperature (120°C). Therefore, the solvent with the higher slope will be the desired one for CO2 capture studies. shows the change in the solubility of some proposed amines for CO2 capture studies with respect to temperature Citation[17,30,31,33–36,41,42,53,56,60,61,64]. It can be observed that the sterically hindered amines have higher slopes, followed by aqueous PZ (cyclic diamine). The general trend of the slopes is sterically hindered amines > tertiary amines > secondary amines > primary amines.

▪ Effect of co-solvent on the solubility of CO2

shows data of mixtures of ionic liquids (ILs) and amines to produce better solvents for CO2 capture operations Citation[65–67]. The maximum loading (1.14) is achieved for the mixed solvent with 20 wt% MEA and 20 wt% 1-butyl-3-methylimidazolium tetraflluoroborate at 40°C and 1500 kPa. It shows that the increase in amine concentration increases the solubility of CO2 in the mixed solvent, and that the solubility of CO2 in this IL mixture is more than the task-specific IL with an amine added to the cation. Although this kind of mixed solvent involving ILs looks profitable in CO2 capture loading capacity, they have high viscosity and are expensive Citation[67]. Zhao et al. reported the changes in the physical properties (densities, viscosities and CO2 capture capacity) during the addition of amines and water in ILs Citation[68]. It was concluded from their work that PZ added to aqueous IL had a good CO2 capture capacity, and did not decrease after four capture and regeneration cycles. shows the CO2 capture capacities of these systems at 40°C and 1500 kPa Citation[68]. Industrially attractive hybrid solvent for effective CO2 capture could be developed by integrating the desirable properties of both ILs as well as amines through the combination of the targeted ILs with selective amines.

VLE modeling

Theoretical modeling of the VLE of the CO2–alkanolamine system is challenging due to the highly non-ideal nature of the liquid phase arising from the complex interactions such as hydrogen bonding among polar molecules, long-range interaction between ions and short-range interaction among molecule–molecule, ion–ion and ion–molecules, lack of accurate model parameters, and availability and quality of the solubility data in the region of operating conditions. A thermodynamically consistent and rigorous model correlates available experimental data, facilitates process modeling and simulation in wider a range of thermodynamic conditions, and guides in screening out questionable data Citation[69].

▪ Thermodynamic framework & modeling approach

The chemical reaction in the liquid phase drives the dissolution of CO2 from vapor to liquid phase. The VLE of CO2–alkanolamine is typically represented as a combination of phase (diffusion) equilibrium between vapor and liquid phase and chemical equilibrium in the liquid phase. Edwards et al. developed the basic thermodynamic framework to treat aqueous solution of weak electrolytes such as CO2, which couples VLE, dissociation balances, mass balances, electroneutrality and deviations from ideality Citation[70]. The phase and chemical equilibrium equations typically used in modeling for a general monoamine are summarized in Supplementary Table 8. At low pressure, the vapor phase is usually treated as ideal, but at higher pressure, the fugacity coefficient is calculated using an equation of state (EoS). The liquid phase non-ideality is thermodynamically quantified by the activity coefficients. The equilibrium constants relate the species concentration with the activity coefficients. In rigorous modeling, the activity coefficients are calculated from an expression of excess Gibbs energy model (Equation 6 in Supplementary Table 8), which are developed from solution theory of electrolytes (e.g., the model of Deshmukh–Mather [DM] Citation[71], Pitzer’s excess Gibbs energy-based models Citation[72], electrolyte nonrandom two-liquid [e-NRTL] model Citation[73–76] and extended universal quasichemical [e-UNIQUAC] model Citation[77,78] or from an expression of residual Helmholtz free energy obtained from an EoS Citation[79–81]. Such rigorous thermodynamic models have emerged in the last few decades and some were further refined or modified. Different researchers have chosen different models for the same system, which shows the comparative predictive power of these models. The models used for solubility of CO2 (or a mixture of gases including CO2) in industrially important amines such as MEA, DEA, MDEA, AMP, PZ and their mixtures and blends with physical solvents are given in Supplementary Tables 2–6, respectively. A brief overview of the popular rigorous models is also given in the Supplementary Information. We summarize the general features of rigorous thermodynamic models in the section titled ‘General features of rigorous thermodynamic models’ and highlight some of the recent applications of popular thermodynamic models in the following sections. Details of these models can be found in Newman et al.Citation[82], and a recent review of thermodynamic modeling of CO2–amine system published by Georgios and Georgios Citation[83].

▪ General features of rigorous thermodynamic models

Representation of species interactions

At the fundamental level, the models differ in their viewpoints in attributing the liquid phase non-idealities to the type of interactions between species which determines the subsequent modeling approach. Long-range ion–ion interactions are invariably considered by all rigorous models through either one of the forms of modified Debye–Hückel expression or mean spherical approximation. However, the treatment of short- or medium-range interactions is usually different in different modeling approaches:

▪ DM model: only the molecule–molecule, molecule–ion binary interaction, Bij:

Equation 1

▪ Pitzer model: binary, ternary or many-body interactions between the solute and ionic species;

▪ e-NRTL: binary interactions between molecule–molecule, molecule–electrolyte, electrolyte–electrolyte interactions only;

▪ e-UNIQUAC: molecule–molecule and molecule–ion interaction; in some variations (molecule–molecule and molecule–ion pair interaction);

▪ Furst and Renon EoS: ion–ion and ion–molecule interaction (usually cation–molecule and cation–anion) interaction.

Physical property requirement

A number of thermodynamic parameters/properties are required to solve these equations, and in general are taken from the literature:

▪ Some equilibrium constants are available in the literature as a function of temperatures. Kim et al. have presented different expressions for the equilibrium constants of the reactions from literature Citation[84];

▪ Solubility of N2O data in pure amine are used to estimate the solubility of CO2 in pure or blended amines. Penttilä et al.Citation[85] have presented a new correlation for Henry’s law constant for CO2 using the N2O analogy and by regression of a large number of solubility data points taken from literature. Henry’s law constants for mixed solvents are estimated from Henry’s law constants in pure solvents. Henry’s law constant for CO2 in water is usually taken from the literature;

▪ Vapor pressure, molar volume and dielectric constants of solvents are also taken from the literature;

▪ Partial molar volume of CO2 at infinite dilution in solvent is computed from correlations.

Interaction parameter regression requirement

Regression of the interaction parameters will depend on the particular model in hand. To enhance the accuracy, additional thermodynamic data (pure component data, e.g., vapor pressure and calorimetric data for binary system: excess enthalpy and VLE of binary system) are used in regression in addition to the VLE data of the actual ternary (CO2–amine–H2O) or quaternary system with blended amines.

▪ Rigorous thermodynamic models & recent applications

Applications of the DM model

Application to 2-((2-aminoethyl)amino)ethanol

Recently, Zoghi et al. applied the DM method to correlate and predict the CO2 solubility in aqueous 2-((2-aminoethyl)amino)ethanol (AEEA) solutions Citation[33]. For correlation purposes, 96 data points for CO2 solubility in 30 wt% of AEEA at temperatures ranging from 313.2 to 368.2 K and CO2 partial pressures ranging from above atmospheric to 4400 kPa were considered. AEEA is a diamine and 14 species are considered to be present at equilibrium participating in ten chemical reactions in the liquid phase. The activity of water is set to its mole fraction in the liquid phase. They imposed the following assumptions and a sensitivity analysis that left only eight binary interaction parameters to be considered:

▪ Symmetry Bij = Bji;

▪ Interaction between like-charged ions are neglected;

▪ Self interaction between molecular species are neglected;

▪ All interactions of molecular and ionic species associated with low concentration species (H+, OH-, +HAEEH+, -OOCAEEACOO- and CO3--) are ignored.

Correlation of data was successful with a percentile absolute average deviation (AAD%) for all the experimental data points of 8.2. Finally, the enthalpy of solution was calculated using the Gibbs–Helmholtz equation at different loading with an average uncertainty of approximately 4.6%. Ma’mun et al. considered 14 Bij values allowing for solute interaction with water, and found the prediction was qualitatively satisfying (data ranged from 0.01 to 220 kPa) Citation[21].

Application to AMP

We cite another example (Pahlavanzadeh et al.) where the application of the DM model was found to be unsatisfactory Citation[86]. Only seven temperature-dependent Bij parameters were considered (Bij = aij + BijT). The average absolute relative error is equal to 30.35%.

Application to diethylenetriamine & MDEA

The diethylenetriamine–H2O–CO2 system consists of 22 species due to the three amine functionalities and, therefore, Harnoto et al. attempted to model the system with all binary interaction parameters equal to zero and the prediction was successful up to loading of 1.0 Citation[87]. They fitted 12 equilibrium constants as functions of first order polynomials (total 24 coefficients) setting all Bij to zero. With Bij equal to zero, the model represents the experimental data reasonably well up to a loading of approximately 1.0; however, at loadings above 1.0 the discrepancies are large. For further agreement, consideration of binary interaction parameters was necessary as shown by a sensitivity test, but they did not do so due to insufficiency of data. The estimated value for the differential heat absorption using the Gibbs–Helmholtz equation of aqueous 2.5 M diethylenetriamine solution deviated greatly in the low loading range up to a loading of approximately 0.6 while the agreement is reasonable for loadings between 0.6 and 1.2. The deviation in the low loading region was attributed to the presence of very few experimental data points in the low loading region.

Dicko et al. found that the DM model was in good agreement correlating solubility data up to a total CO2 + H2S loading of 1.0 in 50 wt% MDEA Citation[88]. Kumar et al. used a thermodynamic framework by employing the virial EoS and the extended Debye–Hückel equation to address the vapor and liquid phase nonideality to model the CO2 solubility data at temperatures of 303.1, 313.1 and 323.1 K and in the CO2 pressure range of 1 to 350 kPa Citation[89]. A total of 12 binary pairs of the CO2 + DEA + AMP + H2O system, and 12 binary pairs of CO2 + DEA + MDEA + H2O were correlated with AADs% of 5.3 and 8, respectively.

Applications of Pitzer model

Application to PZ

Bougie and Iliuta presented a new set of interaction parameters for the Pitzer model for the CO2 + PZ + H2O system by regressing 354 experimental data to cover a wider range of temperature (286.1–395.1 K), pressure and amine concentrations (0.1–8 mol/kg) with an average deviation of 26.1% Citation[90]. Eleven dissolved species are considered to be present in the liquid phase (two molecular species [CO2 and PZ], three cations [PZH+, PZH2+ and H+] and six anions [HCO3-, CO3-- , PZCOO-, PZ(COO-)2, PZH+COO- and OH-]. All ternary interaction parameters were assumed to be zero in order to reduce the binary interaction parameters from a possible total 66 (allowing symmetry); they invoked the following assumptions:

▪ Interaction parameters associated with H+, OH-, PZH2+ and CO3-- are set to zero due to low concentration;

▪ Interaction with species with same sign of charge (including neutral solute–solute) are neglected.

Finally, from the remaining parameters only those which have significant influence on the data are kept, which are a total of eight binary interaction parameters. They were expressed as first order polynomial of temperature (total 16 coefficients).

Application to the AHPD system

Bougie and Iliuta also modeled the solubility of CO2 in an aqueous AHPD system covering a large range of amine concentrations (between 0.0125 and 4 mol kg-1), temperature (between 283.15 and 333.15 K) and total pressure (between 1.85 and 2640.8 kPa) Citation[63]. Using seven temperatures dependent binary interaction parameters (14 coefficients) plus two temperature-independent ternary interaction parameters, they correlated 177 selected experimental data points with an average relative deviation of 22.7%. Generally, higher deviations were obtained at very high amine concentration and very high pressures. Speciation in an aqueous AHPD based on the equilibrium model was also presented.

Application to AMP

Silkenbäumer et al. correlated the solubility of CO2 in aqueous solutions containing AMP in the temperature range from 313 to 353 K at total pressures up to 2.7 MPa, with the Pitzer excess Gibbs energy model mostly within the experimental uncertainty Citation[46].

Applications of e-UNIQUAC model

Application to MEA, MDEA & blend

Faramarzi et al. have applied the e-UNIQUAC model for CO2 absorption in the MEA + H2O, MDEA + H2O and MDEA + MEA + H2O systems Citation[91]. Water was considered the only solvent and all other species are considered solute. Mole fraction-based symmetric and asymmetric convention is adopted for solvent and solute species, respectively. Standard chemical potential was calculated from data taken from the literature or estimated from the thermodynamic relationship (e.g., Gibbs–Helmholtz equation), equilibrium constants and assumptions (heat capacity of some ions are set to zero). Fugacity coefficients were calculated using the Soave–Redlich–Kwong EoS, using the standard mixing rule, and no additional parameters were needed. No adjustable parameters were required for as water was the only solvent and density and dielectric constant of water was required as in the DM or Pitzer models presented above. Evaluation of the term requires size and shape-related pure species properties; that is, UNIQUAC volume parameters (r) and surface area parameters (q). The volume and surface area parameters for MEA, MEAH+, MEACOO-, MDEA and MDEAH+ are determined by fitting to experimental data, and parameters for other species are taken from the literature. Finally, the enthalpic contribution requires energy interaction parameters (and ) among the species present in the system (molecule–molecule, molecule–ion and ion–ion). Due to the like-ion repulsion assumption, only ion–ion parameters with opposite signs were considered. A wide range of data set was used to fine-tune the model parameters, for example:

▪ Excess enthalpy;

▪ Freezing point of the water–MEA system;

▪ CO2–MEA–water and CO2–MDEA–water VLE data;

▪ CO2–MEA–MDEA–H2O VLE data.

Thirteen binary interaction parameters (26 coefficients) were regressed for both MEA and MDEA systems and good agreement was found across a wide number of thermodynamic properties such as excess enthalpy, vapor pressure and speciation for a wide range of condition.

Application to MEA

Aronu et al. represented solubility of CO2 in 15, 30, 45 and 60 wt% MEA from 40 to 120°C using an extended UNIQUAC framework Citation[92]. The Soave–Redlich–Kwong EoS was used to calculate the gas-phase properties. Chemical equilibria in the liquid phase was represented by five chemical reactions involving nine species (three molecular: H2O, CO2 and RNH2; two cations: H3O+ and RNH3+; and four anions: OH-, HCO3-, CO3-- and RNHCOO-). Model parameters were either taken from literature or regressed:

▪ Mole fraction-based temperature-dependent equilibrium constants and Henry’s law constant for CO2 were taken from the literature;

▪ Volume and surface area parameters, ri and qi, for H2O, MEA, CO2, H3O+ and OH- were taken from literature and those of MEAH+, HCO3-, CO3-- and MEACOO- were regressed;

▪ The binary energy interaction parameters are assumed to be symmetric, therefore considering interaction between all species, a maximum of 81 parameters and 81 is required. Many of them were taken from the literature and the remaining ones were found from regression;

▪ Experimentally determined CO2 partial pressures, total pressures and loadings, as well as N2O solubility data, were used in regression.

The thermodynamic properties of the H2O–MEA system was represented well: total pressure in 2%, activity coefficient and excess enthalpy in 11%, freezing point depression in 3%. The partial pressure of CO2, total pressure and physical CO2 solubility were correlated with an average absolute relative deviation of 24.3, 11.7 and 2.7%, respectively. The estimated heat of absorption and the amine volatility were also found to be satisfactory.

Application to NH3

Thomsen and Rasmussen modeled the NH3 + CO2 + H2O system in the temperature range 0–110°C, pressure up to 10 MPa and the concentration range up to 80 M ammonia Citation[78]. In a later work by Darde et al., the validity of the e-UNIQUAC model was extended up to 150°C by refitting 43 model parameters with a larger experimental database Citation[93].

Applications of the e-NRTL model

Application to MDEA

Zhang et al. used the symmetric e-NRTL model to correlate a wide variety of thermodynamic properties of the system MDEA–H2O–CO2 at temperatures from 313 to 393 K, MDEA concentrations up to 8 M (∼50 wt%), and CO2 loadings up to 1.32 molCO2/molamineCitation[69]. The vapor phase is assumed to be a ternary mixture of water, CO2 and MDEA and the fugacity coefficients of these species were calculated using the perturbed chain statistical associating fluid theory EoS. Eight species (molecular: MDEA, H2O and CO2; cations: H3O+ and MDEAH+; anions: OH-, HCO3- and CO3--) and four chemical reactions are considered in modeling the liquid phase. The activity coefficient of the species in the liquid phase was calculated from the symmetric e-NRTL model Citation[76]. In VLE calculations, the asymmetric mixed-solvent reference state for the molecular solute CO2 was applied. The reference states for aqueous phase chemical equilibrium calculations were pure liquid for the solvents (water and MDEA) and aqueous-phase infinite dilution for the solutes (ionic and molecular). Standard-state chemical potentials as a function of temperature are calculated from physical property data mostly taken from the literature and using thermodynamic relationships. Physical properties of MDEAH+ were obtained from data regression. All molecule–molecule binary parameters and electrolyte–electrolyte binary parameters were defaulted to zero. All molecule–electrolyte binary parameters were defaulted to 8 and -4 for the NRTL model. The nonrandomness factor (R) was fixed at 0.2. The dominant binary NRTL parameters were obtained by regression from the experimental data. Experimental and e-NRTL prediction of CO2 partial pressure of the MDEA–H2O–CO2 system at an MDEA concentration of approximately 8 M are shown in Citation[69]. Prediction using the same model parameters was obtained from Zhang and Chen Citation[69]; the 1986 framework is also shown. The two versions of the model yield similar results at low MDEA concentrations with the difference increasing with increasing MDEA concentration. Slight discrepancy in the prediction of partial pressure of CO2 over MEA + CO2 + H2O by the refined e-NRTL and the original e-NRTL was found by Hessen et al.Citation[94]. Earlier, Austgen et al. applied the unsymmetrical e-NRTL model to MDEA and in blends (MDEA + MEA or DEA) with a thorough discussion of modeling details Citation[20].

Application to AMP, PZ, MEA & NH3

Dash et al. modeled CO2 + AMP + H2O using the unsymmetrical e-NRTL model Citation[34]. Nonrandomness factors for molecule–molecule and molecule–electrolyte have been fixed at 0.2 Citation[34]. The ion-pair–ion-pair parameters were assigned a default value of zero. Molecule–ion-pair, ion-pair molecule and one molecule–molecule parameters were regressed as a function of temperature. The modeling results were compared with the solubility data of their own and those from the literature. Their own data at 2.5, 3.4 and 4.9 M AMP concentration in a temperature range of 298 to 328 K were correlated within 10.5%. This model also predicts the speciation, heat of absorption, pH of CO2-loaded AMP solution and the amine volatility.

The high-pressure CO2 + PZ + H2O system was modeled with the unsymmetrical e-NRTL model by Kadiwala et al.Citation[19] and Dash et al.Citation[60]. Dash et al. adjusted the Henry’s law constants in aqueous solution of PZ for a better prediction of the CO2 partial pressure over aqueous PZ solution. Kadiwala et al. set the ion-pair–ion-pair parameters to zero, took molecule–molecule binary interaction parameters from the literature and regressed three water–electrolyte and five electrolyte–water temperature-independent interaction parameters Citation[19]. A number of data sets including their own high-pressure data were correlated with AAD deviation less than 3% in partial pressure of CO2. The AAD% deviation reported in the work of Kadiwala et al. and Dash et al. in partial pressure of CO2 for data from various literature sources is less than the overall AAD (26%) reported by Bougie et al., who used the Pitzer’s excess model for the same system using selected literature data Citation[90].

Vivier et al. correlated a variety of literature data on the MEA + CO2 + H2O system with the unsymmetric e-NRTL model with a differential evolution algorithm Citation[95]. The symmetric e-NRTL model was used by Que and Chen to satisfactorily represent the thermodynamic properties (VLE, SLE, speciation, heat of solution and heat capacity) of the NH3 + CO2 + H2O system with a temperature up to 473 K, pressure up to 7 MPa, NH3 concentration up to 30 wt% and CO2 loading up to unity Citation[96].

▪ Application of nonrigorous models

The nonrigorous model of Kent and Eisenberg is attractive due to its limited input information requirement, mathematical simplicity and ability to well correlate the partial pressure of CO2 in a limited range of experimental conditions Citation[97]. The equilibrium constants were treated as apparent equilibrium constants:

Equation 2

The non-ideality of the system was overcome by fitting two equilibrium constants (one of them was for carbamate reversion) to the solubility data while the other equilibrium constants are usually taken from the literature as function of temperature. In the original approach, all equilibrium constants were expressed only as function of temperature. Later, one or more equilibrium constants were expressed as function of temperature, amine concentration, loading or partial pressure of CO2 to add further flexibility.

The modified Kent and Eisenberg model from Li and Shen was applied Citation[98], in more recent works, to the CO2 + H2O + MIPA system by Rebolledo-Morales et al.Citation[38], 2-amino-2-methyl-1-propanol + PZ + water + CO2 and by Yang et al.Citation[62], and aqueous 4-(diethylamino)-2-butanol by Sema et al.Citation[99], where the two equilibrium constants are expressed as a function of temperature, amine concentration and loading and, subsequently, the constants ai and bi are obtained from fitting to experimental data:

Equation 3

In the work of Haider et al. for the modeling of the CO2 + H2O + 2-(methylamino)-ethanol system, only one equilibrium constant (carbamate reversion) was expressed as function of temperature, partial pressure and amine concentration, and the rest of the equilibrium constants were considered as function of temperature only Citation[100]. The correlation of partial pressure was deemed to be satisfactory. Gabrielsen et al. considered only one reaction in the aqueous amine system (carbamate formation for MEA and DEA and bicarbonate formation for MDEA) and the partial pressure of CO2 is expressed as a function of mole fraction CO2, loading of CO2 and a combined Henry’s law and equilibrium constant Citation[101]. The parameters in the expression of the equilibrium constant were regressed to experimental data and accurate representation of partial pressure was obtained, although in a limited range of pressure and concentration.

Conclusion

In terms of solubility at low and high pressures, and a low concentration of aqueous sterically hindered amine, AHPD seems to be one of the best solvents. At low pressure and high concentration, an aqueous diamine, AEEA, has higher solubility than other amines but is subject to high rates of degradation. At low pressure and low concentration, an aqueous diamine, 1,6-hexanediamine, has more CO2 capacity than other available aqueous amines. Unfortunately, the environmental impact of these amines remains questionable.

Considering only the flue gas conditions (low-pressure solubility data), it was observed that 1,6-hexanediamine has the highest loading capacity followed by a mixed solvent with 40 wt% MDEA and a task-specific IL (tetramethylammonium glycinate [N1111Gly]). Mixed solvents with sterically hindered amines have the advantages of maximum CO2 loading capacity at lower temperatures and low loading at higher temperatures. Adding physical solvents enhances the solubility at high pressures and reduces the solubility at high temperatures, which is desirable for the absorption/stripping cycle.

Rigorous thermodynamic modeling of a reactive CO2–aqueous alkanolamine system is a challenging task. While the long-range interaction is usually taken care of by an extended Debye–Hückel equation, great varieties exist in modeling the array of short-range interactions.

The nonlocal composition-based models such as the DM model, Pitzer’s excess Gibbs energy model, and local composition-based models such as e-UNIQUAC and e-NRTL and the modified (still nonrigorous) Kent–Eisenberg model are frequently chosen. The nonrigorous models perform well in a limited range of temperature, pressure and loading. The performance of the DM model may deteriorate at higher amine concentration.

The e-NRTL model has been applied to numerous systems. A recently proposed symmetric e-NRTL is thermodynamically consistent for treating an aqueous or non-aqueous system and mixed solvent system. It has also been applied to model VLE, liquid–liquid equilibrium and solid–liquid equilibrium in a single (ammonia) system. Prediction using the unsymmetrical and symmetrical e-NRTL may vary slightly for some systems.

The e-UNIQUAC requires ion-specific parameters (size and shape parameters) in addition to binary interaction parameters and can model complex systems including VLE, liquid–liquid equilibrium and solid–liquid equilibrium. Many assumptions are introduced to reduce the theoretically possible hundreds of binary interaction parameters to a manageable number in the applications of the DM, Pitzer, e-UNIQUAC, e-NRTL, and Fürst and Renon EoS models. Extensive experimental data on pure component property and binary subsystems (H2O + CO2, H2O + amine) are required to enhance the capability of the model. Nuclear magnetic resonance speciation data are useful to verify the model predictive ability for speciation.

Binary interaction parameters are usually transferable from previous work or from similar systems and can be used as an initial guess. EoS models require molecule-specific and ion-specific parameters.

Future perspective

Large-scale carbon capture for flue gases is still not available commercially. The first CO2 capture plant for a 110 MW coal power plant unit is being built in Saskatchewan, Canada, in 2012 for Saskpower. The process is still quite costly and, therefore, there is much room for potential improvement in the design of new solvents. It is for this reason that we have left the door open in our review to some solvents that for now are not considered probable candidates for flue gas application. Many researchers have focused their efforts on mixed or blended amines to overcome the deficiencies of single solvents. Some recent screening studies show the potential of finding aqueous single monoamines, diamines or task-specific ILs for industrial applications. However, the optimal solvent has to pass the test of kinetics, cyclic capacity, energy of regeneration, volatility, degradation, corrosion, toxicity, cost and pilot plant demonstration studies. Future research work is expected to continue to further optimize the currently promising solvents, focused on overcoming their limitations. Research is expected to intensify in dealing with novel cyclic and tertiary amines, amine blends with ILs, amine functionalized ILs or amino acids, in order to find an economical solvent for CO2 capture.

Equilibrium solubility

Study of the solubility of a gas in a liquid when temperature, pressure and concentration of vapor and liquid have no net change over time.

CO2 loading

Ratio of mole of CO2 absorbed per mole of solvent.

Executive summary

Desired properties of a CO2 capture solvent

▪ High specific absorption capacity for the acid gases.

▪ High boiling point and consequently low vapor pressure to prevent solvent loss.

▪ High degree of stability of the molecule in order to prevent decomposition.

▪ Reduced corrosion.

Vapor–liquid equilibrium or equilibrium solubility

▪ Vapor–liquid equilibrium (VLE) data are required for the design and operation of the plants.

▪ VLE data can be used for testing and development of theoretical models, correlations and process design software.

Experimental methods

▪ In CO2 capture research, the existing experimental set-up can be classified into two main experimental methods: analytical and synthetic.

▪ Three different types of VLE cells are used extensively to obtain equilibrium solubility data. Supplementary Table 1 summarizes the experiments conducted during the past decades with the development made in phase analysis.

Experimental data on the solubility of CO2

▪ General advantages and disadvantages of different industrial solvents are reviewed.

▪ Solubility data in a number of industrially important solvents monoethanolamine, diethanolamine, methyldiethanolamine, 2-amino-2-methyl-1-propanol, piperazine and their mixtures and blends with physical solvents are reviewed.

Analysis of the published solubility data

▪ Trends in the effect of pressure and temperature on solubility are analyzed.

▪ Pure, mixed and blended solvents are compared in terms of capacity.

VLE modeling

▪ Both rigorous and nonrigorous models are applied in data correlation of CO2–alkanolamine systems.

▪ Thermodynamic models used by different researchers for different systems are tabulated.

▪ Overview of the models and recent applications are highlighted.

Supplemental material

tcmt_a_10816303_sm0001.docx

Download MS Word (658.1 KB)

Supplementary data

To view the supplementary data that accompany this paper please visit the journal website at: www.future-science.com/doi/suppl/10.4155/CMT.12.47

Financial & competing interests disclosure

The authors have no relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

No writing assistance was utilized in the production of this manuscript.

References

  • Singh P, Niederer JPM, Versteeg GF. Structure and activity relationships for amine based CO2 absorbents – I. Int. J. Greenh. Gas Cont.1,5–10 (2007).
  • Singh P, Niederer JPM, Versteeg GF. Structure and activity relationships from amine-based CO2 absorbents – II. Chem. Eng. Res. Design87,135–144 (2009).
  • Puxity G, Rowland R, Allport A et al. Carbon dioxide post combustion capture: a novel screening study of the carbon dioxide absorption performance of 76 amines. Environ. Sci. Technol.43(16),6427–6433 (2009).
  • Chowdhury FA, Okabe H, Shimizu S, Onoda M, Fujioka Y. Development of novel tertiary amine absorbents for CO2 capture. Energy Procedia1,1241–1248 (2009).
  • Chowdhury FA, Okabe H, Yamada H, Onoda M, Fujioka Y. Synthesis and selection of hindered new amine absorbents for CO2 capture. Energy Procedia4,201–208 (2011).
  • Porcheron F, Gibert A, Jacquin M et al. High throughput screening of amine thermodynamic properties applied to post-combustion CO2 capture process evaluation. Energy Procedia4,15–22 (2011).
  • Porcheron F, Gibert A, Mougin P, Wender A. High throughput screening of CO2 solubility in aqueous monoamine solutions. Environ. Sci. Technol.45(6),2486–2492 (2011).
  • Rochelle GT, Bishnoi S, Chi S, Dang H, Santos J. Research Needs for CO2 Capture from Flue Gas by Aqueous Absorption/Stripping. US Department of Energy, Pittsburgh, PA, USA (2000).
  • Anufrikov YA, Kuranov GL, Smirnova NA. Solubility of CO2 and H2S in alkanolamine-containing aqueous solutions. Russ. J. Appl. Chem.80(4),515–527 (2007).
  • Dohrn R. Thermophysical properties-industrial directions. Fluid Phase Equilib.194–197,15–29 (2002).
  • Jou F-Y, Mather AE, Otto FD. The solubility of CO2 in a 30 mass percent monoethanolamine solution. Can. J. Chem. Eng.73,140–147 (1995).
  • Sidi-Boumedine R, Horstmann S, Fisher K. Experimental determination of carbon dioxide solubility data in aqueous alkanolamine solutions. Fluid Phase Equilib.218(1),85–94 (2004).
  • Tong D, Trusler JPM, Maitland GC, Gibbins J, Fennell PS. Solubility of carbon dioxide in aqueous solution of monoethanolamine or 2-amino-2-methyl-1-propanol: experimental measurements and modeling. Int. J. Greenh. Gas Cont.6,37–47 (2012).
  • Dohrn R, Brummer G. Thermophysical properties-industrial directions. Fluid Phase Equilib.194–197,15–29 (1995).
  • Weir RD, de Loos TWW. Measurement of Thermodynamic Properties of Multiple Phases: Experimental Thermodynamics (Volume 7). Elsevier, Amsterdam, The Netherlands (2005).
  • Ma’mun S, Nilsen R, Svendsen HF. Solubility of carbon dioxide in 30 mass% monoethanolamine and 50 mass% methyldiethanolamine solutions. J. Chem. Eng. Data50(2),630–634 (2005).
  • Lawson JD, Garst AW. Gas sweetening data: equilibrium solubility of hydrogen sulfide and carbon dioxide in aqueous monoethanolamine and aqueous diethanolamine solutions. J. Chem. Eng. Data21,20–30 (1976).
  • Addicks J, Owren GA. Solubility of carbon dioxide and methane in aqueous methyldiethanolamine solutions. J. Chem. Eng. Data47(4),855–860 (2002).
  • Kadiwala S, Rayer AV, Henni A. High pressure solubility of carbon dioxide (CO2) in aqueous piperazine solutions. Fluid Phase Equilib.292(1–2),20–28 (2010).
  • Austgen DM, Rochelle GT, Chen CC. Model of vapor–liquid equilibria for aqueous acid gas–alkanolamine systems. 2. Representation of H2S and CO2 solubility in aqueous MDEA and CO2 solubility in aqueous mixtures of MDEA with MEA or DEA. Ind. Eng. Chem. Res.30(3),543–555 (1991).
  • Ma’mun S, Jakobsen JP, Svendsen HF. Experimental and modeling study of the solubility of carbon dioxide in aqueous 30 mass% 2-((2-aminoethyl) amino) ethanol solution. Ind. Eng. Chem. Res.45(8),2505–2512 (2006).
  • Daneshvar N, Moattar MTZ, Abdi MA, Aber S. Carbon dioxide equilibrium absorption in the multi-component systems of CO2 + TIPA + MEA + H2O, CO2 + TIPA + PZ + H2O and CO2 + TIPA + H2O at low partial pressures: experimental solubility data, corrosion study and modeling with artificial neural network. Sep. Purf. Technol.37,135–147 (2004).
  • Mason JW, Dodge BF. Equilibrium absorption of carbon dioxide by solutions of the ethanolamines. Trans. Am. Inst. Chem. Eng.32,27–48 (1936).
  • Ji L, Miksche SJ, Rimpf LM, Farthing GA. CO2 chemical solvent screening. Presented at: 8th Annual Conference on Carbon Conference and Sequestration. Pittsburgh, PA, USA, 4–7 May 2009.
  • Rogers WJ, Bullin JA, Davidson RR, Frazier RE, Marsh KN. FTIR method for VLE measurements of acid–gas–alkanolamine systems. AIChE J43(12),3223–3231 (1997).
  • Hartono A, da Silva EF, Grasdalen H, Svendsen HF. Qualitative determination of species in DETA–H2O–CO2 system using 13C NMR spectra. Ind. Eng. Chem. Res.46(1),249–254 (2007).
  • Souchon V, de Oliveira Aleixo M, Delpoux O et al.In situ determination of species distribution in alkanolamines–H2O–CO2 systems by Raman spectroscopy. Energy Procedia4,554–561 (2011).
  • Park MK, Sandall OC. Solubility of carbon dioxide and nitrous oxide in 50 mass methyldiethanolamine. J. Chem. Eng. Data46(1),166–168 (2001).
  • Chakma A, Tontiwachwuthikul P. Designer solvents for energy efficient CO2 separation from flue gas streams. In: Greenhouse Gas Control Technologies. Eliasson B, Riemer PWF, Wokaum A (Eds). Elsevier, Amsterdam, The Netherlands, 35–42 (1999).
  • Martin JL, Otto FD, Mather AE. Solubility of hydrogen sulfide and carbon dioxide in a diglycolamine solution. J. Chem. Eng. Data23,163–164 (1978).
  • Jou FJ, Mather AE, Otto FD. Solubility of H2S and CO2 in aqueous methyldiethanolamine solutions. Ind. Eng. Chem. Process Des. Dev.21,539–544 (1982).
  • Jou FY, Mather AE, Otto FD. The solubility of CO2 in a 30 mass percent monoethanolamine solution. Can. J. Chem. Eng.73,140–147 (1995).
  • Zoghi AT, Feyzi F, Zarrinpashneh S. Equilibrium solubility of carbon dioxide in a 30wt. % aqueous solution of 2-((2-aminoethyl) amino) ethanol at pressures between atmospheric and 4400 kPa: an experimental and modeling study. J. Chem. Thermodyn.44,66–74 (2012).
  • Dash SK, SamantaAN, Bandyopadhyay SS. (Vapour + liquid) equilibria (VLE) of CO2 in aqueous solutions of 2-amino-2-methyl-1-propanol: new data and modeling using eNRTL-equation. J. Chem. Thermodyn.43,1278–1285 (2011).
  • Park JY, Yoon SJ, Lee H. Density, viscosity and solubility of CO2 in aqueous solutions of 2-amino-2-hydroxymethyl-1,3-propanediol. J. Chem. Eng. Data47,970–973 (2002).
  • Baek JI, Yoon JH. Solubility of carbon dioxide in aqueous solutions of 2-amino-2-methyl-1,3-propanediol. J. Chem. Eng. Data43,635–637 (1998).
  • Carroll JJ, Slupsky JD, Mather AE. The solubility of carbon dioxide in water at low pressure. J. Phys. Chem. Ref. Data20,1201–1209 (1991).
  • Rebolledo-Morales MA, Rebolledo-Libreros ME, Trejo A. Equilibrium solubility of CO2 in aqueous solutions of 1-amino-2-propanol as function of concentration, temperature and pressure. J. Chem. Thermodyn.43,690–695 (2011).
  • Rayer AV. Screening of solvents for CO2 capture: kinetics, solubility and calorimetric studies. PhD Thesis, University of Regina, Regina, Canada (2012).
  • Jou FY, Otto FD, Mather AE. Solubility of H2S, CO2 and their mixtures in an aqueous solution of 2-piperidineethanol and sulfolane. J. Chem. Eng. Data43,409–412 (1998).
  • Park JY, Yoon SJ, Lee H et al. Solubility of carbon dioxide in aqueous solutions of 2-amino-2-ethyl-1,3-propanediol. Fluid Phase Equilib.202,359–366 (2002).
  • Chauhan RK, Yoon SJ, Lee H et al. Solubilities of carbon dioxide in aqueous solutions of triisopropanolamine. Fluid Phase Equilib.208,239–245 (2003).
  • Lee JI, Otto FD, Mather AE. Equilibrium between carbon dioxide and aqueous monoethanolamine solutions. J. Appl. Chem. Biotechnol.26,541–549 (1976).
  • Isaacs EE, Otto FD, Mather AE. Solubility of hydrogen sulfide and carbon dioxide in an aqueous diisopropanolamine solution. J. Chem. Eng. Data22,77–73 (1977).
  • Jou FJ, Mather AE, Otto FD. Solubility of H2S and CO2 in aqueous methyldiethanolamine solutions. Ind. Eng. Chem. Process Des. Dev.21,539–544 (1982).
  • Silkenbäumer D, Rumpf B, Lichtenthaler RN. Solubility of carbon dioxide in aqueous solutions of 2-amino-2-methyl-1-propanol and n-methyldiethanolamine and their mixtures in the temperature range from 313 to 353 K and pressures up to 2.7 MPa. Ind. Eng. Chem. Res.37,3133–3141 (1998).
  • Maneeintr K, Idem RO, Tontiwachwuthikul P, Wee AGH. Synthesis, solubilities and cyclic capacities of amino alcohols for CO2 capture from flue gas streams. Energy Procedia1,1327–1334 (2009).
  • Singh P, Brilman DWF, Groeneveld MJ. Solubility of CO2 in aqueous solution of newly developed absorbents. Energy Procedia1,1257–1264 (2009).
  • Haji-Sulaiman MZ, Aroua MK, Benamor A. Analysis of equilibrium data of CO2 in aqueous solutions of diethanolamine (DEA), Methyldiethanolamine (MDEA) and their mixtures using the modified Kent Eisenberg Model. Trans IChemE Part A76,961–968 (1998).
  • Cheng MD, Caparanga AR, Soriano AN, Li MH. Solubility of CO2 in the solvent system (water + monoethanolamine + triethanolamine). J. Chem. Thermodyn.42,342–347 (2010).
  • Ermatchkov V, Pérez-Salado Kamps Á, Speyer D, Maurer G. Solubility of carbon dioxide in aqueous solutions of piperazine in the low gas loading region. J. Chem. Eng. Data51,1788–1796 (2006).
  • Idriss AI. Experimental equilibrium between carbon dioxide and aqueous solutions of diglycolamine and morpholine. MSc Thesis. King Fahd University of Petroleum & Minerals, Dhahran, Saudi Arabia (1989).
  • Kumar G, Kundu M. Vapor–liquid equilibrium of CO2 in aqueous solutions of n-methyl-2-ethanolamine. Can. J. Chem. Eng.90,627–630 (2011).
  • Dong L, Chen J, Gao G. Solubility of carbon dioxide in aqueous solutions of 3-amino-1-propanol. J. Chem. Eng. Data55,1030–1034 (2010).
  • Jenab MH, Abdi MA, Najibi SH, Vahidi M, Matin NS. Solubility of carbon dioxide in aqueous mixtures of n-methyldiethanolamine + piperazine + sulfolane. J. Chem. Eng. Data50,583–586 (2005).
  • Shen KP, Li MH. Solubility of carbon dioxide in aqueous mixtures of monoethanolamine with methyldiethanolamine. J. Chem. Eng. Data37(1),96–100 (1992).
  • Murrieta GF, Rebolledo LME, Romero-Martínez A, Trejo A. Solubility of CO2 in aqueous mixtures of diethanolamine with methyl diethanolamine and 2-amino-2-methyl-1-propanol. Fluid Phase Equilib.150–151,721–729 (1998).
  • Jenab MH, Vahidi M, Mehrabi M. Solubility of carbon dioxide in aqueous mixtures of DIPA + MDEA and DIPA + PZ solutions. J. Chin. Chem. Soc.53,283–286 (2006).
  • Mondal MK. Solubility of carbon dioxide in an aqueous blend of diethanolamine and piperazine. J. Chem. Eng. Data54,2381–2385 (2009).
  • Dash SK., Samanta A, Samanta AN, Bandyopadhyay SS. Vapor liquid equilibria of carbon dioxide in dilute and concentrated aqueous solutions of piperazine at low to high pressure. Fluid Phase Equilib.300,145–154 (2011).
  • Chung PY, Soriano AN, Leron RB, Li MH. Equilibrium solubility of carbon dioxide in the amine solvent system of (triethanolamine + piperazine + water). J. Chem. Thermodyn.42,802–807 (2010).
  • Yang ZY, Soriano AN, Caparanga AR, Li MH. Equilibrium solubility of carbon dioxide in (2-amino-2-methyl-1-propanol + piperazine + water). J. Chem. Thermodyn.42,659–665 (2010).
  • Bougie F, Iliuta MC. CO2 absorption into mixed aqueous solutions of 2-amino-2-hydroxymethyl-1,3-propanediol and piperazine. Ind. Eng. Chem. Res.49,1150–1159 (2010).
  • Bishnoi S, Rochelle GT. Absorption of carbon dioxide into aqueous piperazine: reaction kinetics, mass transfer and solubility. Chem. Eng. Sci.55,5531–5543 (2000).
  • Lee JI, Otto FD, Mather AE. The solubility of H2S and CO2 in aqueous monoethanolamine solutions. Can. J. Chem. Eng.52,803–805 (1974).
  • Feng Z, Jing-Wen M, Zheng Z, You-Ting W, Zhi-Bing Z. Study on the absorption of carbon dioxide in high concentrated MDEA and ILs solutions. Chem. Eng. J.181–182,222–228 (2011).
  • Taib MM, Murugesan T. Solubilities of CO2 in aqueous solutions of ionic liquids (ILs) and monoethanolamine (MEA) at pressures from 100 to 1600 kPa. Chem. Eng. J. doi:10.1016/j.cej.09.048 (2011) (Epub).
  • Zhao Y, Zhang X, Zeng S et al. Density, viscosity and performances of carbon dioxide capture in 16 absorbents of amine + ionic liquid + H2O, ionic liquid + H2O and amine + H2O systems. J. Chem. Eng. Data55(9),3513–3519 (2010).
  • Zhang Y, Chen CC. Thermodynamic modeling for CO2 absorption in aqueous MDEA solution with electrolyte NRTL model. Ind. Eng. Chem. Res.50(1),163–175 (2011).
  • Edwards TJ, Newman J, Prausnitz JM. Thermodynamics of aqueous solutions containing volatile weak electrolytes. AIChE J.21(2),248–259 (1975).
  • Desmukh R, Mather AE. A mathematical model for equilibrium solubility of hydrogen sulphide and carbon dioxide in aqueous alkanolamines solutions. Chem. Eng. Sci.36(2),355–362 (1981).
  • Pitzer KS. Thermodynamics of electrolytes. I. Theoretical basis and general equations. J. Phys. Chem.77,268–277 (1973).
  • Chen CC, Britt HI, Boston JF, Evans LB. Local composition model for excess Gibbs energy of electrolytes systems, Part 1: Single solvent, single completely dissociated electrolyte systems. AIChE J.28(4),588–596 (1982).
  • Chen C-C, Evans LB. A local composition of model for the excess Gibbs energy of aqueous electrolyte systems. AIChE J.32(3),444–454 (1986).
  • Bollas GM, Chen C-C, Barton PI. Refined electrolyte-NRTL model: Activity coefficient expressions for application to multi-electrolyte systems. AIChE J.54,1608–1624 (2008).
  • Song Y, Chen C-C. Symmetric electrolyte nonrandom two-liquid activity coefficient model. Ind. Eng. Chem. Res.48,7788–7797 (2009).
  • Sander B, Fredenslund A, Rasmussen P. Calculation of vapour–liquid equilibria in mixed solvent/salt systems using an extended UNIQUAC equation. Chem. Eng. Sci.41,1171 (1986).
  • Thomsen K, Rasmussen P. Modeling of vapor–liquid–solid equilibrium in gas–aqueous electrolyte system. Chem. Eng. Sci.54,1787 (1999).
  • Fürst W, Renon H. Representation of excess properties of electrolyte solutions using a new equation of state. AIChE J.39(2),335–343. (1993).
  • Derks PWJ, Dijkstra HBS, Hogendoorn JA, Versteeg GF. Solubility of carbon dioxide in aqueous piperazine solutions. AIChE J.51,2311–2327 (2005).
  • Dowell NM, Llovell F, Adjiman CS, Jackson G, Galindo A. Modeling the fluid phase behavior of carbon dioxide in aqueous solutions of monoethanolamine using transferable parameters with the SAFT-VR approach. Ind. Eng. Chem. Res.49,1883–1899 (2010).
  • Newman SA. Thermodynamics of Aqueous Systems with Industrial Applications. ACS Symposium Series. American Chemical Society, Washington, DC, USA (1980).
  • Georgios MK, Georgios KF. Thermodynamic Models for Industrial Applications: from Classical and Advanced Mixing Rules to Association Theories. John Wiley and Sons Ltd, Hoboken, NJ, USA (2010).
  • Kim I, Hoff KA, Hessen ET, Haug-Warberg T, Svendsen HF. Enthalpy of absorption of CO2 with alkanolamine solutions predicted from reaction equilibrium constants. Chem. Eng. Sci.64(9),2027–2038 (2009).
  • Penttilä A, Dell’Era C, Uusi-Kyyny P, Alopaeus V. The Henry’s law constant of N2O and CO2 in aqueous binary and ternary amine solutions (MEA, DEA, DIPA, MDEA and AMP). Fluid Phase Equilib.311,59–66 (2011).
  • Pahlavanzadeh H, Nourani S, Saber M. Experimental analysis and modeling of CO2 solubility in AMP (2-amino-2-methyl-1-propanol) at low CO2 partial pressure using the models of Deshmukh–Mather and the artificial neural network. J. Chem. Thermodyn.43,1775–1783 (2011).
  • Hartono A, Hoff KA, Mejdell T, Svendsen HF. Solubility of carbon dioxide in aqueous 2.5 m of diethylenetriamine (DETA) solution. Energy Procedia4,179–186 (2011).
  • Dicko M, Coquelet C, Jarne C, Northrop S, Rochon D. Acid gases partial pressures above a 50 wt% aqueous methyldiethanolamine solution: experimental work and modeling. Fluid Phase Equilib.289(2),99–109 (2010).
  • Kumar G, Mondal TK, Kundu M. Solubility of CO2 in aqueous blends of (diethanolamine + 2-amino-2-methyl-1-propanol) and (diethanolamine +n-methyldiethanolamine). J. Chem. Eng. Data57,670–680 (2012).
  • Bougie F, Iliuta MC. CO2 absorption in aqueous piperazine solutions: experimental study and modeling. J. Chem. Eng. Data56(4),1547–1554 (2011).
  • Faramarzi L, Kontogeorgis GM,Thomsen K, Stenby EH. Extended UNIQUAC model for thermodynamic modeling of CO2 absorption in aqueous alkanolamine solutions. Fluid Phase Equilibria282(2),121–132 (2009).
  • Aronu UE, Gondal S, Hessen ET et al. Solubility of CO2 in 15, 30, 45 and 60 mass% MEA from 40 to 120°C and model representation using the extended UNIQUAC framework. Chem. Eng. Sci.66,6393–6406 (2011).
  • Darde LV, van Well WJM, Stenby EH, Thomsen K. Modeling of carbon dioxide absorption by aqueous ammonia solutions using the extended UNIQUAC model. Ind. Eng. Chem. Res.49,12663–12674 (2010).
  • Hessen ET, Haug-Warberg T, Svendsen HF. The refined e-NRTLmodel applied to CO2–H2O–alkanolamine systems. Chem. Eng. Sci.65,3638–3648 (2010).
  • Vivier J, Kamalpour S, Mehablia A. NRTL with differential evolution algorithm modeling of CO2–MEA–H2O system equilibria. J. Thermodyn. Catal.3,111 (2012).
  • Que H, Chen C-C. Thermodynamic modeling of the NH3-CO2-H2O system with electrolyte NRTL model. Ind. Eng. Chem. Res.50,11406–11421 (2011).
  • Kent RL, Eisenberg B. Better data for amine treating. Hydrocarbon Proc.55,87–89 (1976).
  • Li MH, Shen KP. Calculation of equilibrium solubility of carbon dioxide in aqueous mixtures of monoethanolamine with methyldiethanolamine. Fluid Phase Equilib.85,129–140 (1993).
  • Sema T. Naami A, Idem R, Tontiwachwuthikul P. Correlations for equilibrium solubility of carbon dioxide in aqueous 4-(diethylamino)-2-butanol solutions. Ind. Eng. Chem. Res.50(24),14008–14015 (2011).
  • Haider HAM, Yusoff R, Aroua MK. Equilibrium solubility of carbon dioxide in 2 (methyamino) ethanol. Fluid Phase Equilib.303(2),162–167 (2011).
  • Gabrielsen J, Michelsen ML, Stenby EH, Kontogeorgis GM. A model for estimating CO2 solubility in aqueous alkanolamines. Ind. Eng. Chem. Res.44,3348–3354 (2005).

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.