Carbon Dioxide Absorption into Aqueous Potassium Salt Solutions of

Aug 1, 2016 - The calculated values from two models agree well with the experimental data within 7.5% AAD. The Supporting Information is available fre...
1 downloads 0 Views 2MB Size
Article pubs.acs.org/EF

Carbon Dioxide Absorption into Aqueous Potassium Salt Solutions of Arginine for Post-Combustion Capture Shufeng Shen* and Ya-nan Yang School of Chemical and Pharmaceutical Engineering, Hebei University of Science and Technology, Shijiazhuang 050018, P.R. China S Supporting Information *

ABSTRACT: The mass transfer and kinetics of CO2 absorption into aqueous potassium argininate (ArgK) solutions were investigated using a wetted-wall column at concentrations ranging from 0.26 to 1.23 M and temperatures from 298 to 333 K. Absorption rates of CO2 were compared by the absorption flux and a fall-in-pressure method among basic amino acid salts (AAS) and two amine standards for capture technology. Results showed that ArgK has comparable kinetics toward CO2 with the standard monoethanolamine (MEA) but lower kinetics than piperazine (PZ). The kinetic results were interpreted using two models based on zwitterion mechanism and termolecular mechanism for the reactive absorption of CO2. It was found that the partial reaction orders vary from 1.25 to 1.60 with respect to ArgK. Under the pseudo-first-order regime, the overall reaction rate constants were obtained and the individual reaction rate constants in these models were derived. The calculated values from two models agree well with the experimental data within 7.5% AAD. summarized.14,25 Lysine, histidine, and arginine have been investigated as promising promoters for aqueous potassium carbonate solutions in our studies.7 We have recently investigated the detailed kinetics of CO2 absorption into aqueous potassium lysinate (LysK) and potassium histidine (HisK) in a wetted wall column.25,27 It was found that LysK has higher chemical reactivity toward CO2 than HisK and the industrial standard MEA. It was also noted that these amino acid salts possess α-amino group with similar pKa but quite different side chains. The effect of side chains with different pKa on CO2 absorption is not clear. Arginine contains a primary αamino group (pKa 9.0) and a basic side chain with a guanidinium group (pKa 12.1),7,28 as shown in Figure 1. The CO2 solubility in and some physical properties of aqueous potassium argininate solutions were measured.29,30 For aqueous arginine with equimolar KOH neutralization, the pH of solutions is usually above 13.0, part of them with uncharged side chains are expected. However, kinetic information of CO2 absorption in these solutions is still rare in literature. In this study, the absorption rates of CO2 were initially compared among five absorbents by absorption flux using the wetted-wall column and a fall-in-pressure method for a stirredcell absorber. Molecular structures of the studied AAS and amines for comparison are given in Figure 1.The identification of the species in the CO2-loaded solution was also obtained from 13C NMR spectroscopic analysis. The kinetics of CO2 absorption into aqueous ArgK solutions were investigated at concentrations (0.26−1.23 M) and temperatures (298−333 K) using a wetted wall column. Under the pseudo-first-order regime, the reaction rate parameters were correlated from the experimental data using zwitterion and termolecular mechanisms.

1. INTRODUCTION With global demands for energy from fossil fuels expected to rise, carbon dioxide (CO2) emissions from fossil fuel-fired power plants will produce increasingly worrying environmental impacts, such as global warming. Accordingly, intensive researches on CO2 capture and storage have been conducted to reduce CO2 emissions into the atmosphere.1 Postcombustion CO2 capture is currently the only and promising method. It can be easily integrated into existing power plants with few process changes. The most established technique for capturing CO2 is solvent absorption in which aqueous amine-based solvents, typically monoethanolamine (MEA), react chemically with CO2 in the flue gas to form carbamates and/or bicarbonates.2,3 This process has already been in commercial use and is generally effective in postcombustion capture plants where the CO2 partial pressures in flue gas are very low (about 5−15 kPa). However, the use of these solvents suffers several drawbacks: high energy requirement for solvent regeneration, toxicity, corrosiveness, easy thermal and oxidation degradation, and high volatility.3 Due to having same reactive groups as alkanolamines, amino acid salts (AAS) can be expected to have similar absorption kinetics with CO2. They also have more superior physicochemical properties than alkanolamines, such as low vapor pressure, better resistance to oxidative degradation, and no environmental or toxic issues.4−7 Therefore, aqueous alkaline salts of amino acids have been regarded as possible replacements to the conventional solvents for CO2 capture. In recent years, there have already been reported regarding their physicochemical properties and the reaction kinetics of CO2 absorption in aqueous alkaline salts of different amino acids, such as of glycine,5,6,8−13 taurine,6,14,15 proline,4,10,14,16,17 threonine,18,19 sarcosine,20,21 alanine,22,23 lysine,24−26 and histidine.26,27 In our previous work, a comprehensive list of aqueous AAS proposed for CO2 capture and the kinetic data for each salt with CO 2 available in the literature were © 2016 American Chemical Society

Received: May 6, 2016 Revised: July 25, 2016 Published: August 1, 2016 6585

DOI: 10.1021/acs.energyfuels.6b01092 Energy Fuels 2016, 30, 6585−6596

Article

Energy & Fuels

Figure 1. Molecular structures of the studied absorbents. MEA (pKa: 9.50): monoethanolamine; PZ (pKa: 5.33, 9.73): piperazine; LysK (pKa: 9.16, 10.67): potassium lysinate; HisK (pKa: 9.09, 6.04): potassium histidinate; ArgK (pKa: 9.00, 12.10): potassium argininate. The pKa is for the α-NH2/ NH group and groups in side chain, respectively.

HCO2

2. THEORETICAL BACKGROUND 2.1. Mass Transfer with Chemical Reactions in Aqueous AAS Solution. The absorption flux of CO2 into ArgK solutions can be described by the following equation:7,25 * )= NCO2 = K G(PCO2,b − PCO 2

* PCO2,b − PCO 2 1 kg′

+

1 kg



kovDCO2

+

⎛ HCO ⎞ 2 ⎟⎟ = log⎜⎜ ⎝ HCO2,H2O ⎠

E∞ =

(5)

(6)

∑ (hion + hg)Cion (7)

where HCO2 and HCO2,H2O are, respectively, the Henry’s constants of gases in the ArgK solutions and water. hion is the ion-specific parameter for ion i (i.e., K+ and Arg−), hg is the gas specific parameter, and Cion is the concentration of ion i in the solution. In this work, Cion = CArgK. The gas side mass transfer coefficient and liquid side physical mass transfer coefficient were estimated by empirical correlations reported in our previous work.27 2.2. Reaction Mechanisms in Aqueous AAS Solution. Zwitterion mechanism (ZM) was initially proposed by Caplow34 and is generally accepted to model the CO2 absorption using the aqueous amino acid salt solutions.5−25 CO2 reacts with the AAS (ArgK, in the present case) via the formation of a zwitterion as an intermediate (see eq 8) and this is followed by deprotonation by base species:

(2)

kovDCO2

DCO2 ⎛ H C ⎞ D ⎜⎜1 + AAS × CO2 AAS ⎟⎟ DAAS ⎝ DCO2 νAASPCO2 ⎠

1 kg

where DCO2,H2O and ηH2O are the diffusion coefficient and viscosity of CO2 in water, respectively, which can be obtained from open literature.7,31 The solubility of CO2 in aqueous ArgK solutions can be estimated by Schumpe model using ion and gas specific parameters:29,32,33

where

kL0



2

kg

where NCO2 is the molar flow of CO2 entering the liquid solution. PCO2,b is the CO2 partial pressure in the gas bulk phase. PCO2,b will change in the circulus as the CO2 is absorbed, therefore it can be estimated and expressed by the log mean pressure (Pln,CO2) obtained from the top PCO2,b and the bottom PCO2,b in the wetted-wall column.7 P*CO2 is the equilibrium partial pressure and can be neglected (i.e., PCO2,b ≫ P*CO2) by using CO2-free ArgK solutions for each run of experiments and the CO2 loading at the outlet of WWC for each experiment is low. k0L is the physical mass transfer coefficient in the liquid. kg is the gas-side mass transfer coefficient. HCO2 is the Henry’s constant of CO2 in solution. E is the enhancement factor, which represents the ratio between the rate of chemical absorption and that of physical absorption. When the absorption is in the fast pseudo-first-order reaction regime if eq 2 is fulfilled, the E equals Hatta number (Ha).9,17,25

Ha =

NCO2

0.6 DCO2ηHisK = DCO2,H2OηH0.6O

(1)

3 < Ha = E < 11.0; (c) pH < 8.5 with acid adjustment) and CO2loaded KOH solution (d), in D2O.

Table 2. Values of Apparent Reaction Rate Constant, kapp, for the Absorption of CO2 into ArgK Solutions kapp (s−1) CArgK (M)

298 K

0.26

2156 ± 13

0.50

7002 ± 557 13182 ± 1856 20478 ± 1638 28750 ± 2300

0.75 1.05 1.23

303 K

313 K

323 K

333 K

2769 ± 131 8818 ± 693 15187 ± 1490 25060 ± 1253 35236 ± 1762

3917 ± 106 11318 ± 1020 19395 ± 2753 32784 ± 1639 42927 ± 2146

5199 ± 395 14438 ± 1940 26818 ± 4002 40441 ± 2025 53962 ± 2698

7142 ± 463 19364 ± 1863 33255 ± 3604 51770 ± 2588 58939 ± 2947

ArgK and CO2 if the formation of zwitterion is ratedetermining step for zwitterion mechanism. So, the overall reaction order is two. However, the overall reaction order will be in the range between 2 and 3 when the zwitterion deprotonation is the rate-determining step. In this case, the partial reaction order for ArgK species is in between 1 and 2. According to termolecular mechanism, it will be a first-order reaction with respect to ArgK when H2O is dominant over other basic species (i.e., OH− and ArgK). When the contributions of H2O and ArgK to the reaction in eqs 12 and 13 are comparable, a partial reaction order between 1 and 2 is expected for ArgK. Partial reaction orders presented in this work suggest that the zwitterion deprotonation steps (eq 9) are not faster than zwitterion formation and most likely are going

Figure 6. CO2 normalized flux in aqueous ArgK solutions as a function of temperature.

to be rate-controlling. Therefore, the contribution of basic species in ArgK solutions to zwitterion deprotonation are taken into consideration. The zwitterion transition region model (eq 11) is suitable for interpreting the experimental data of this work. The contributions of H2O and ArgK are also considered in termolecular mechanism model (eq 13). The individual reaction rate constants in above-mentioned models were derived from the experimental data. As observed in Figure 8, the kapp shows a strong temperature and concentration dependence. The kapp results were non6592

DOI: 10.1021/acs.energyfuels.6b01092 Energy Fuels 2016, 30, 6585−6596

Article

Energy & Fuels

The fitting ZM and TM model predictions are shown in dash lines along with the experimental data in Figures 8 and 9,

Figure 7. Overall kinetic constants for aqueous ArgK solutions compared with other absorbents at 298 K. Figure 9. Apparent reaction rate constants as a function of ArgK concentrations and at different temperatures: experimental data and lines from TM model.

respectively. Both models match well with the experimental data. The average absolute deviation (AAD) between model predictions and experimental results was found to be 3.6% for ZM model and within 7.5% for TM model, respectively. It was noted that the experimental data at low ArgK concentrations (e.g., for 0.26 M ArgK) were slightly lower than the predictions from the TM model. It should be pointed out that these models are not taking the ion strength effect into consideration. If ArgK is used as a promoter in highly concentrated salt solutions (e.g., potassium carbonate), the nonidealities of the solution should not be neglected. The calculated rate constants in this work and other AAS reported in the literature are with the same order of magnitude.25,27 It was noted that water contributes significantly to the carbamate and bicarbonate formation. Take TM model as example, the partial reaction order with respect to ArgK can be derived by estimating the individual contribution of ArgK, kTArgKCArgK, to kapp from eq 13. The individual contribution of ArgK accounted for about 26% for 0.50 M and 61% for 1.05 M at 313 K. Increasing the ArgK concentration, the contribution of ArgK to the formation of carbamate will be more. Its average contribution (49%) for all runs indicates that the partial order should be about 1.49 with respect to ArgK. This result is in line with the value (i.e., 1.42) in eq 19. Therefore, water is still acting as a dominant base and its contribution to the overall reaction rate is significant, especially for the absorbent solutions at low concentrations of ArgK. For the zwitterion mechanism, if k−1/∑ kZM B CB≪ 1, the second-order rate constants (k2) can be obtained from a second-order kinetics model. k2 and the acid dissociation constants (pKa) of amines can be correlated by the reported Brønsted relationships.8,47 It was found that the correlations proposed by Versteeg et al.47 and Penny and Ritter8 give prediction values with large variation. A comparison of k2* for amino acid salts and amines at 298 K can be found in our previous work.27 k2* is defined as the apparent second order rate constant (kov/CAAS, m3 mol−1 s−1), which is obtained from the scattered kinetic data at about 1.0 mol L−1 AAS reported in the literature (see SI Table S1). Brønsted relationships at 298 and 313 K were correlated, respectively, and shown in Figure

Figure 8. Apparent reaction rate constants as a function of ArgK concentrations and at different temperatures: experimental data and model lines from ZM model.

linearly regressed to obtain the individual rate constants expressed as the Arrhenius equations by minimizing the sum of the relative residues. The individual rate constants are given as eqs 20−24. Zwitterion mechanism (ZM): ⎛ −731 ⎞ ⎟ k 2ZM = 6.476 × 105exp⎜ ⎝ T ⎠ ZM kArgK

k −1

k HZM 2O k −1

(20)

⎛ −3083 ⎞ ⎟ = 1.772 × 104exp⎜ ⎝ T ⎠

(21)

⎛ −6340 ⎞ ⎟ = 7.998 × 105exp⎜ ⎝ T ⎠

(22)

Termolecular mechanism (TM): ⎛ −1407 ⎞ T ⎟ kArgK = 1.62 × 106exp⎜ ⎝ T ⎠

(23)

⎛ −3637 ⎞ ⎟ k HT2O = 2.75 × 107exp⎜ ⎝ T ⎠

(24) 6593

DOI: 10.1021/acs.energyfuels.6b01092 Energy Fuels 2016, 30, 6585−6596

Article

Energy & Fuels

spectra. The partial reaction order was found to be an average of 1.42 with respect to ArgK. Water and ArgK are acting as dominant bases to form carbamate and should be considered in both ZM and TM models. Correlations of the individual rate constants from two mechanisms can predict well the CO2 absorption rate in the ArgK solutions. Moreover, a rough Brønsted relationship for amino acid salts was proposed based on the reported rate constants.

10. It should be pointed out that the pKa values of amine groups in Figure 10b are also obtained at 298 K. 3793 ln k 2* = 1.05pK a + 5.26 − T

(25)



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.energyfuels.6b01092. The empirical correlations for physicochemical properties and physical mass transfer coefficients in ArgK solutions, comparison of apparent second-order rate constants for amino acid salts and amines, the reported Brønsted relationships (DOC)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]; Tel.: +86 311 88632183; Fax: +86 311 88632183. Funding

The authors acknowledge National Natural Science Foundation of China (No.21206029), Hebei Provincial Natural Science Foundation for Distinguished Young Scholars of China (B2015208067), and Hebei Provincial Science and Technology Research Project of College and University (QN2015070) for financial support. Notes

The authors declare no competing financial interest.

■ Figure 10. Brønsted plots for amino acid salts and MEA at 298 K (a) and 313 K (b). k2* = kov/CAAS.

It can be seen that the data points of k2* are generally in line with the pKa of amine groups. However, the calculated values from the correlations eq 25 underestimate for ArgK and PZ, and overestimate for AlaK. The high basicity of side chain (pKa 12.10) in ArgK molecules might improve the formation of carbamate and bicarbonate (eqs 9 and 12). The fact that PZ, ArgK, and LysK have faster rate constants than expected from the correlations may be attributed to their molecules containing basic side chain or cyclic diamine structure. More work on the kinetic rate constants under the same conditions is required to obtain a rigorous correlation for AAS in the future.

5. CONCLUSION The reaction kinetics between ArgK and CO2 was investigated in a wetted wall column at concentrations of 0.26−1.23 M and temperatures of 298−333 K. ArgK shows fast kinetics toward CO2, but has limited solubility in water. ArgK carbamate and bicarbonate are the main products of CO2 absorption into the deprotonated ArgK solutions identified by the 13C NMR 6594

NOMENCLATURE AAS=amino acid salt B=base species such as H2O, OH−, and AAS C=concentration, M or kmol m−3 D=diffusion coefficient, m2 s−1 E=enhancement factor, dimensionless E∞=infinite enhancement factor in instantaneous reaction regime, dimensionless hion, hg=ion and gas specific constants in the Shumpe equation, m3 kmol−1 hArg−=ion-specific parameter of Arg− H=Henry coefficient, kPa m3 kmol−1 Ha=Hatta number, dimensionless HCO2=Henry coefficient of CO2 in the ArgK solution, kPa m3 kmol−1 HCO2, H2O=Henry coefficient of CO2 in the water, kPa m3 kmol−1 k−1=reverse reaction rate constant, L mol−1 s−1 k2ZM=second order rate constant, L mol−1 s−1 kapp=apparent reaction rate constant, s−1 kArgKT=termolecular reaction rate constants contributed by ArgK, L2 mol−2 s−1 kArgKZM=zwitterion deprotonation rate constant by ArgK, L mol−1 s−1 kBZM=zwitterion deprotonation rate constant by bases, L mol−1 s−1 kH2OT=termolecular reaction rate constants contributed by water, L2 mol−2 s−1 DOI: 10.1021/acs.energyfuels.6b01092 Energy Fuels 2016, 30, 6585−6596

Article

Energy & Fuels kH2OZM=zwitterion deprotonation rate constant by H2O, L mol−1 s−1 kg=mass transfer coefficient in gas phase, mol m−2 s−1 Pa−1 kg’=normalized flux (a partial pressure driving force across the liquid film), mol m−2 s−1 Pa−1 KG=overall gas phase mass transfer coefficient, mol m−2 s−1 Pa−1 kL0=liquid phase physical mass transfer coefficient, m s−1 kOH−=reaction rate constant with hydroxide ion, L mol−1 s−1 kov=overall kinetic constant, s−1 NCO2=CO2 absorption flux, mol m−2 s−1 P1, P2=gas pressure, kPa PCO2=CO2 partial pressure, kPa pCO2,t=partial pressure of CO2 at time t in the reactor, kPa P*CO2=equilibrium partial pressure Pln=log mean pressure (Pln) at the top and bottom of the wetted-wall column, kPa Ps=vapor pressure of solvent, kPa −rCO2=rate of reaction, mol s−1 R=universal gas constant, J mol−1 K−1 Re=Reynolds number Sc=Schmidt number Sh=Sherwood number T=temperature, K

(10) Thee, H.; Nicholas, N. J.; Smith, K. H.; da Silva, G.; Kentish, S. E.; Stevens, G. W. A kinetic study of CO2 capture with potassium carbonate solutions promoted with various amino acids: glycine, sarcosine and proline. Int. J. Greenhouse Gas Control 2014, 20, 212− 222. (11) Guo, D.; Thee, H.; Tan, C. Y.; Chen, J.; Fei, W.; Kentish, S. E.; Stevens, G. W.; da Silva, G. Amino acids as carbon capture solvents: chemical kinetics and mechanism of the glycine + CO2 reaction. Energy Fuels 2013, 27, 3898−3904. (12) Lee, S.; Song, H. J.; Maken, S.; Park, J. W. Kinetics of CO2 absorption in aqueous sodium glycinate solutions. Ind. Eng. Chem. Res. 2007, 46, 1578−1583. (13) Park, S. -W.; Son, Y. -S.; Park, D. -W.; Oh, K. -J. Absorption of carbon dioxide into aqueous solution of sodium glycinate. Sep. Sci. Technol. 2008, 43, 3003−3019. (14) Sodiq, A.; Rayer, A. V.; Olanrewaju, A. A.; Abu Zahra, M. R. M. Reaction kinetics of carbon dioxide (CO2) absorption in sodium salts of taurine and proline using a stopped-flow technique. Int. J. Chem. Kinet. 2014, 46, 730−745. (15) Wei, C.-C.; Puxty, G.; Feron, P. Amino acid salts for CO2 capture at flue gas temperatures. Chem. Eng. Sci. 2014, 107, 218−226. (16) Majchrowicz, M. E.; Kersten, S.; Brilman, W. Reactive absorption of carbon dioxide in L-prolinate salt solutions. Ind. Eng. Chem. Res. 2014, 53, 11460−11467. (17) Paul, S.; Thomsen, K. Kinetics of absorption of carbon dioxide into aqueous potassium salt of proline. Int. J. Greenhouse Gas Control 2012, 8, 169−179. (18) Hwang, K.-S.; Park, D.-W.; Oh, K.-J.; Kim, S.-S.; Park, S.-W. Chemical absorption of carbon dioxide into aqueous solution of potassium threonate. Sep. Sci. Technol. 2010, 45, 497−507. (19) Portugal, A. F.; Magalhães, F. D.; Mendes, A. Carbon dioxide absorption kinetics in potassium threonate. Chem. Eng. Sci. 2008, 63, 3493−3503. (20) Simons, K.; Brilman, W.; Mengers, H.; Nijmeijer, K.; Wessling, M. Kinetics of CO2 absorption in aqueous sarcosine salt solutions: influence of concentration, temperature, and CO2 loading. Ind. Eng. Chem. Res. 2010, 49, 9693−9702. (21) Aronu, U. E.; Hartono, A.; Hoff, K. A.; Svendsen, H. F. Kinetics of carbon dioxide absorption into aqueous amino acid salt: potassium salt of sarcosine solution. Ind. Eng. Chem. Res. 2011, 50, 10465−10475. (22) Kim, M.; Song, H.; Lee, M.; Jo, H.; Park, J. Kinetics and steric hindrance effects of carbon dioxide absorption into aqueous potassium alaninate solutions. Ind. Eng. Chem. Res. 2012, 51, 2570−2577. (23) Lim, J.; Kim, D. H.; Yoon, Y.; Jeong, S. K.; Park, K. T.; Nam, S. C. Absorption of CO2 into aqueous potassium salt solutions of Lalanine and L-proline. Energy Fuels 2012, 26, 3910−3918. (24) Shen, S.; Yang, Y.; Wang, Y.; Ren, S.; Han, J.; Chen, A. CO2 absorption into aqueous potassium salts of lysine and proline: density, viscosity and solubility of CO2. Fluid Phase Equilib. 2015, 399, 40−49. (25) Shen, S.; Yang, Y.; Bian, Y.; Zhao, Y. Kinetics of CO2 absorption into aqueous basic amino acid salt: potassium salt of lysine solution. Environ. Sci. Technol. 2016, 50, 2054−2063. (26) Muñoz, D. M.; Portugal, A. F.; Lozano, A. E.; de la Campa, J. G.; de Abajo, J. New liquid absorbents for the removal of CO2 from gas mixtures. Energy Environ. Sci. 2009, 2, 883−891. (27) Shen, S.; Yang, Y.; Zhao, Y.; Bian, Y. Reaction kinetics of carbon dioxide absorption into aqueous potassium salt of histidine. Chem. Eng. Sci. 2016, 146, 76−87. (28) CRC Handbook of Chemistry and Physics, 90th Edition (CDROM Version 2010); Lide, D. R. Ed.; CRC Press: Taylor and Francis: Boca Raton, FL, 2010. (29) Bian, Y.; Shen, S.; Zhao, Y.; Yang, Y. Physicochemical properties of aqueous potassium salts of basic amino acids as absorbents for CO2 capture. J. Chem. Eng. Data 2016, 61, 2391−2398. (30) Shen, S.; Feng, X.; Ren, S. Effect of arginine on carbon dioxide capture by potassium carbonate solution. Energy Fuels 2013, 27, 6010− 6016. (31) Laliberté, M. Model for calculating the viscosity of aqueous solutions. J. Chem. Eng. Data 2007, 52, 321−335.

Greek letters

α=CO2 loading, mol CO2 per mol of amino acid salt, dimensionless η=solution viscosity, mPa s ρ=solution density, g cm−3 ν=the stoichiometric coefficient for ArgK (ν = 2)

Subscripts

AAS=amino acid salt ArgK=potassium salt of arginine CO2=Carbon dioxide H2O=water



REFERENCES

(1) MacDowell, N.; Florin, N.; Buchard, A.; Hallett, J.; Galindo, A.; Jackson, G.; Adjiman, C. S.; Williams, C. K.; Shah, N.; Fennell, P. An overview of CO2 capture technologies. Energy Environ. Sci. 2010, 3, 1645−1669. (2) Yildirim, Ö .; Kiss, A. A.; Hüser, N.; Leßmann, K.; Kenig, E. Y. Reactive absorption in chemical process industry: a review on current activities. Chem. Eng. J. 2012, 213, 371−391. (3) Mumford, K. A.; Wu, Y.; Smith, K. H.; Stevens, G. W. Review of solvent based carbon-dioxide capture technologies. Front. Chem. Sci. Eng. 2015, 9, 125−141. (4) van Holst, J.; Versteeg, G. F.; Brilman, D. W. F.; Hogendoorn, J. A. Kinetic study of CO2 with various amino acid salts in aqueous solution. Chem. Eng. Sci. 2009, 64, 59−68. (5) Vaidya, P. D.; Konduru, P.; Vaidyanathan, M.; Kenig, E. Y. Kinetics of carbon dioxide removal by aqueous alkaline amino acid salts. Ind. Eng. Chem. Res. 2010, 49, 11067−11072. (6) Kumar, P. S.; Hogendoorn, J. A.; Versteeg, G. F.; Feron, P. H. M. Kinetics of the reaction of CO2 with aqueous potassium salt of taurine and glycine. AIChE J. 2003, 49, 203−213. (7) Shen, S.; Feng, X.; Zhao, R.; Ghosh, U. K.; Chen, A. Kinetic study of carbon dioxide absorption with aqueous potassium carbonate promoted by arginine. Chem. Eng. J. 2013, 222, 478−487. (8) Penny, D. E.; Ritter, T. J. Kinetic study of the reaction between carbon dioxide and primary amines. J. Chem. Soc., Faraday Trans. 1 1983, 79, 2103−2109. (9) Portugal, A. F.; Derks, P. W. J.; Versteeg, G. F.; Magalhães, F. D.; Mendes, A. Characterization of potassium glycinate for carbon dioxide absorption purposes. Chem. Eng. Sci. 2007, 62, 6534−6547. 6595

DOI: 10.1021/acs.energyfuels.6b01092 Energy Fuels 2016, 30, 6585−6596

Article

Energy & Fuels (32) van Holst, J.; Kersten, S. R. A.; Hogendoorn, J. A. Physiochemical properties of several aqueous potassium amino acid salts. J. Chem. Eng. Data 2008, 53, 1286−1291. (33) Weisenberger, S.; Schumpe, A. Estimation of gas solubilities in salt solutions at temperatures from 273 to 363 K. AIChE J. 1996, 42, 298−300. (34) Caplow, M. Kinetics of carbamate formation and breakdown. J. Am. Chem. Soc. 1968, 90, 6795−6803. (35) Crooks, J. E.; Donnellan, J. P. Kinetics and mechanism of the reaction between carbon dioxide and amines in aqueous solution. J. Chem. Soc., Perkin Trans. 2 1989, 2, 331−333. (36) da Silva, E. F.; Svendsen, H. F. Ab Initio study of the reaction of carbamate formation from CO2 and alkanolamines. Ind. Eng. Chem. Res. 2004, 43, 3413−3418. (37) Astarita, G.; Savage, D. W.; Bisio, A. Gas treating with chemical solvents; John Wiley & Sons: New York, 1983. (38) Olofsson, G.; Hepler, L. G. Thermodynamics of ionization of water over wide ranges of temperature and pressure. J. Solution Chem. 1975, 4, 127−143. (39) Nagai, H.; Kuwabara, K.; Carta, G. Temperature dependence of the dissociation constants of several amino acids. J. Chem. Eng. Data 2008, 53, 619−627. (40) Kierzkowska-Pawlak, H.; Chacuk, A.; Siemieniec, M. Reaction kinetics of CO2 in aqueous 2-(2-aminoethylamino)ethanol solutions using a stirred cell reactor. Int. J. Greenhouse Gas Control 2014, 24, 106−114. (41) Jiru, Y.; Eimer, D. A. A study of mass transfer kinetics of carbon dioxide in (monoethanolamine + water) by stirred cell. Energy Procedia 2013, 37, 2180−2187. (42) Rochelle, G. T.; Chen, E.; Freeman, S. A.; Van Wagener, D. H.; Xu, Q.; Voice, A. K. Aqueous piperazine as the new standard for CO2 capture technology. Chem. Eng. J. 2011, 171, 725−733. (43) Ciftja, A. F.; Hartono, A.; Svendsen, H. F. 13C NMR as a method species determination in CO2 absorbent systems. Int. J. Greenhouse Gas Control 2013, 16, 224−232. (44) Yan, S.; He, Q.; Zhao, S.; Zhai, H.; Cao, M.; Ai, P. CO2 removal from biogas by using green amino acid salts: performance evaluation. Fuel Process. Technol. 2015, 129, 203−212. (45) Laddha, S. S.; Danckwerts, P. V. Reaction of CO2 with ethanolamines: Kinetics from gas absorption. Chem. Eng. Sci. 1981, 36, 479−482. (46) Derks, P. W. J.; Kleingeld, T.; van Aken, C.; Hogendoorn, J. A.; Versteeg, G. F. Kinetics of absorption of carbon dioxide in aqueous piperazine solutions. Chem. Eng. Sci. 2006, 61, 6837−6854. (47) Versteeg, G. F.; van Dijck, L. A. J.; Van Swaaij, W. P. M. On the kinetics between CO2 and alkanolamines both in aqueous and nonaqueous solutions. An overview. Chem. Eng. Commun. 1996, 144, 113−158.

6596

DOI: 10.1021/acs.energyfuels.6b01092 Energy Fuels 2016, 30, 6585−6596