Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Article

pubs.acs.org/IECR

Facile Method for Determination of Amine Speciation in CO2 Capture


Solutions
Naser S. Matin, Joseph E. Remias, James K. Neathery, and Kunlei Liu*
Center for Applied Energy Research, University of Kentucky, Lexington, Kentucky 40511, United States
*
S Supporting Information

ABSTRACT: A simple and quantitatively reliable method for determination of amine speciation is introduced. The method
employs three experimental methods that should be readily accessible. The results for CO2 loaded aqueous solutions of 30 wt %
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

monoethanolamine (MEA) are used for demonstration of the method and show promising agreement with the more
Downloaded via KTH ROYAL INST OF TECHNOLOGY on June 19, 2024 at 14:00:03 (UTC).

complicated spectroscopic methods reported in the literature. The measurements were done at ambient temperature and
atmospheric pressure, since theoretical calculations and experimental data from the open literature revealed no considerable
difference in speciation at different temperatures. The procedure is based on acid and base titration of the CO2 loaded amine
solution along with the determination of total CO2 loading. The quantitative results for the different species concentration in the
example MEA solution is in agreement with other available spectroscopic methods, mainly NMR, particularly for the free amine,
carbamate, and protonated amine concentrations. Aspen Plus was also used for further assessment of the experimental data.

■ INTRODUCTION
Removal of acidic gases, particularly CO2, is a very important
necessary.9 They used the combination of proton and carbon
NMR to study speciation in monoethanolamine (MEA)−
operation from an industrial and environmental point of view. H2O−CO2 and diethanolamine (DEA)−H2O−CO2 solutions
Primary, secondary, and tertiary amines and their blends have at a wide CO2 loading range and temperatures between 20 and
found widespread application in the absorption and removal of 80 °C. Of course, as noted, due to the fast proton transfer
carbon dioxide from process gases.1,2 In the designing of between molecular and protonated amine, only the sum of their
absorption or desorption columns for a CO2 capture process by concentrations can be observed. Furthermore, the NMR
amine or other reactive solvents, the precise knowledge of the method is time-consuming and requires expensive equipment.
reaction chemistry is crucial. In this regard, having an accurate Jakobsen et al. applied the same technique in the experimental
determination of liquid phase species concentration is critical in part of their work to study species distribution in 2-[(2-
CO2 absorption processes modeling vapor−liquid equilibria aminoethyl)amino]-ethanol (AEEA)−H2O−CO2 system.10
data. The contribution of the protonated amine, carbamate, and Fourier transform infrared spectroscopy (FTIR) was also
bicarbonate formed in amine solvents to, for example, the CO2 used for the speciation study of CO2 absorbed into aqueous
heat of absorption makes their quantities critical for alkanolamines by Jackson et al.11 Their work was not
thermodynamic understanding.3 Furthermore, providing a quantitative, but they assigned the major peaks corresponding
good estimation of the free amine concentration is important to a few important species in the liquid phase for MEA−CO2
for mass transfer study of the CO2 absorption process, where and 2-amino-2-methyl-1-propanol (AMP)−CO2 system. On
the free amine concentration explicitly appears in the liquid film the other hand Derks et al. recently applied the FTIR technique
mass transfer coefficient.4 Finally, with good understanding of to quantitative measurements of speciation in CO2 loaded
the above concepts it is possible to better design new solvents aqueous solutions of N-methyldiethanolamine (MDEA).12 For
and blends. performing the analysis, they initially prepared calibration
Various thermodynamic models, mainly based on activity curves for the species under study. Souchon et al. employed in
coefficients, have been proposed for the CO2−H2O−alkanol- situ Raman spectroscopy for the quantitative determination of
amine systems.5−7 However, many of the models employ the species distribution in alkanolamines (MEA, DEA, and
computational models, which are not readily accessible or MDEA)−H2O−CO2 systems at equilibrium at 40 °C.13 A gas
require parameters not necessarily available for all solvents, chromatography technique also has been proposed.14
making experimental methods of merit.5 Furthermore, reliable A titration method for the concentration measurement of
experimental values for species concentration in the liquid different species in AMP−H2O−CO2 and DEA−H2O−CO2
phase can improve the thermodynamic models. The activity was used by Haji-Sulaiaman et al.15 They employed a strong
coefficient models and their parameters can be better evaluated base (i.e., NaOH) to titrate the solution, followed by solving
by the available amine speciation data.3 the equilibrium equations simultaneously to determine solution
The NMR approach for speciation study of alkanolamine
solutions containing dissolved CO2 is one of the basic methods Received: February 3, 2012
for this purpose.8−10 As it was discussed by Böttinger et al., the Revised: April 20, 2012
quantification study in NMR spectroscopy is easier than optical Accepted: April 27, 2012
spectroscopy or gas chromatography because no calibration is Published: April 27, 2012

© 2012 American Chemical Society 6613 dx.doi.org/10.1021/ie300230k | Ind. Eng. Chem. Res. 2012, 51, 6613−6618
Industrial & Engineering Chemistry Research Article

speciation. Their approach is simple and reliable but still 0.1 M sulfuric acid solution was utilized here with sample size
requires solving the simultaneous system of equations and selected to require a minimum of 10 mL of titrant to minimize
requires the necessary thermodynamic parameters, which may volumetric delivery error.
not be available for all amines. For the base titrations, the dynamic equivalence titration
In this work, a new method based on the total alkalinity (DET) method incorporated in the Metrohm Titrando 836
measurements (acid titration), strong base titration, and CO2 software was used as previously described.15 The method was
loading of the solution was developed to calculate the different validated by adding known amounts of potassium bicarbonate
species concentrations in the MEA (monoethanolamine)− to 30 wt % MEA solutions and titrating using the method. A
H2O−CO2 system. The method has several advantages in that nominally 0.1 M sodium hydroxide solution was used for the
it is fast, repeatable, and employs methods that are typically base titrations with sample size selected to require a minimum
available in any laboratory studying CO 2 absorption. of 10 mL of titrant. The base titrations were performed at
Furthermore, the approach does not require the knowledge ambient temperature. For the base titration, eq 1 can be used
of any thermodynamic parameters to determine speciation. The with M, n, v as the base molarity, number of hydroxide ions per
developed system is compared to experimentally determined base molecule, and volume of base (in mL) used for titration,
values obtained using NMR for the same solution and to an respectively. (see Figure S.2, Supporting Information for typical
Aspen Plus equilibrium model for MEA. Good agreement was DET titration curve)
found between the three methods.

■ EXPERIMENTAL SECTION
■ RESULTS AND DISCUSSION
Background. Liquid bulk concentrations of chemical
The aqueous solutions of 30 wt % MEA with different CO2 species and the partial pressure of solutes in the gas phase
loadings were prepared and used for all experiments by are required for all calculations including kinetic analysis and
contacting the aqueous MEA solution with a CO2 containing system simulation. Therefore, the independent chemical
gas stream for different time periods. From each sample, five reactions, which produce or consume stable species in the
pieces of data were collected: pH, density, CO2 loading, total bulk phase, and phase equilibrium equations have to be known
alkalinity, and base titration. The pH was measured at ambient for the system under study. Having such information makes it
temperature using a temperature corrected probe and 2 point possible to calculate the desired quantities mentioned above.
calibration. The density was measured by dispensing 1 mL of The following reactions may occur when CO2 absorbs into
solution (calibrated pipet to water) in quadruplicate and and reacts with aqueous primary and secondary amines.17
determining the mass. Ionization of water:
The CO2 solution loading was measured by an adaptation of
2H 2O ↔ OH− + H3O+ (2)
a total inorganic carbon method previously described.16 In this
method, phosphoric acid liberates the dissolved CO2 in amine Dissociation of dissolved CO2 through carbonic acid:
solutions. The CO2 gas is stripped out with a nitrogen carrier
gas and is directed to a HORIBA CO2 analyzer (VIA-510). The CO2 + 2H 2O ↔ HCO3− + H3O+ (3)
area under the curve is integrated, and CO2 concentration is
Dissociation of bicarbonate:
determined using a calibration curve from a known analytical
standard of potassium carbonate. Approximately 1 mL of HCO3− + H 2O ↔ CO32 − + H3O+ (4)
sample (mass determined) was injected into the phosphoric
acid reactor yielding CO2 concentration in mol CO2/kg sln. Carbamate reversion to bicarbonate (hydrolysis reaction):
The HORIBA CO2 analyzer was calibrated with a certified RNHCOO− + H 2O ↔ RNH 2 + HCO−3 (5)
CO2/N2 gas mixture (PurityPlus, 14.00% CO2) each day. The
uncertainty was checked with a known analytical standard prior Dissociation of protonated amine:
and after each set of unknowns with the allowable discrepancy
in CO2 measurement (|(expected − measured)/expected| × RNH3+ + H 2O ↔ RNH 2 + H3O+ (6)
100) set at less than ±2% absolute. The standard deviation for Considering the above reactions, the potential species
repeated measurements was ±2.7%. concentrations in question are OH−, H3O+, HCO3−, CO2,
Total alkalinity measurements were conducted using a CO 3 − , RNH 3 + , RNH 2 (unreacted free amine), and
Metrohm Titrando 836 and a standard equivalence point RNHCOO−. One approach to finding the concentrations of
determination method, (with accuracy of ±0.003 in pH, and the above-mentioned species is solving eight of the equations
standard deviation at ±3.6% in end point measurements). below simultaneously. The chemical reaction equilibriums, the
Having the consumed acid volume at the final equivalence mass and charge balances in the aqueous phase, and the phase
point (typically two were observed, see Figure S.1, Supporting equilibrium between CO2 in the gas and aqueous phase give
Information) and its molarity, the following equation is used for enough (eqs 7−15) to perform the calculations.17
alkalinity determination (and also base titration) of sample Amine balance:
solution. A typical total alkalinity titration curve is shown in
Figure S.1, Supporting Information. [RNH 2]f + [RNH3+] + [RNHCOO−] = [RNH 2]t (7)
nMv Carbon balance:
Total alkalinity (moleCO2 /kg sln) =
wsample (1)
[CO2 ] + [HCO3−] + [CO32 −] + [RNHCOO−]
Where M, n, v, and wsample are molarity of acid, number of = α[RNH 2]t (8)
proton per acid molecule, volume of acid was used during
titration (mL), and sample mass (g), respectively. A nominal Charge balance:
6614 dx.doi.org/10.1021/ie300230k | Ind. Eng. Chem. Res. 2012, 51, 6613−6618
Industrial & Engineering Chemistry Research Article

Table 1. Equilibrium Constants18,19,a


A B C D
K1, (mol/dm3)2 132.89888 −13445.9 −22.4773 0
K2, (mol/dm3) 231.465439 −12092.1 −36.7816 0
K3, (mol/dm3) 216.050446 −12431.7 −35.4819 0
K4, (mol/dm3) −0.52135 −2545.53 0 0
K5, (mol/dm3) −3.038325 −7008.357 0 −0.0031348
a
ln (K) = A + B/T + C ln(T) + DT, where T is in Kelvin.

[RNH3+] + [H3O+] CO2 + RNH 2 + H 2O ↔ H3O+ + RNHCOO− (17)

= [HCO3−] + [OH−] + 2[CO32 −] + [RNHCOO−] Therefore, it can be assumed that, during titration for total
(9) alkalinity, the free amine also includes all of the carbamate in
the solution. Considering the pH for carbonated alkanolamine
Independent equilibrium constants: solutions, which are generally between 8 to 10, and the
K1 = [OH−][H3O+] dissociation constant (pK2) of dissolved CO2 through carbonic
(10)
acid at ambient temperature is ∼6.0 (see Table 1); it is a
plausible approximation that the bicarbonate−carbon dioxide
K 2 = [HCO3−][H3O+]/[CO2 ] (11)
equilibrium, eq 3 is shifted toward the formation of bicarbonate.
Thus, the concentration of free CO2 is generally very low.15
K3 = [CO32 −][H3O + ]/[HCO3−] (12) The alternative form of eq 8, considering carbon balance, gives
K4 = [RNH 2][HCO3−]/[RNHCOO−] (13) [CO2 ]t = [CO2 ]f + [HCO3−] + [CO32 −]

K 5 = [RNH 2][H3O + ]/[RNH3+] + [RNHCOO−] (8-a)


(14)
Therefore, from eq 16, 8-a, and discussion followed by eq 17,
PCO2 = He × [CO2 ] (15) the following equation can be reached,
Where [RNH2]f, [RNH2]t, α, and He are unreacted free amine Alka t = {[RNH 2]f + [RNHCOO−]} + 2[CO2 ]t
and total amine concentrations, carbon loading, and Henry’s
constant. The equilibrium constants are given in Table 1. − [HCO3−] − [RNHCOO−] (16-a)
Using eqs 2−15 is the approach applied to thermodynamic which is simplified as follows,
modeling of CO2 loaded solutions (alkanolamines) to calculate
each species concentration in the liquid and gas phase, and it is Alka t = [RNH 2]f + 2[CO2 ]t − [HCO3−] (16-b)
necessary to solve and optimize the system of nonlinear Considering eq 4, at basic conditions, the following equilibrium
equations over the experimental solubility data. In the next reaction can be considered between bicarbonate and carbonate
section, a simple and reliable titration method is introduced to ions
obtain the speciation data in alkanolamine solutions based on
the mass and charge balance without the need for equilibrium CO32 − + H 2O ↔ HCO3− + OH− (4-a)
constants. Using the dissociation constant for bicarbonate to carbonate
Method Description. For total alkalinity measurement, the ion and proton at room temperature (pK3 ≅ 10.3), and the
reactions 3, 4, and 6 can take place with approximate pK values water dissociation constant (pKw ≅14), the equilibrium
of 7, 11, and 10, respectively (see Table 1). Therefore, constant for eq 4-a can be expressed as
considering the pK of the mentioned reactions, with addition of
proton (acid) the bicarbonate ion and protonated amine K3‐a = K w /K3 = [HCO3−][OH−]/[CO32 −] ≅ 10−3
formation take place at approximately pH ≅ 6 (pH of 7.3 in (18)
experiments), that is, the first equivalence point. Further
Considering eq 18, the basicity determines the bicarbonate/
addition of acid results in the bicarbonate ion converting to
carbonate ratio in the solution. Therefore, it can be inferred at
CO2 through eq 3 at pH ≅ 4 (pH of 4.4 in experiments), that
the mentioned solution’s pH (8 ≤ pH ≤ 10) that
is, the second equivalence point. Depending on the amine
protonation equilibrium constant of various alkanolamines, the 10 ≤ [HCO3−]/[CO32 −] ≤ 1000
pH versus acid volume graph might display three equivalence
points, that is, a large difference in equilibrium constants of eqs So, for most CO2 loading ranges considered and particularly at
4 and 6, but the pattern and interpretation remain the same. the loadings above 0.25 molCO2/molamine, the concentration of
The total alkalinity of the solution can be calculated using an CO32− compared to HCO3− is very low, and it can be discarded
acid titration, as follows: from the eq 8 or 8-a. Therefore, the simplified form of the eq 8-
a is
Alka t = [RNH 2]f + [HCO3−] + 2[CO32 −]
[CO2 ]t = [HCO3−] + [RNHCOO−] (8-b)

+ [RNHCOO ] (16)
where [CO2]t was determined experimentally using the above-
The addition of the acid (proton) in alkalinity measurement described method.
causes protonation of the carbamate to produce free amine and Moving from the alkalinity measurements, as the acid
CO2 through the reverse reaction titration gives the free amine concentration, the dynamic
6615 dx.doi.org/10.1021/ie300230k | Ind. Eng. Chem. Res. 2012, 51, 6613−6618
Industrial & Engineering Chemistry Research Article

titration of the amine solution with a strong base yields the agreement with the Aspen Plus predicted values. The pH is
concentration of bicarbonate, protonated amine and free defined as the activity of the proton in the solution. On the
CO2,15 so that the following balanced equation can be written other hand, the single ion activity coefficients are related to the
activity coefficient of the neutral electrolyte (i.e., γ+γ− = γN2).21
[Base] = [RNH3+] + [HCO3−] + [CO2 ]f (19) Therefore, as it can be deduced from the fundamental
with the same reasoning for eqs 16 to 16-a and 8-a to 8-b, equations for ionic species activity in the electrolyte systems,
and as it has also been mentioned in the literature that the
[Base] = [RNH3+] + [HCO3−] (19-a) single ion activity cannot be measured experimentally using pH
Considering the charge balance equation (eq 9) and the and compared to predicted activity coefficient models for high
assumption that the concentrations of H3O+, OH−, and CO32− ionic strength solutions.22 However, Chan et al. displayed that
species are negligible compared to other species (i.e., pH ≅ 9), the pH measurement using a glass electrode even at high ionic
then the charge balance can be written as follows, strength comparing to the Pitzer model for single ion activity
coefficient of hydrogen ion is in the acceptable range of error
[RNH3+] = [HCO3−] + [RNHCOO−] (9-a) (generally less than 1.4%).23
Then, from eq 8-b and the simplified charge balance (eq 9-a), Speciation Comparisons. Figure 2 displays the speciation
the following useful equation can be derived: measured by the method developed in this work compared to

[Base] = [CO2 ]t + [HCO3−] (19-b)


Therefore, titration of a sample solution with strong base (e.g.,
NaOH) and using eq 19-b gives bicarbonate concentration.
Then, applying eqs 19-a and 8-b the protonated amine and
carbamate concentrations can be calculated, respectively. Using
the above, along with the alkalinity measurements (acid
titration), total inorganic carbon content of the solution and
using eq 16-b the free amine concentration can be determined.
Table S.1 (Supporting Information) shows the experimental
values for the concentration of the different species in MEA
solution with different CO2 loading. As it can be seen from
Table S.1 (Supporting Information), in terms of the carbamate,
protonated amine, and free amine, the agreement is good. In
the case of bicarbonate concentration, the values determined in
this work generally give higher values compared to the work
done by Jakobsen et al., which is discussed later in this section.8 Figure 2. MEA speciation as a function of carbon loading: this work at
Table S.2 (Supporting Information) displays the uncertainty in 21 °C, blank symbols; Jakobsen et al.,8 filled symbols; and Aspen Plus,
this work compared to literature values. dotted and filled lines at 20 °C.
Aspen Model Verification. For further verification of the
experimental method, Aspen Plus with the ENRTL-RK activity the Aspen Plus prediction and literature.8 The agreement
coefficient model was chosen to compare the results. Figure 1 between data produced in this work using total carbon and
HCO3− analytical measurement with the data from Jakobsen et
al.8 and the Aspen Plus prediction are fairly good. As can be
seen at the intermediate CO2 loading (i.e., 0.3 to 0.45) the
predicted values for the total bicarbonate and carbonate
concentration reveals slightly different values. At increased
CO2 loading, 0.45 molCO2/molMEA, the predicted values for the
three different sources come closer. The bicarbonate
concentration determined in this work is higher than the data
from Jakobsen et al.,8 particularly in the range of 0.3 to 0.45
CO2 loading. However, as it is discussed by Jakobsen et al.,8
their data shows the average uncertainties of 10.1% and 11.7%
and the maximum uncertainties were 29.2% and 34.4% for data
at 20 and 40 °C, respectively, in electroneutrality. They also
pointed out that because of the sensitivity of their results to the
Figure 1. Experimental pH from this work compared to the Aspen calibration method used, for the species with fractions <10%,
Plus predicted values, at 21 °C temperature. such as bicarbonate ion at low CO2 loadings, the quantitative
values are not valid.
shows the comparison of the experimental pH compared with As explained in the previous section, some approximations
AspenPlus ENRTL-RK model predicted values for the same were made for the method in this work. At low CO2 loading
MEA solution at different CO2 loadings. The model parameters (i.e. <0.3), the concentration of either carbonate or bicarbonate
in the Aspen Plus data bank were selected from recent work are very low. Therefore, at the CO2 loadings in the range 0−
done by Zhang et al.20 As it can be seen in Figure 1 at the 0.3, in fact, ignoring one of them (i.e. carbonate) and
working range for CO2 capture (i.e., CO2 loading of higher than incorporating its portion for the bicarbonate concentration
0.1 up to 0.5), the experimental pH values are in good does not affect the general trend of the speciation for four
6616 dx.doi.org/10.1021/ie300230k | Ind. Eng. Chem. Res. 2012, 51, 6613−6618
Industrial & Engineering Chemistry Research Article

major species concentrations in the solution including course, as it has been discussed by them, the mass balance in
bicarbonate (which is actually close to zero in this range), the reported values for MEA speciation by NMR spectroscopy
free amine, protonated amine, and carbamate. At the CO2 displays considerable deviation at 40 °C temperature (about
loading between 0.3 to about 0.45, the bicarbonate to carbonate 30% AAD).9 The speciation data reported by Böttinger et al.9
concentration ratio rises from 2 to 10. This ratio might not be did not include the free amine and protonated amine
high enough, so discarding the carbonate and/or incorporating concentration separately. Figure 3 shows the comparison of
its portion in the solution as the bicarbonate mainly does not the carbamate and total free amine plus protonated amine
affect the bicarbonate concentration in the general speciation concentrations determined in this work with data from
profile. With the above description and considering Figure 2, at Böttinger et al. showing good agreement.9
the CO2 loading higher than 0.45, ignoring the carbonate Since the experimental speciation was conducted at room
concentration compared to the other species is completely temperature, while CO2 absorption conditions are typically
reasonable. above this temperature, it was of interest how much
Figure 3 shows that the CO2 loading has a strong effect on temperature affects speciation. As can be deduced from the
speciation. As can be seen from Figure 3, and discussed in detail speciation data obtained by Jakobsen et al.8 at 20 and 40 °C
and mentioned in other published works,8,9,25 temperature has
a minimal effect on speciation. So, the same trends for the
amine speciation can be applied to typical absorption
temperatures. In terms of temperature effect, AspenPlus also
quantitatively shows no different patterns for speciation.
Therefore, based on this work and the literature, it can be
inferred that the speciation patterns has “no” or “very low”
temperature dependency.
Difference in Free Amine When Not Considering
Bicarbonate. Considering eq 16-b, the bicarbonate concen-
tration in the solution will affect the free amine concentration
in solution. Figure 4 shows the effects of bicarbonate

Figure 3. MEA speciation at 21 °C, this work, blank symbols;


Böttinger et al.,9 filled symbols, temperature 20 °C.

by Böttinger et al.,9 the species distribution of MEA−H2O−


CO2 shows that at low loadings practically all carbon dioxide is
chemically bound as carbamate. The carbamate concentration
increases linearly with carbon dioxide loading. With increasing
carbon dioxide loading, the concentration of protonated amine
increases while the unprotonated free amine concentration
decreases. This trend is almost linear up to a loading of about
0.5 molCO2/molMEA. At a loading of about 0.5 molCO2/molMEA,
the mole fraction of molecular amine has dropped to almost
zero. The amine is now either protonated or has reacted to
carbamate. With further increasing carbon dioxide loading, the Figure 4. Calculated free amine (MEA) concentration with and
concentration of the protonated amine continues to rise, but without considering the bicarbonate concentration term in eq 16-a at
this can only occur at the expense of the carbamate, so that its 21 °C.
concentration begins to drop. The carbon dioxide now released
from the carbamate reacts to bicarbonate.
In fact, in the reported experimental values for the MEA concentration in free amine concentration in the solution as a
speciation data by NMR spectroscopy, generally up to the CO2 function of CO2 loading using speciation in comparison to a
loading of about 0.4 (depending on the amine concentration), simple approximation where each CO2 reacts with 2 amines. As
the carbonate and bicarbonate mole fraction are close or equal it is obvious from Figure 4, the bicarbonate portion in the total
to zero.9 The values for the bicarbonate concentrations alkalinity of the solution is increased with CO2 loading. While
determined in this work has at least an order of magnitude at lean loading the impact is small, the difference at rich loading
difference with the reported experimental values mentioned in is substantial. For example, considering a loading of 0.4 molCO2/
the Figures 2 and 3 and Table S.1 in the Supporting molMEA, a 29% increase in the amount of free amine is
calculated.


Information, despite the acceptable prediction for free amine
and protonated and carbamate concentration at the normal
CO2 loading. Regardless, Aspen over predicts the bicarbonate CONCLUSIONS
concentration based on the ENRTL-RK model beyond even The versatile and quantitatively reliable method for amine
the values obtained in this work. The same over prediction by speciation determination is introduced. The procedure utilizes
extended UNIQUAC was mentioned by Faramarzi et al.24 Of two commonly used analytical methods for CO2 capture
6617 dx.doi.org/10.1021/ie300230k | Ind. Eng. Chem. Res. 2012, 51, 6613−6618
Industrial & Engineering Chemistry Research Article

solvents total alkalinity and total CO2 loading. Another titration (12) Derks, P. W. J.; Huttenhuis, P. J. G.; van Akena, C.; Marsman, J.
using strong base is used for bicarbonate concentration H.; Versteeg, G. F. Determination of the liquid-phase speciation in the
allowing for speciation to be determined. The quantitative MDEA-H2O-CO2 system. Energy Procedia 2011, 4, 599.
results for the different species concentration in MEA solution (13) Souchon, V.; de Oliveira Aleixo, M.; Delpoux, O.; Sagnard, C.;
Mougin, P.; Wender, A.; Raynal, L. In situ determination of species
is in good agreement with other available spectroscopic distribution in alkanolamine−H2O−CO2 systems by Raman spectros-
methods, mainly NMR, particularly for the free amine, copy. Energy Procedia 2011, 4, 554.
carbamate, and protonated amine concentrations. The (14) Shahi, P.; Hu, Y. F.; Chakma, A. Gas chromatographic analysis
AspenPlus predicted speciation also compared favorably with of acid gases and single/mixed alkanolamines. J. Chromatogr., A 1994,
the experimental data to further assess the data. 687, 121.


*
ASSOCIATED CONTENT
S Supporting Information
(15) Haji-Sulaiman, M. Z.; Aroua, M. K.; Md lIyas Pervez, Md.
Equilibrium concentration profiles of species in CO2−alkanolamine−
water systems. Gas Sep. Purif. 1996, 10, 13.
(16) Goyet, C.; Snover, A. K. High-accuracy measurements of total
Additional tables and figures. This material is available free of dissolved inorganic carbon in the ocean: Comparison of alternate
charge via the Internet at http://pubs.acs.org.


detection methods. Marine Chem. 1993, 44, 235.
(17) Aboudheir, A.; Tontiwachwuthikul, P.; Chakma, A.; Idem, R.
AUTHOR INFORMATION Kinetics of the reactive absorption of carbon dioxide in high CO2-
Corresponding Author loaded, concentrated aqueous monoethanolamine solutions. Chem.
*E-mail: kunlei.liu@uky.edu . Eng. Sci. 2003, 58, 5195.
(18) Edwards, T. J.; Maurer, G.; Newman, J.; Prausnitz, J. M. Vapor−
Notes liquid equilibria in multicomponent aqueous solutions of volatile weak
The authors declare no competing financial interest. electrolytes. AlChE J. 1978, 24, 966.

■ ACKNOWLEDGMENTS
The authors acknowledge the Carbon Management Research
(19) Austgen, D. M.; Rochelle, G. T.; Peng, X.; Chen, C. C. A model
of vapor−liquid equilibria in the aqueous acid gas−alkanolamine
system using the electrolyte-NRTL equation. New Orleans AIChE
Meeting, March 1988.
Group (CMRG) members, including Big Rivers Electric (20) Zhang, Y.; Que, H.; Chen, C.-C. Thermodynamic modeling for
Corporation, Duke Energy, East Kentucky Power Cooperative CO2 absorption in aqueous MEA solution with electrolyte NRTL
(EKPC), Electric Power Research Institute (EPRI), Illinois model. Fluid Phase Equilib. 2011, 311, 67.
Clean Coal Institute (ICCI), Kentucky Department of Energy (21) Nesbitt, H. W. pH-Electrode measurements of single-ion activity
Development and Independence (KY-DEDI), Kentucky Power coefficients consistent with a pH convention. Chem. Geol. 1981, 32,
(AEP), and LG&E and KU Energy, for their financial support. 207.


(22) Hessen, E. T.; Haug-Warberg, T.; Svendsen, H. F. The refined
REFERENCES e-NRTL model applied to CO2−H2O−alkanolamine systems. Chem.
Eng. Sci. 2010, 65, 3638.
(1) Figueroa, J. D.; Fout, T.; Plasynski, S.; McIlvried, H.; Srivastava, (23) Chan, C.-Y.; Eng, Y. W.; Eu, K.-S. Pitzer single-ion activity
R. D. Advances in CO2 capture technologyThe U.S. Department of coefficients and pH for aqueous solutions of potassium hydrogen
Energy’s Carbon Sequestration Program. Int. J. Greenhouse Gas Control phthalate in mixtures with KC1 and with NaCl at 298.15 K. J. Chem.
2008, 2, 9. Eng. Data 1995, 40, 685.
(2) Olajire, A. A. CO2 capture and separation technologies for end- (24) Faramarzi, L.; Kontogeorgis, G. M.; Thomsen, K.; Stenby, E. H.
of-pipe applicationsA review. Energy 2010, 35, 2610. Extended UNIQUAC model for thermodynamic modeling of CO2
(3) Kim, I.; Hoff, K. A.; Hessen, E. T.; Warberg, T. H.; Svendsen, H. absorption in aqueous alkanolamine solutions. Fluid Phase Equilib.
F. Enthalpy of absorption of CO2 with alkanolamine solutions 2009, 282, 121.
predicted from reaction equilibrium constants. Chem. Eng. Sci. 2009, (25) Cullinane, J. T.; Rochelle, G. T. Carbon dioxide absorption with
64, 2027. aqueous potassium carbonate promoted by piperazine. Chem. Eng. Sci.
(4) Dugas, R.; Rochelle, G. Absorption and desorption rates of 2004, 59, 3619.
carbon dioxide with monoethanolamine and piperazine. Energy
Procedia 2009, 1, 1163.
(5) Ermatchkov, V.; Maurer, G. Solubility of carbon dioxide in
aqueous solutions of N-methyldiethanolamine and piperazine:
Prediction and correlation. Fluid Phase Equilib. 2011, 302, 338.
(6) Haghtalab, A.; Shojaeian, A. Modeling solubility of acid gases in
alkanolamines using the nonelectrolyte Wilson-nonrandom factor
model. Fluid Phase Equilib. 2010, 289, 6.
(7) Dugas, R. E.; Rochelle, G. T. Modeling CO2 absorption into
concentrated aqueous monoethanolamine and piperazine. Chem. Eng.
Sci. 2011, 66, 5212.
(8) Jakobsen, J. P.; Krane, J; Svendsen, H. F. Liquid-phase
composition determination in CO2−H2O−alkanolamine systems: An
NMR study. Ind. Eng. Chem. Res. 2005, 44, 9894.
(9) Bö t tinger, W.; Maiwald, M.; Hasse, H. Online NMR
spectroscopic study of species distribution in MEA−H2O−CO2 and
DEA−H2O−CO2. Fluid Phase Equilib. 2008, 263, 131.
(10) Jakobsen, J. P.; da Silva, E. F; Krane, J.; Svendsen, H. F. NMR
study and quantum mechanical calculations on the 2-[(2-aminoethyl)-
amino]-ethanol−H2O−CO2 system. J. Magn. Reson. 2008, 191, 304.
(11) Jackson, P.; Robinson, K.; Puxty, G.; Attalla, M. In situ Fourier
transform-infrared (FT-IR) analysis of carbon dioxide absorption and
desorption in amine solutions. Energy Procedia 2009, 1, 985.

6618 dx.doi.org/10.1021/ie300230k | Ind. Eng. Chem. Res. 2012, 51, 6613−6618

You might also like