Carbon dioxide solubility in aqueous potassium salt solutions of l-proline and dl-α-aminobutyric acid at high pressures

https://doi.org/10.1016/j.jct.2014.12.010Get rights and content

Highlights

  • CO2 solubility in aqueous potassium salt solutions of l-proline and dl-α-aminobutyric acid were studied.

  • The CO2 partial pressures studied was up to 1000 kPa.

  • The temperatures studied were (313.2, 333.2, 353.2) K.

  • The measured data were represented satisfactorily by using the applied correlations.

  • The CO2 absorption capacity of the studied systems was high and comparable with monoethanolamine.

Abstract

In the present work, the solubility of CO2 in aqueous solutions of potassium prolinate (KPr) and potassium α-aminobutyrate (KAABA) was measured at temperatures (313.2, 333.2, and 353.2) K and CO2 partial pressures up to 1000 kPa for amino acid salt concentrations: KPr, w = (7.5, 14.5, and 27.4 wt%) and KAABA, w = (6.9, 13.4, and 25.6 wt%). It was found that the CO2 absorption capacities of the studied amino acid salt systems were considerably high and comparable with that of industrially important alkanolamines including monoethanolamine. The CO2 loadings in aqueous potassium α-aminobutyrate at high pressures were also found to be generally higher than the loadings in aqueous potassium prolinate. A modified Kent–Eisenberg model was applied to correlate the CO2 solubility in the amino acid salt solution as function of CO2 partial pressure, temperature, and concentration. The model gave good representation of the (vapour + liquid) equilibrium data obtained for the amino acid salt systems studied, and provided accurate predictions of the solubility.

Introduction

The leading technology today for post-combustion CO2 capture is based on the reversible chemical absorption of CO2 in amine-based solvents. The process most commonly utilises aqueous solutions of alkanolamines particularly monoethanolamine (MEA). The wide use of MEA as CO2 absorbent is due to its fast CO2 reaction kinetics, low solvent cost, high alkalinity, and other desirable characteristics. However, the use of MEA also entails a number of process limitations and other disadvantages. Aqueous MEA is corrosive at high concentrations, forms degradation products due to long exposure with CO2 and O2, and exhibits high vapour pressure causing significant solvent losses during processing. Also, it has a high heat of reaction with CO2, which leads to high energy requirement during solvent regeneration [1], [2]. Thus, development of new and better solvents than aqueous MEA has always been necessary.

Amino acid salts are proposed as suitable alternatives to aqueous alkanolamine solvents [3], [4], [5], [6]. Due to the presence of amino functional groups in their molecules, amino acid salts, like alkanolamines, have fast reactivity and appreciable absorption capacity for CO2 [4], [7], [8], [9], [10], [11]. The reaction kinetics of potassium salts of some amino acids were found to be even faster than MEA’s [12]. Additional advantages of aqueous amino acid salt solutions include low volatility and resistance to oxidative degradation [13]. They also have surface tension and viscosity that are similar to those of water, which makes their application more practical [14], [15], [16], [17]. It has also been demonstrated that the utilisation of this type of solvents would lead to a more environment-friendly and energy-effective separation process for CO2 capture [18].

In the present work, we investigated the CO2 absorption capacities of two amino acid salts that are potential CO2 absorbents—the potassium salts of l-proline and dl-α-aminobutyric acid. Proline is a secondary amino acid with a distinctive cyclic structure, which includes an amine group while dl-α-aminobutyric acid is a sterically-hindered amino acid having an ethyl group substituent. The potassium salt of l-proline (KPr) is proposed by several authors as a candidate absorbent for CO2 capture due to its fast reaction rate and high absorption capacity for CO2 [10], [12], [19]. van Holst et al. [12] examined the absorption kinetics of CO2 in potassium salts of several amino acids including 6-aminohexanoic acid, β-alanine, l-arginine, l-glutamic acid, dl-methionine, l-proline, and sarcosine at T = 298 K. They reported that the potassium salt of l-proline (and sarcosine) has relatively high apparent rate constant and low pKa making it a promising CO2 absorbent. In a more recent study, Paul and Thomsen [10], used a zwitterionic mechanism, which generally applies for the reaction of CO2 with primary and secondary alkanolamines, to describe the kinetics of reaction of CO2 with KPr. They found that the overall reaction rate constant for the latter is much higher than that of potassium threonate (KThr) and potassium taurate (KTau). The solvent has also been shown to have higher absorption capacity compared with the benchmark alkanolamine, MEA, and (vapour + liquid) equilibrium (VLE) data for low CO2 partial pressures (up to 30 kPa) have been reported [5], [19]. On the other hand, the potassium salt of dl-α-aminobutyric acid (KAABA) has been found to have a significantly higher net cyclic capacity compared with MEA and potassium salts of some other amino acids [5]. The results of the study suggest that KAABA could be an energy-efficient alternative to MEA. However, except for the results of the (absorption + desorption) experiments at 15 kPa presented by Song et al. [5], no other VLE data has been reported for CO2 absorption in KAABA.

The objective of this work was to provide VLE data for the absorption of CO2 in aqueous KPr and aqueous KAABA solutions at high pressures. The solubility of CO2 in the solvents was measured at temperatures (313.2, 333.2, and 353.2) K and pressures up to 1000 kPa. In addition, a modified Kent–Eisenberg model [20] was used to represent the correlation of the measured CO2 solubility (αCO2, expressed as mole CO2 per mole amine) with CO2 partial pressure, temperature, and amino acid salt (AAS) concentration.

Section snippets

Chemicals

The aqueous potassium salt solutions of l-proline (Pr, mass fraction purity > 0.99), dl-α-aminobutyric acid (AABA, purity > 99 wt%), and sarcosine (Sr, mass fraction purity > 0.98) were prepared by adding an equimolar amount of KOH (pellets, mass fraction purity > 0.85) to the amino acid in a volumetric flask with deionised-distilled water. The description and sources of all chemicals used, including those used in the validation experiments (Sr and MEA), are listed in table 1. They were used without

Validation of experimental method

To validate the experimental method used in this work, CO2 solubility in 15.3 wt% MEA solution at T = 313.2 K and 42.4 wt% potassium sarcosinate solution (KSr) at T = 353.2 K were measured. The VLE values obtained are given in TABLE 2, TABLE 3 along with those reported by Lee et al. [24] and Park et al. [25] for CO2–MEA–H2O and Kang et al. [26] for CO2–KSr–H2O systems, respectively. Graphical representations of the results are also provided in figure 1 showing that the experimental values are in

Modelling

In this work, we used a modified form of the Kent–Eisenberg model proposed by Li and Shen [20] to correlate the CO2 loading in the aqueous amino acid salt solution with temperature, CO2 partial pressure, and AAS concentration. This model allows, with simplicity, the calculation of CO2 solubility in amine solvents using equilibrium constant expressions derived from equilibrium reactions, which are assumed to occur between CO2 and the amine. It has been shown that the absorption of CO2 in aqueous

Conclusions

The CO2 solubility in aqueous potassium salts of l-proline and dl-α-aminobutyric acid were measured at temperatures (313.2, 333.2, and 353.2) K and CO2 partial pressures up to 1000 kPa for amino acid salt concentrations (0.5, 1.0, and 2.0) M. Results show that CO2 loadings in the solvents increase with increasing CO2 partial pressure and decrease as temperature increases. It was also found that the absorption capacities of the amino acid salt systems studied are in the same order of magnitude as

Acknowledgement

This research was supported by Grant MOST-103-2221-E-033-067-MY3 from the Ministry of Science and Technology of the Republic of China. We also thank Chung Yuan Christian University for financial support.

References (31)

  • A.F. Portugal et al.

    Chem. Eng. Sci.

    (2009)
  • H.-J. Song et al.

    Int. J. Greenhouse Gas Control

    (2012)
  • S. Paul et al.

    Int. J. Greenhouse Gas Control

    (2012)
  • A.F. Portugal et al.

    Chem. Eng. Sci.

    (2008)
  • S.S. Navarro et al.

    J. Chem. Thermodyn.

    (2014)
  • L.A. Tirona et al.

    J. Chem. Thermodyn.

    (2014)
  • E. Sanchez-Fernandez et al.

    Energy Proc.

    (2013)
  • M.-H. Li et al.

    Fluid Phase Equilib.

    (1993)
  • M.-D. Cheng et al.

    J. Chem. Thermodyn.

    (2010)
  • P.-Y. Chung et al.

    J. Chem. Thermodyn.

    (2010)
  • J.-Y. Park et al.

    Fluid Phase Equilib.

    (2002)
  • A.L. Kohl et al.

    Gas Purification

    (1997)
  • R. Idem et al.

    Ind. Eng. Chem. Res.

    (2006)
  • P.H.M. Feron et al.
    (2005)
  • J. van Holst, P.P. Politiek, J.P.M. Niederer, G.F. Versteeg, CO2 Capture from flue Gas using Amino Acid Salt Solutions,...
  • Cited by (33)

    • Solubility of N<inf>2</inf>O and CO<inf>2</inf> in nonaqueous blends of 2-(methylamino)ethanol and 2-methoxyethanol

      2021, Journal of Molecular Liquids
      Citation Excerpt :

      In this case, the model underpredicts the CO2 solubility at low temperature-high PCO2 conditions. It was also mentioned this issue in several modified Kent-Eisenberg model for aqueous amine/amino acid salt-CO2 systems, showing the AARDs in the range of 7.0–22.0% [32,41–44]. The concentration profiles of the different species in the system for 39.1 mass% MAE at 333.2 K were also evaluated using the proposed model, as shown in Fig. 5.

    • Three-phase hybrid facilitated transport hollow fiber membranes for enhanced CO<inf>2</inf> separation

      2020, Applied Materials Today
      Citation Excerpt :

      The presence of these species formed upon interaction with CO2 with the mobile carrier (represented in Scheme 2B) was identified by NMR studies (Figure S5). Studies indicate that the formation of carbonate/bicarbonate species is more favoured in ProK than the carbamate intermediate due to the steric hindrance effect [49,50]. Increasing the content of ProK mobile carrier led to increasing CO2 permeance from 407 GPU to 730 GPU at 20 wt% with no significant change in CO2/N2 separation factor, as seen in Fig. 7.

    View all citing articles on Scopus
    View full text