You are on page 1of 13

Ind. Eng. Chem. Res.

2011, 50, 163–175 163

Thermodynamic Modeling for CO2 Absorption in Aqueous MDEA Solution with


Electrolyte NRTL Model
Ying Zhang
AspenTech, Limited, Pudong, Shanghai 201203, People’s Republic of China

Chau-Chyun Chen*
Aspen Technology, Inc., Burlington, Massachusetts 01803

Accurate modeling of thermodynamic properties for CO2 absorption in aqueous alkanolamine solutions is
essential for the simulation and design of such CO2 capture processes. In this study, we use the Electrolyte
Nonrandom Two-Liquid activity coefficient model to develop a rigorous and thermodynamically consistent
representation for the MDEA-H2O-CO2 system. The vapor-liquid equilibrium (VLE), heat capacity, and
excess enthalpy data for the binary aqueous amine system are used to determine the NRTL interaction
parameters for the MDEA-H2O binary. The VLE, heat of absorption, heat capacity, and NMR spectroscopic
data for the MDEA-H2O-CO2 ternary system are used to identify the NRTL interaction parameters for the
molecule-electrolyte binaries and the previously unavailable standard-state properties of the amine ion, MDEA
protonate. The calculated VLE, heat of absorption, heat capacity, and the species concentrations for the
MDEA-H2O-CO2 system are compared favorably to experimental data.

1. Introduction Arcis et al.15 also fitted the VLE data with Pitzer’s equation
and used the thermodynamic model to estimate the enthalpy of
CO2 capture by absorption with aqueous alkanolamines is solution of CO2 in aqueous MDEA. Faramarzi et al.16 used the
considered an important technology to reduce CO2 emissions extended UNIQUAC model17 to represent VLE for CO2
from fossil-fuel-fired power plants and to help alleviate global absorption in aqueous MDEA, MEA, and mixtures of the two
climate change.1 Methyldiethanolamine (MDEA), which is an alkanolamines. Furthermore, they predicted the concentrations
alternative to monoethanolamine (MEA) for bulk CO2 removal, of the species in both MDEA and MEA solutions containing
has the advantage of relatively low heat of reaction of CO2 with CO2 and in the case of MEA, compared to NMR spectroscopic
MDEA.2 To properly simulate and design the absorption/ measurements.18,19
stripping processes with MDEA-based aqueous solvents, it is
essential to develop a sound process understanding of the In the present work, we expand the scope of the work of
transfer phenomena3 and accurate thermodynamic models4 to Austgen et al.8 and Posey9 to cover all thermodynamic proper-
calculate the driving forces for heat and mass transfer. In other ties. We use the 2009 version10 of the electrolyte NRTL
words, scalable simulation, design, and optimization of the CO2 model10-12 as the thermodynamic framework to correlate
capture processes start with modeling of the thermodynamic recently available experimental data for CO2 absorption in
properties, specifically vapor-liquid equilibrium (VLE) and aqueous MDEA solution. Much new data for thermodynamic
chemical reaction equilibrium, as well as calorimetric properties. properties and calorimetric properties have become available
in recent years, and they cover wider ranges of temperature,
Accurate modeling of thermodynamic properties requires
pressure, MDEA concentration, and CO2 loading. The binary
availability of reliable experimental data. Earlier literature
NRTL parameters for MDEA-water binary are regressed from
reviews5,6 suggested that, while there are extensive sets of
the binary VLE, excess enthalpy, and heat capacity data. The
experimental data available for the MDEA system, some of the
binary NRTL parameters for water-electrolyte pairs and
published CO2 solubility data for the aqueous MDEA system
MDEA-electrolyte pairs and the standard-state properties of
may be questionable. The use of a thermodynamically consistent
protonated MDEA ion are obtained by fitting to the ternary VLE,
framework makes it possible to correlate available experimental
heat of absorption, heat capacity, and NMR spectroscopic data.
data, to screen out questionable data, and to morph these diverse
This expanded model should provide a comprehensive thermo-
and disparate data into a useful and thermodynamically con-
dynamic representation for the MDEA-H2O-CO2 system over
sistent form for process modeling and simulation.
a broader range of conditions and give more reliable predictions
Excess Gibbs energy-based activity coefficient models pro-
over previous works.
vide a practical and rigorous thermodynamic framework to
model thermodynamic properties of aqueous electrolyte systems, In conjunction with the use of the electrolyte NRTL model
including aqueous alkanolamine systems for CO2 capture.4,7 For for the liquid-phase activity coefficients, we use the PC-SAFT20,21
example, Austgen et al.8 and Posey9 applied the electrolyte equation of state (EOS) for the vapor-phase fugacity coefficients.
NRTL model10-12 to correlate CO2 solubility in aqueous MDEA While both PC-SAFT EOS and typical cubic EOS would give
solution and other aqueous alkanolamines. Kuranov et al.,5 reliable fugacity calculations at low to medium pressures, we
Kamps et al.,6 and Ermatchkov et al.13 used Pitzer’s equation14 choose PC-SAFT for its ability to model vapor-phase fugacity
to correlate the VLE data of the MDEA-H2O-CO2 system. coefficients at high pressures, which is an important consider-
ation for modeling CO2 compression. The PC-SAFT parameters
* To whom correspondence should be addressed. Tel.: 781-221-6420. used in this model are given in Table 1. The parameters for
Fax: 781-221-6410. E-mail: chauchyun.chen@aspentech.com. water and CO2 are taken from the literature21 and the Aspen
10.1021/ie1006855  2011 American Chemical Society
Published on Web 08/05/2010
164 Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011

Table 1. Parameters for PC-SAFT Equation of State Table 2. Parameters of the Characteristic Volume for the
Brelvi-O’Connell Modela
MDEA H2O CO2
Characteristic Volume (m3/kmol)
source this work Gross and Aspen
Sadowski 21 Databank22 parameter MDEA H2O CO2
segment number 3.3044 1.0656 2.5692 24
source this work Brelvi and O’Connell Yan and Chen25
parameter, m
V1,i 0.369b 0.0464 0.175
segment energy 237.44 K 366.51 K 152.10 K
V2,i 0 0 -3.38 × 10-4
parameter, ε
segment size 3.5975 Å 3.0007 Å 2.5637 Å a
The Brelvi-O’Connell model has been described in ref 24. The
parameter, σ correlation of the characteristic volume for the Brelvi-O’Connell model
association energy 3709.9 K 2500.7 K 0K (ViBO) is given as follows: ViBO ) V1,i + V2,iT, where T is the temperature
parameter, εAB (given in Kelvin). b Here, the critical volume was used as the
kAB 0.066454 Å3 0.034868 Å3 0 Å3 characteristic volume for MDEA.

Databank.22 The parameters for MDEA are obtained from the Table 3. Parameters for Henry’s Constant (Expressed in Units of Pa)a
regression of experimental data on vapor pressure, liquid density, solute i CO2 CO2
and liquid heat capacity. solvent j H2O MDEA
source Yan and Chen25 this work
aij 91.344 19.8933
2. Thermodynamic Framework bij -5876.0 -1072.7
cij -8.598 0.0
2.1. Chemical and Phase Equilibrium. CO2 solubility in dij -0.012 0.0
aqueous amine solutions is determined by both its physical
solubility and the chemical equilibrium for the aqueous phase
a
The correlation for Henry’s constant is given as follows: ln Hij ) aij
+ bij/T + cij ln T + dijT, where T is the temperature (given in Kelvin).
reactions among CO2, water, and amines.
2.1.1. Physical Solubility. Physical solubility is the equi- Henry’s law constants for CO2 with water and for CO2 with
librium between gaseous CO2 molecules and CO2 molecules in MDEA are required. The former has been extensively studied,25
the aqueous amine solutions: although knowledge about the latter is relatively limited.
Because it is not feasible to directly measure CO2 physical
CO2(V) T CO2(l) (1)
solubility in pure amines, because of the reactions between them,
the common practice is to derive the CO2 physical solubility in
It can be expressed by Henry’s law:
amines from that of N2O for their analogy in molecular structure
PyCO2φCO2 ) HCO2xCO2γ*CO2 (2) and, thus, in physical solubility as believed:26

HCO2,MDEA HCO2,water
where P is the system pressure, yCO2 the mole fraction of CO2 ) (5)
in the vapor phase, φCO2 the CO2 fugacity coefficient in the vapor HN2O,MDEA HN2O,water
phase, HCO2 the Henry’s law constant of CO2 in the mixed
solvent of water and amine, xCO2 the equilibrium CO2 mole In 1992, Wang et al.27 reported the solubility of N2O in pure
fraction in the liquid phase, and γ*CO2 the unsymmetric activity MDEA solvent as follows:
coefficient of CO2 in the mixed solvent of water and amine.
-1312.7
The Henry’s constant in the mixed solvent can be calculated
from those in the pure solvents:23
HN2O,MDEA (kPa m3 kmol-1) ) (1.524 × 105) exp
T ( )
(6)

ln
() Hi
γ∞i
) ∑x
A
A ln
( )
HiA

γiA
(3) Based on the work of Versteef and van Swaaij,28 we obtained
the following two equations for the solubilities of N2O and CO2
in water:
where Hi is the Henry’s constant of supercritical component i
in the mixed solvent, HiA the Henry’s constant of supercritical -2284
component i in pure solvent A, γ∞i the infinite dilution activity
HN2O,water (kPa m3 kmol-1) ) (8.5470 × 106) exp
T ( )
coefficient of supercritical component i in the mixed solvent, (7)

γiA the infinite dilution activity coefficient of supercritical
component i in pure solvent A, and xA the mole fraction of -2044
solvent A.
HCO2,water (kPa m3 kmol-1) ) (2.8249 × 106) exp
T ( )
We use wA in lieu of xA in eq 3 to weigh the contributions (8)
from different solvents.22 The parameter wA is calculated using
eq 4: We use eqs 5-8 to determine HCO2,MDEA and the parameters
are summarized in Table 3.
∞ 2/3 The Henry’s constant of CO2 in pure solvent A is corrected
xA(ViA )
wA ) (4) with the Poynting term for pressure:25
∑ x (V B
∞ 2/3
iB)


B
HCO2,A(T, P) ) HCO2,A(T, po,l
A ) exp ( RT1 ∫ p◦,l
A
P ∞
VCO2,A
dp ) (9)
Here, ViA represents the partial molar volume of supercritical

component i at infinite dilution in pure solvent A. ViA is where HCO2,A(T,P) is the Henry’s constant of CO2 in pure solvent
24
calculated from the Brelvi-O’Connell model with the char- A at system temperature and pressure, HCO2,A(T,p°,lA ) the Henry’s
BO
acteristic volume for the solute (VCO 2
) and solvent (VsBO), which constant of CO2 in pure solvent A at system temperature and

are listed in Table 2. the solvent vapor pressure, and VCO 2,A
the partial molar volume
Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011 165
Table 4. Parameters for the Reference States Properties
ig ig ∞,aq ∞,aq
property ∆fG298.15 (J/kmol) ∆fH298.15 (J/kmol) ∆fG298.15 (J/kmol) ∆fH298.15 (J/kmol) source
MDEA -1.6900 × 10 8
-3.8000 × 10 8
Aspen Databank22
H2O -2.2877 × 108 -2.4200 × 108 Aspen Databank22
CO2 -3.9437 × 108 -3.9351 × 108 Aspen Databank22
H3O+ -2.3713 × 108 -2.8583 × 108 Aspen Databank22
OH- -1.5724 × 108 -2.2999 × 108 Wagman et al.29
HCO-3 -5.8677 × 108 -6.9199 × 108 Wagman et al.29
CO2-
3 -5.2781 × 108 -6.7714 × 108 Wagman et al.29
MDEAH+ -2.5989 × 108 a -5.1422 × 108 a this work
a
The values of MDEAH+ are calculated from the chemical equilibrium constant in Kamps and Maurer,30 which are used as the initial guess to fit
experimental data.

of CO2 at infinite dilution in pure solvent A calculated from Table 5. Parameters for Ideal Gas Heat Capacity
the Brelvi-O’Connell model. Heat Capacity (J/(kmol K))
At low pressures, the Poynting correction is almost unity and parameter MDEA H2O CO2
can be ignored.
22
2.1.2. Aqueous-Phase Chemical Equilibrium. The aqueous- source this work Aspen Databank Aspen Databank22
C1i 2.7303 × 104 3.3738 × 104 1.9795 × 104
phase chemical reactions involved in the MDEA-water-CO2 C2i 5.4087 × 102 -7.0176 7.3437 × 10
system can be expressed as C3i 0 2.7296 × 10-2 -5.6019 × 10-2
C4i 0 -1.6647 × 10-5 1.7153 × 10-5
2H2O T H3O+ + OH- (10) C5i 0 4.2976 × 10-9 0
C6i 0 -4.1696 × 10-13 0
C7i 278 200 300
CO2 + 2H2O T H3O+ + HCO-
3 (11) C8i 397 3000 1088.6
a
The correlation for the ideal gas heat capacity is given as follows:
HCO- +
3 + H2O T H3O + CO3
2-
(12) p ) C1i + C2iT + C3iT + C4iT + C5iT + C6iT , C7i < T < C8i, where
Cig 2 3 4 5

T is the temperature (given in Kelvin).


MDEAH+ + H2O T H3O+ + MDEA (13)
ig ig
The reference-state properties, ∆fGs,298.15 and ∆fHs,298.15 , are
We calculate the equilibrium constants of the reaction from shown in Table 4. The ideal gas heat capacities are shown in
the reference-state Gibbs free energies of the participating Table 5. For water, the Gibbs free energy departure function is
components: obtained from the ASME steam tables. For MDEA, the
departure function is calculated from the PC-SAFT equation
-RT ln Kj ) ∆Goj (T) (14) of state.
For molecular solute CO2, the Gibbs free energy in aqueous
where Kj is the equilibrium constant of reaction j, ∆G°(T)
j the phase infinite dilution is calculated from Henry’s law:
reference-state Gibbs free energy change for reaction j at
temperature T, R the universal gas constant, and T the system
temperature.
For the aqueous phase reactions, the reference states chosen
G∞,aq
i (T) ) ∆fGig
i (T) + RT ln ( )Hi,w
Pref
(17)

are pure liquid for the solvents (water and MDEA), and aqueous where G∞,aq (T) is the mole fraction scale aqueous-phase infinite
i
phase infinite dilution for the solutes (ionic and molecular). dilution Gibbs free energy of solute i at temperature T, ∆fGig
i (T)
The Gibbs free energy of solvents is calculated from that of the ideal gas Gibbs free energy of formation of solute i at
ideal gas and the departure function: temperature T, Hi,w the Henry’s constant of solute i in water,
and Pref the reference pressure.
Gs(T) ) Gsig(T) + ∆Gsigfl(T) (15)
For ionic species, the Gibbs free energy in aqueous-phase
infinite dilution is calculated from the Gibbs free energy of
where Gs(T) is the Gibbs free energy of solvent s at temperature
formation in aqueous-phase infinite dilution at 298.15 K, the
T, Gsig(T) the ideal gas Gibbs free energy of solvent s at
enthalpy of formation in aqueous-phase infinite dilution at
temperature T, and ∆Gsigfl(T) the Gibbs free energy departure
298.15 K, and the heat capacity in aqueous-phase infinite
from ideal gas to liquid at temperature T.
dilution:
The Gibbs free energy of an ideal gas is calculated from the
Gibbs free energy of formation of an ideal gas at 298.15 K, the

T
enthalpy of formation of an ideal gas at 298.15 K, and the ideal G∞,aq
i (T) ) ∆fH∞,aq
i,298.15 + C∞,aq
p,i dT - T ×

( )
298.15

( )
gas heat capacity. ∆fH∞,aq ∞,aq
i,298.15 - ∆fGi,298.15 C∞,aq
∫ 1000
T p,i
+ dT + RT ln
∫ 298.15
T 298.15 T Mw
Gsig(T) ) ∆fHs,298.15
ig
+ Cig dT
298.15 p,s
-T×
(18)

( -
)
ig ig ig
∆fHs,298.15 ∆fGs,298.15 Cp,s

T
+ dT (16) Here, G∞,aq (T) is the mole fraction scale aqueous-phase infinite
298.15 298.15 T i
dilution Gibbs free energy of solute i at temperature T, ∆fG∞,aq
i,298.15
where Gsig(T) is the ideal gas Gibbs free energy of solvent s at the molality scale aqueous-phase infinite dilution Gibbs free
ig ∞,aq
temperature T, ∆fGs,298.15 the ideal gas Gibbs free energy of energy of formation of solute i at 298.15 K, ∆fHi,298.15 the
ig
formation of solvent s at 298.15 K, ∆fHs,298.15 the ideal gas aqueous phase infinite dilution enthalpy of formation of solute
enthalpy of formation of solvent s at 298.15 K, and Cp,s ig
the i at 298.15 K, and C∞,aq
p,i the aqueous-phase infinite dilution heat
ideal gas heat capacity of solvent s. capacity of solute i. The term RT ln (1000/Mw) is added because
166 Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011

Table 6. Parameters for Aqueous-Phase Infinite Dilution Heat Capacitya


Heat Capacity (J/(kmol K))
parameter H3O+ OH- HCO-
3 CO2-
3 MDEAH+
source Aspen Databank22 Aspen Databank22 Criss and Cobble31 Criss and Cobble31 this work
C1 7.5291 × 104 -1.4845 × 105 -2.9260 × 104 b -3.9710 × 105 b 2.9900 × 105 b
∞,aq ∞,aq
a
The aqueous-phase infinite dilution heat capacity is assumed to be constant (Cp,i ) C1). b The Cp,i value of MDEAH+ is calculated from the
∞,aq
30
chemical equilibrium constant in Kamps and Maurer, which is used as the initial guess to fit experimental data. The Cp,i values of HCO- 2-
3 and CO3
are the average values of heat capacity between 298 K and 473 K (taken from Criss and Cobble31).

∞,aq Table 7. Parameters for Heat of Vaporization (Expressed in Units


∆fGi,298.15 , as reported in the literature, is based on molality
of J/kmol)a
concentration scale while G∞,aqi is based on mole fraction scale.
∞,aq ∞,aq ∞,aq component i MDEA
The standard-state properties ∆fGi,298.15 , ∆fHi,298.15 , and Cp,i
are available in the literature for most ionic species, except those source this work
C1i 9.7381 × 107
of MDEAH+. (See Tables 4 and 6.) We calculate the reference- C2i 4.6391 × 10-1
state properties of the MDEAH+ ion from the experimental C3i 0
equilibrium constant of eq 13, as reported in 1996 by Kamps C4i 0
∞,aq ∞,aq
and Maurer.30 The calculated ∆fGi,298.15 and ∆fHi,298.15 values C5i 0
∞,aq
are given in Table 4, and the calculated Cp,i values are given Tci 741.9b
a
in Table 6. As will be shown later, we use these calculated The DIPPR equation for the heat of vaporization is given as follows:
reference-state properties for MDEAH+ as part of the adjustable ∆vapHi ) C1i(1 - Tri)Z, where Z ) C2i + C3iTri + C4iTri2 + C5iTri3 and
Tri ) T/Tci (here, Tci is the critical temperature of component i). The
parameters in the fitting experimental data of thermodynamic temperatures are given in Kelvin. b The Tci value for MDEA is obtained
properties, including VLE, heat of solution, heat capacity, and from Von Niederhausern et al.32
species concentration from NMR spectra. Also given in Table
∞,aq
6 are estimated values of Cp,i for HCO3- and CO32-. They have The liquid enthalpy of pure water is calculated from the ideal
been taken from the 1964 work of Criss and Cobble.31 gas model and the ASME Steam Tables EOS for enthalpy
2.2. Heat of Absorption and Heat Capacity. The CO2 heat departure:
of absorption in aqueous MDEA solutions can be derived from

an enthalpy balance of the absorption process: T
Hwl (T) ) ∆fHw,298.15
ig
+ 298.15
ig
Cp,w dT + ∆Hwigfl(T, p) (21)
l
nFinalHFinal - nInitialHInitial
l
- nCO2HCO
g

∆Habs )
2
(19) where Hwl (T) is the liquid enthalpy of water at temperature T,
nCO2 ig
∆fHw,298.15 the ideal gas enthalpy of formation of water at 298.15
ig
K, Cp,w the ideal-gas heat capacity of water, and ∆Hwigfl(T,p)
l the enthalpy departure calculated from the ASME Steam Tables
where ∆Habs is the heat of absorption per mole of CO2, HFinal
the molar enthalpy of the final solution, HlInitial the molar enthalpy EOS.
g
of the initial solution, HCO2 the molar enthalpy of gaseous CO2 Liquid enthalpy of the nonaqueous solvent s is calculated
absorbed, nFinal the number of moles of the final solution, nInitial from the ideal-gas enthalpy of formation at 298.15 K, the ideal-
the number of moles of the initial solution, and nCO2 the number gas heat capacity, the vapor enthalpy departure, and the heat of
of moles of CO2 absorbed. vaporization:
There are two types of heat of absorption: integral heat of

T
absorption and differential heat of absorption. The integral heat Hls(T) ) ∆fHs,298.15
ig
+ ig
Cp,s dT + ∆HVs (T, p) - ∆vapHs(T)
298.15
of absorption for a certain amine-H2O-CO2 system refers to
(22)
the heat effect per mole of CO2 during the CO2 loading of the
amine solution increasing from zero to the final CO2 loading Here, Hsl(T) is the liquid enthalpy of solvent s at temperature T,
value of that amine-H2O-CO2 system. The differential heat ig
∆fHs,298.15 the ideal-gas enthalpy of formation of solvent s at
of absorption for an amine-H2O-CO2 system refers to the heat 298.15 K, Cigp,s the ideal-gas heat capacity of solvent s, ∆HVs (T,p)
effect per mole of CO2 if a very small amount of CO2 is added the vapor enthalpy departure of solvent s, and ∆vapHs(T) the
into this amine-H2O-CO2 system. heat of vaporization of solvent s.
For calculation of both types of heat of absorption, enthalpy The PC-SAFT EOS is used for the vapor enthalpy departure
calculations for the initial and final amine-H2O-CO2 systems and the DIPPR heat of vaporization correlation is used for the
and for gaseous CO2 are required. The heat capacity of the heat of vaporization. Table 7 shows the DIPPR equation and
MDEA-H2O-CO2 system can be calculated from the temper- the correlation parameters for the heat of vaporization.
ature derivative of enthalpy. The enthalpies of ionic solutes in aqueous phase infinite
We use the following equation for liquid enthalpy: dilution are calculated from the enthalpy of formation at 298.15
K in aqueous-phase infinite dilution and the heat capacity in
Hl ) xwHwl + xsHsl + ∑xH i
∞,aq
i + Hex (20) aqueous-phase infinite dilution:
i


T
Here, Hl is the molar enthalpy of the liquid mixture, Hwl the H∞,aq
i (T) ) ∆fH∞,aq
i,298.15 + 298.15
C∞,aq
p,i dT (23)
molar enthalpy of liquid water, Hsl the molar enthalpy of liquid
nonaqueous solvent s, H∞,aq
i the molar enthalpy of solute i where H∞,aq
i (T) is the enthalpy of solute i in aqueous-phase
∞,aq
(molecular or ionic) in aqueous-phase infinite dilution, and Hex infinite dilution at temperature T, ∆fHi,298.15 the enthalpy of
the molar excess enthalpy. The terms xw, xs, and xi represent formation of solute i in aqueous-phase infinite dilution at 298.15
∞,aq
the mole fractions of water, nonaqueous solvent s, and solute i, K, and Cp,i the heat capacity of solute i in aqueous-phase
respectively. infinite dilution.
Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011 167
a
Table 8. Parameters for Dielectric Constant phase chemical equilibrium calculations, we choose the aqueous-
component i MDEA phase infinite dilution reference state for molecular solute CO2
source Aspen Databank22 and all ionic species.
Ai 21.9957 In applying the electrolyte NRTL model for liquid-phase
Bi 8992.68
Ci 298.15 activity coefficient calculations, the binary NRTL interaction
a
parameters for molecule-molecule binary, molecule-electrolyte
The correlation for the dielectric constant is given as follows: εi(T) binary, and electrolyte-electrolyte binary systems are required.
) Ai + Bi[(1/T) - (1/Ci)], where T is the temperature (given in Kelvin).
Here, electrolytes are defined as cation and anion pairs. In
∞,aq ∞,aq addition, solvent dielectric constants are needed to facilitate
Both ∆fHi,298.15 and Cp,i are also used in the calculation of
calculations of long-range ion-ion interaction contribution to
Gibbs free energy of the solutes, thus impacting chemical
∞,aq ∞,aq activity coefficients. Table 8 shows the dielectric constant
equilibrium calculations. In this study, ∆fHi,298.15 and Cp,i for
+ correlation used in this work for MDEA.
MDEAH are determined by fitting to the experimental phase
equilibrium data, the heat of solution data, and the speciation Unless specified otherwise, all molecule-molecule binary
data, together with molality scale Gibbs free energy of formation parameters and electrolyte-electrolyte binary parameters are
∞,aq
at 298.15 K, ∆fGi,298.15 , and NRTL interaction parameters. defaulted to zero. All molecule-electrolyte binary parameters are
The enthalpies of molecular solutes in aqueous phase infinite defaulted to (8,-4), average values of the parameters as reported
dilution are calculated from Henry’s law: for the electrolyte NRTL model.12 The nonrandomness factor (R)
is fixed at 0.2. The calculated thermodynamic properties of the

H∞,aq
i (T) ) ∆fHig
i (T) - RT
2
( ∂ ln Hi,w
∂T ) (24)
electrolyte solution are dominated by the binary NRTL parameters
associated with the major species in the system. In other
words, the binary parameters for the water-MDEA binary, the
where ∆fHigi (T) is the ideal gas enthalpy of formation of solute water-(MDEAH+, HCO- + 2-
3 ) binary, the water-(MDEAH , CO3 )
i at temperature T, and Hi,w Henry’s constant of solute i in water. + -
binary, and the MDEA-(MDEAH , HCO3 ) binary systems
Excess enthalpy (Hex) is calculated from the activity coef- determine the calculated thermodynamic properties. These binary
ficient model (i.e., the electrolyte NRTL model). parameters, in turn, are identified from fitting to available experi-
2.3. Activity Coefficients. Activity coefficients are required mental data.
in phase equilibrium calculations, aqueous-phase chemical
equilibrium calculations, heat of absorption, liquid heat capacity,
3. Modelling Results
and liquid enthalpy calculations. The activity coefficient of a
component in a liquid mixture is a function of temperature, Table 9 summarizes the model parameters and sources of
pressure, mixture composition, and choice of reference state. the parameters used in the thermodynamic model. Most of the
In VLE calculations, we use the asymmetric mixed-solvent parameters can be obtained from the literature. The remaining
reference state for the molecular solute CO2, and in aqueous- parameters are determined by fitting to the experimental data.

Table 9. Parameters Estimated in Modeling


parameter component source data used for regression
Antoine equation MDEA regression vapor pressure of MDEA
∆vapH MDEA regression heat of vaporization of MDEA, calculated from the vapor
pressure using the Clausius-Clapeyron equation
dielectric constant MDEA Aspen Databank22
Henry’s constant CO2 in H2O Yan and Chen25
CO2 in MDEA this work
NRTL binary parameters CO2-H2O binary Yan and Chen25
MDEA-H2O binary regression VLE, excess enthalpy, and heat capacity for the MDEA-H2O binary
molecule-electrolyte binaries regression VLE, excess enthalpy, heat capacity, and species concentration from
NMR spectra for the MDEA-H2O-CO2 system
ig
∆fG298.15 H2O, MDEA, CO2 Aspen Databank22
ig
∆fH298.15 H2O, MDEA, CO2 Aspen Databank22
Cig
p H2O, CO2 Aspen Databank22
MDEA regression liquid heat capacity of MDEA
∞,aq
∆fG298.15 H3O+, OH-, HCO- 2-
3 , CO3 Aspen Databank22
MDEAH+ regression VLE, excess enthalpy, heat capacity, and species concentration from
NMR spectra for the MDEA-H2O-CO2 system
∞,aq
∆fH298.15 H3O+, OH-, HCO- 2-
3 , CO3 Aspen Databank22
MDEAH+ regression VLE, excess enthalpy, heat capacity, and species concentration from
NMR spectra for the MDEA-H2O-CO2 system
C∞,aq
p H3O+, OH- Aspen Databank22
HCO- 2-
3 , CO3 Criss and Cobble31
MDEAH+ regression VLE, excess enthalpy, heat capacity, and species concentration from
NMR spectra for the MDEA-H2O-CO2 system

Table 10. Experimental Data Used in the Regression for Pure MDEA
data type temperature, T (K) pressure, P (kPa) data points average relative deviation, |∆Y/Y| (%) reference
vapor pressure 293-401 0.0006-1.48 26 1.5 Daubert et al.33
vapor pressure 420-513 3.69-90.4 14 4.0 Noll et al.34
vapor pressure 420-738 3.69-3985 23 2.9 VonNiederhausern et al.32
liquid heat capacity 299-397 5 0.5 Maham et al.35
liquid heat capacity 303-353 11 0.4 Chen et al.36
liquid heat capacity 278-368 19 0.3 Zhang et al.37
168 Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011

Table 11. Antoine Equation Parameters for Pure MDEAa Table 13. Regressed NRTL Parameters for the MDEA-H2O Binary
System with r ) 0.2a
parameter component i value
parameter component i component j value standard deviation
C1i MDEA 1.2276 × 102
C2i MDEA -1.3253 × 104 aij H2O MDEA 8.5092 0.1641
C3i MDEA -1.3839 × 10 aij MDEA H2O -1.7141 0.0566
C4i MDEA 3.20 × 10-6 bij H2O MDEA -1573.9 45.70
a bij MDEA H2O -261.85 22.97
The correlation for the Antoine equation is given as follows: ln Pi*,l
) C1i + C2i/T + C3i ln T + C4iT2, where T is the temperature (given in a
Correlation for the NRTL parameters: τij ) aij + bij/T, where T is
Kelvin). the temperature (given in Kelvin).

3.1. MDEA. Extensive experimental vapor pressure data and


liquid heat capacity data are available for MDEA. The data used
in the regression for MDEA and the correlation results are
summarized in Table 10.
Table 11 shows the Antoine equation parameters regressed from
the recently available vapor pressure data.32-34 The heat of
vaporization (from 293 K to 473 K) generated with the regressed
Antoine equation parameters through the Clausius-Clapeyron
equation are used to determine the DIPPR heat of vaporization
equation parameters (shown in Table 7). The ideal-gas heat capacity
correlation parameters are obtained by fitting to the liquid heat
capacity data35-37 (shown in Table 5). Table 10 shows the excellent
correlation of the experimental data for vapor pressure, with an
average relative deviation of <4%, and liquid heat capacity, with
an average relative deviation of <0.5%.
The PC-SAFT parameters of MDEA (shown in Table 1) are
regressed from the vapor pressure data32-34 (with an average Figure 1. Parity plot for the MDEA-H2O system total pressure, experiment
versus model: (4) Xu et al.,40 (O) Voutsas et al.,41 and (0) Kim et al.42
relative deviation of 13.1%), the liquid heat capacity data35-37
(with an average relative deviation of 22.6%), and the liquid
density data38,39 (with an average relative deviation of 2.7%).
3.2. MDEA-H2O System. Extensive literature data on
VLE,40-42 excess enthalpy,9,35,43 and heat capacity36,37,44 of the
MDEA-H2O binary system are available (see Table 12). These
data cover the complete MDEA-H2O binary concentration
range from room temperature to 458 K. Together, all of these
data are used to identify the NRTL binary parameters, including
their temperature dependencies for the MDEA-H2O binary
system.
The regressed NRTL parameters are summarized in Table
13. The experimental data for the binary MDEA-H2O system
are well-represented. The average relative deviations between
the calculated values and the experimental data are summarized
in Table 12. Figure 1 shows the parity plot, while Figure 2 shows
the comparison for the experimental total pressure data and the
calculated results from the model. Figure 3 shows the compari- Figure 2. Comparison of the experimental data from Kim et al.42
son results for the MDEA vapor composition. The excess (represented by symbols: (O) T ) 373 K, (4) T ) 353 K, (0) T ) 333 K,
and (×) T ) 313 K) for total pressure of the MDEA-H2O binary solution
enthalpy fit is given in Figure 4. Both the experimental excess and the model results (represented by lines).
enthalpy data from Posey9 and those of Maham et al.35,43 are
represented very well. Figure 5 shows the model also provides activity coefficient remains relatively constant, the model
satisfactory representation of the heat capacity data. suggests that the MDEA activity coefficient varies strongly with
Figure 6 shows the model predictions for water and MDEA MDEA concentration and temperature, especially in dilute
activity coefficients at 313, 353, and 393 K. While the water aqueous MDEA solutions.
Table 12. Experimental Data Used in the Regression for the MDEA-H2O System
data average relative
data type temperature, T (K) pressure, P (kPa) MDEA mole fraction points deviation, |∆Y/Y| (%) reference
a
VLE (isoconcentration), TP 326-381 13-101 0.016-0.26 34 3.4 Xu et al.40
VLE (isobaric), Tx 349-458 40.0-66.7 0-0.93 30 3.3a Voutsas et al.41
VLE (isothermal), TPxy 313-373 6-100 0-0.36 61 3.0a Kim et al.42
excess enthalpy (isothermal) 298, 342 0.02-0.74 19 6.3 Posey9
excess enthalpy (isothermal) 298, 303 0.03-0.93 26 5.4 Maham et al.35
excess enthalpy (isothermal) 338 0.10-0.90 9 11.2 Maham et al.43
heat capacity (isobaric) 303-353 100 0.13-0.80 44 2.2 Chiu and Li44
heat capacity (isobaric) 303-353 100 0.20-0.80 44 2.2 Chen et al.36
heat capacity (isobaric) 278-368 100 0-1.0 228 2.8 Zhang et al.37
a
The average relative deviation is that of total pressure.
Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011 169

Figure 3. Comparison of the experimental data from Kim et al.42


(represented by symbols: (O) T ) 373 K, (4) T ) 353 K, (0) T ) 333 K, Figure 6. Model predictions of water and MDEA activity coefficients at
and (×) T ) 313 K) for the vapor composition of the MDEA-H2O binary 313, 353, and 393 K over the entire mole fraction range; solid lines represent
solution and the model results (represented by lines). water activity coefficients and dashed lines represent MDEA activity
coefficients.

∞,aq ∞,aq
The terms ∆fG298.15 , ∆fH298.15 , and Cp∞,aq of MDEAH+ and
the binary NRTL parameters for major molecule-electrolyte
pairs are regressed from selected experimental data of the
MDEA-H2O-CO2 system. Table 14 summarizes the VLE,5,6,13
heat of absorption,65,66 heat capacity,67 and species concentra-
tion68 data used to obtain these parameters.
Species concentration data from NMR spectra are very useful
to validate the model predictions for the species distribution in
the ternary system. Calculated heat of absorption of CO2 by
the MDEA solution also strongly depends on the species
distribution.
For VLE data, we choose the total pressure data of Kuranov
et al.,5 Kamps et al.,6 and the CO2 partial pressure data of
Ermatchkov et al.13 in the regression. Together, these data cover
Figure 4. Comparison of the experimental data from Posey9 (represented temperatures from 313 K to 413 K, pressures from 0.1 kPa to
by full symbols: (b) T ) 298 K, ([) T ) 342 K) and Maham et al.35,43 6000 kPa, MDEA mole fractions from 0.03 to 0.13, and CO2
(represented by empty symbols: (O) T ) 298 K, (4) T ) 313 K, (0) T )
338 K) for excess enthalpy of the MDEA-H2O binary solution and the
loadings from 0.003 to 1.32. The CO2 partial pressure data of
model results (represented by lines). Jou et al.45 also cover wide ranges for temperature, pressure,
MDEA concentration, and CO2 loading. However, considering
the reported inconsistency5,6,13 between these data45 and those
of Kuranov et al.,5 Kamps et al.,6 and Ermatchkov et al.,13 we
choose to exclude the data of Jou et al.45 from the regression.
The Jou et al. data45 and all other available literature VLE
data8,46-64 are used only for model validation.
The average relative deviations between the correlation results
and the various experimental data are shown in Table 14. The
regressed parameters for the MDEA-H2O-CO2 system are
summarized in Table 15. As expected, the regressed values of
∞,aq ∞,aq
∆fG298.15 , ∆fH298.15 , and Cp∞,aq for MDEAH+ in Table 15 are
comparably close to the estimated values reported in Tables 4
and 6.
Figures 7 and 8 show that most of the total pressure data of
Kuranov et al.5 and Kamps et al.6 are fitted within (20%.
Figures 9-11 show the excellent correlation results for the total
pressure data for MDEA concentration from 2 m to 8 m, CO2
Figure 5. Comparison of the experimental data from Chen et al.36 loading from 0.11 to 1.32, temperature from 313 K to 413 K,
(represented by symbols: (O) MDEA mole fraction ) 0.8, (4) MDEA mole and pressure up to 6000 kPa. Figures 12 and 13 show that most
fraction ) 0.6, (0) MDEA mole fraction ) 0.4, and (×) MDEA mole
fraction ) 0.2) for heat capacity of the MDEA-H2O binary solution and
of the CO2 partial pressure data of Ermatchkov et al.13 are fitted
the model results (represented by lines). within (30%. Figure 12 suggests that there is a slight systematic
deviation that changes from negative to positive as the CO2
3.3. MDEA-H2O-CO2 System. Extensive VLE,5,6,8,13,45-64 loading increases. Figures 14-16 show the satisfactory cor-
heat of absorption,65,66 heat capacity,67 and NMR spectro- relation results for the CO2 partial pressure data for MDEA
scopic68 data of the ternary MDEA-H2O-CO2 system are concentration from 2 m to 8 m, CO2 loading from 0.003 to 0.78,
available. temperature from 313 K to 393 K, and pressure from 0.1 kPa
170 Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011

Table 14. Experimental Data Used in the Regression for the MDEA-H2O-CO2 System
temperature, pressure, MDEA data average relative
data type T (K) P (kPa) mole fraction CO2 loading points deviation, |∆Y/Y| (%) reference
VLE, TPx, total pressure 313-413 70-5000 0.035-0.067 0-1.32 82 6.8 Kuranov et al.5
VLE, TPx, total pressure 313-393 200-6000 0.126 0.13-1.15 23 10.5 Kamps et al.6
VLE, TPx, CO2 pressure 313-393 0.1-70 0.033-0.132 0.003-0.78 101 17.7 Ermatchkov et al.13
heat of solution 313-393 0.06 0.1-1.4 112 6.8 Mathonat65
heat of solution 298 0.017-0.061 0.02-0.25 40 2.1 Carson et al.66
heat capacity (isobaric) 298 0.061-0.185 0-0.64 39 3.0 Weiland et al.67
species concentration 293-313 0.04 0.1-0.7 8 47.5 Jakobsen et al.68

Table 15. Regressed Parameters for the MDEA-H2O-CO2 System with r ) 0.2
parameter component i component j value standard deviation
∞,aq +
∆fG298.15 (J/kmol) MDEAH -2.5951 × 108
2.1986 × 105
∞,aq
∆fH298.15 (J/kmol) MDEAH+ -5.1093 × 108 5.8718 × 105
∞,aq
Cp (J/(kmol K)) MDEAH+ 3.3206 × 105 1.2799 × 104
τij H2O (MDEAH+, HCO-3) 8.7170 0.2246
τij (MDEAH+, HCO-3) H2O -4.2995 0.0836
τij H2O (MDEAH+, CO2-
3 ) 10.4032 0.3676
+
τij 2-
(MDEAH , CO3 ) H2O -4.9252 0.1248
τij MDEA (MDEAH+, HCO-3) 5.2964 0.2746
τij (MDEAH+, HCO-3) MDEA -0.8253 0.0685

to 70 kPa. The correlation results match the experimental data Table 16 shows a comparison of model predictions and
well, except that the calculated CO2 pressure at 313 K is slighter experimental VLE data from numerous other sources not
higher than the measured values for the 8 m MDEA solution at included in the regression. The results highlight the fact that
high CO2 loading (see Figure 16). Upon further examination, we cannot match all the VLE data because the experimental
we find that, under the same conditions, the predicted total
pressure matches the experimental data of Sidi-Boumedine et
al.62 well.

Figure 9. Comparison of the experimental data from Kuranov et al.5


(represented by symbols: (O) T ) 413 K, (4) T ) 393 K, (×) T ) 373 K,
(0) T ) 353 K, and (]) T ) 313 K) for total pressure of the
MDEA-H2O-CO2 system and the model results (represented by lines);
the MDEA concentration is ∼2 m.
Figure 7. Ratio of experimental total pressure to calculated total pressure,
as a function of CO2 loading ((4) data from Kuranov et al.5 and (O) data
from Kamps et al.6).

Figure 10. Comparison of the experimental data from Kuranov et al.5


(represented by symbols: (O) T ) 413 K, (4) T ) 393 K, (×) T ) 373 K,
(0) T ) 353 K, and (]) T ) 313 K) for total pressure of the
Figure 8. Parity plot for the MDEA-H2O-CO2 system total pressure: MDEA-H2O-CO2 system and the model results (represented by lines);
experiment versus model ((4) Kuranov et al.5 and (O) Kamps et al.6). the MDEA concentration is ∼4 m.
Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011 171

Figure 14. Comparison of the experimental data for CO2 partial pressure
Figure 11. Comparison of the experimental data from Kamps et al.6 of the MDEA-H2O-CO2 system and the model results; the MDEA
(represented by symbols: (O) T ) 393 K, (0) T ) 353 K, and (]) T ) 313 concentration is ∼2 m. Empty symbols are data from Ermatchkov et al.13
K) for total pressure of the MDEA-H2O-CO2 system and the model results ((O) T ) 393 K, (0) T ) 353 K, (]) T ) 313 K), solid lines represent the
(represented by lines); the MDEA concentration is ∼8 m. 2009 eNRTL model results, and dashed lines represent the 1986 eNRTL
model results.

Figure 12. Rato of experimental CO2 partial pressure ((O) Ermatchkov et Figure 15. Comparison of the experimental data for CO2 partial pressure
al.13) to calculated CO2 partial pressure (line), as a function of CO2 loading. of the MDEA-H2O-CO2 system and the model results; the MDEA
concentration is ∼4 m. Empty symbols are data from Ermatchkov et al.13
((O) T ) 393 K, (0) T ) 353 K, (]) T ) 313 K), solid lines represent the
2009 eNRTL model results, and dashed lines represent the 1986 eNRTL
model results.

Figure 13. Parity plot for CO2 partial pressure of the MDEA-H2O-CO2
system: experiment ((O) Ermatchkov et al.13) versus model (line).

data from different sources can be inconsistent. With the


exception of the data from Jou et al.,45 Silkenbaeumer et al.,56 Figure 16. Comparison of the experimental data for CO2 partial pressure
of the MDEA-H2O-CO2 system and the model results; the MDEA
and Ali and Aroua,61 the model predictions are very satisfactory, concentration is ∼8 m. Empty symbols are data from Ermatchkov et al.13
with the average relative deviation on pressure (either total ((O) T ) 393 K, (0) T ) 353 K, (]) T ) 313 K), solid lines represent the
pressure or CO2 partial pressure) in the range of 7%-80%. It 2009 eNRTL model results, and dashed lines represent the 1986 eNRTL
is particularly significant that the model predictions give an model results.
excellent match with the recent data of Kamps et al.,60 Sidi-
Boumedine et al.,62 and Ma’mum et al.63
Figure 17 shows the species distribution as a function of CO2 concentrations of the species are consistent with the experimental
loading for a 23 wt % MDEA solution at 293 K. The calculated NMR measurements from Jakobsen et al.68
172 Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011

Table 16. Comparison between Experimental Data and Model Predictions for Total Pressure or CO2 Partial Pressure of the MDEA-H2O-CO2
System
source data points temperature, T (K) pressure, P (kPa) MDEA concentration CO2 loading |∆P/P|a (%)
Jou et al.45 118 298-393 0.001-6000 0.044-0.128 0.0004-1.68 204
Chakma and Meisen46 76 373-473 100-5000 0.03-0.12 0.01-0.95 31.5
Maddox et al.47 99 310-388 20-6000 0.02-0.04 0.17-1.51 21.1
Austgen et al.8 14 313 0.005-100 0.045-0.13 0.003-0.67 32.2
MacGregor and Mather48 5 313 1-4000 0.04 0.12-1.2 22.1
Jou et al.49 37 313-373 0.004-260 0.07 0.002-0.80 37.5
Dawodu and Meisen50 12 373-393 160-4000 0.12 0.09-0.8 13.3
Liu et al.51 16 303-363 20-350 0.09 0.09-0.85 28.5
Mathonat et al.52 9 313-393 2000-10000 0.06 0.5-1.3 57.4
Rho et al.53 103 323-373 0.1-268 0.008-0.31 0.006-0.68 78.7
Baek and Yoon54 12 313 1-2000 0.06 0.12-1.13 53.6
Rogers et al.55 34 313-323 0.00007-1 0.04-0.13 0.0002-0.12 27.1
Silkenbaeumer et al.56 11 313 12-4000 0.07 0.2-1.3 135
Xu et al.57 65 328-363 4-800 0.07-0.13 0.04-0.9 20.2
Lemoine et al.58 13 298 0.02-1.64 0.04 0.02-0.26 11.9
Bishnoi and Rochelle59 3 313 0.1-0.7 0.13 0.01-0.03 17.9
Kamps et al.60 5 313 80-5000 0.03 1.06-1.41 8.9
Ali and Aroua61 15 313-353 0.08-100 0.04 0.05-0.8 495
Sidi-Boumedine et al.62 103 298-348 2.7-4500 0.05-0.11 0.008-1.30 7.7
Ma’mun et al.63 34 328-358 66-813 0.12 0.17-0.81 6.5
Dicko et al.64 5 323 6-434 0.12 0.1-0.9 48.5
a
Experimental pressure expressed either as total pressure or CO2 partial pressure.

Figure 17. Comparison of the experimental data for species concentration Figure 18. Integral CO2 heat of absorption in 30 wt % MDEA aqueous
in MDEA-H2O-CO2 and the model results at T ) 293 K. MDEA solution at 313 K. Symbols (0) represent experimental data from Matho-
concentration is 23 wt %. Symbols represent experimental data from nat;65 lines represent model results ((s) integral heat of absorption,
Jakobsen et al.68 ((O) MDEA, (4) HCO- + 2-
3 , (0) MDEAH , (×) CO3 , (]) (- · · -) differential heat of absorption, ( · · · ) contribution of reactions,
CO2); lines represent model results ((s) MDEA, (- - -) HCO- 3 , (- · -) (- · -) contribution of CO2 dissolution, (- - -) contribution of excess
MDEAH+, (- · · -) CO2- 3 , and ( · · · ) CO2. enthalpies).

Figures 18-20 show comparisons of the model correlations


and the experimental data of Mathonat65 for the integral heat The heat of CO2 dissolution (∆Hdissolution) is calculated as the
of CO2 absorption in aqueous MDEA solution at 313, 353, and enthalpy difference between 1 mol of CO2 in the vapor phase
393 K, respectively. The calculated values are in reasonable and 1 mol of CO2 in aqueous-phase infinite dilution. The
agreement with the experimental data. Also shown in Figures contribution of excess enthalpies (∆Hex) is computed as the
18-20 are the predicted differential heats of CO2 absorption. excess enthalpy difference between the final and initial com-
The integral heat and the differential heat overlap at low CO2 positions of the solution per mole of CO2 absorbed.
loadings, and then diverge much at high CO2 loadings (i.e., The results in Figures 18-20 show that the heat of absorption
>0.8), where the differential heat decreases by >50%. We further is dominated by MDEAH+ dissociation and excess enthalpy.
show the computed integral heat of absorption as the sum of In addition, CO2 dissolution is important near room temperature,
the various contributions from reactions 10-13, CO2 dissolution, whereas CO2 dissociation becomes more important at higher
and excess enthalpy. temperatures.
k Figure 21 shows a comparison of the model correlations and
∆Habs ) ∑ ∆n ∆H° + ∆H
i i dissolution + ∆Hex (25) the experimental data of Weiland et al.67 for heat capacity of
i)1 the MDEA-H2O-CO2 system. The model results are consistent
where ∆Habs is the integral heat of absorption per mole of CO2, with the data.
∆H°i the standard heat of reaction for reaction i per mole of To show the impact of the different versions of the
key component reacted, and ∆ni the reaction extent of the electrolyte NRTL model to the model results, we perform
reaction key component for reaction i when 1 mol CO2 is VLE predictions with the same model parameter values given
absorbed. in Table 15 with the 1986 version of the electrolyte NRTL
Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011 173
concentrations. The difference increases slightly with increas-
ing MDEA concentration.

4. Conclusion
To support process modeling and simulation of the CO2
capture process with MDEA, the electrolyte NRTL model has
been successfully applied to correlate the available experimental
data on thermodynamic properties of the MDEA-H2O-CO2
system. The model has been validated for predictions of vapor-
liquid equilibrium (VLE), heat capacity, and CO2 heat of
absorption of the MDEA-H2O-CO2 system with temperatures
from 313 K to 393 K, MDEA concentrations up to 8 m (∼50
wt %), and CO2 loadings up to 1.32. This model should provide
a comprehensive thermodynamic property representation for the
MDEA-H2O-CO2 system over a broader range of conditions
Figure 19. Integral CO2 heat of absorption in 30 wt % MDEA aqueous
solution at 353 K. Symbols (4) represent experimental data from Matho-
and give more-reliable predictions than those from previous
nat;65 lines represent model results ((s) integral heat of absorption, works.
(- · · -) differential overall absorption heat, ( · · · ) contribution of
reactions, (- · -) contribution of CO2 dissolution, (- - -) contribution
of excess enthalpies). Acknowledgment
The authors thank Huiling Que and Joseph DeVincentis for
their support in preparing the manuscript.

Literature Cited
(1) Zhang, Y.; Chen, H.; Chen, C.-C.; Plaza, J. M.; Dugas, R.; Rochelle,
G. T. Rate-Based Process Modeling Study of CO2 Capture with Aqueous
Monoethanolamine Solution. Ind. Eng. Chem. Res. 2009, 48, 9233–9246.
(2) Kohl. A. L.; Riesenfeld, F. C. Gas Purification, 4th ed.; Gulf
Publishing: Houston, TX, 1985.
(3) Taylor, R.; Krishna, R.; Kooijman, H. Real-World Modeling of
Distillation. Chem. Eng. Prog. 2003, 99, 28–39.
(4) Chen, C.-C.; Mathias, P. M. Applied Thermodynamics for Process
Modeling. AIChE J. 2002, 48, 194–200.
(5) Kuranov, G.; Rumpf, B.; Smirnova, N. A.; Maurer, G. Solubility of
Single Gases Carbon Dioxide and Hydrogen Sulfide in Aqueous Solutions
of N-Methyldiethanolamine in the Temperature Range 313-413 K at
Figure 20. Integral CO2 heat of absorption in 30 wt % MDEA aqueous Pressures up to 5 MPa. Ind. Eng. Chem. Res. 1996, 35, 1959–1966.
solution at 393 K. Symbols (O) represent experimental data from Matho- (6) Kamps, Á. P.-S.; Balaban, A.; Jödecke, M.; Kuranov, G.; Smirnova,
nat;65 lines represent model results ((s) integral heat of absorption, N. A.; Maurer, G. Solubility of Single Gases Carbon Dioxide and Hydrogen
(- · · -) differential heat of absorption, ( · · · ) contribution of reactions, Sulfide in Aqueous Solutions of N-Methyldiethanolamine at Temperatures
(- · -) contribution of CO2 dissolution, (- - -) contribution of excess from 313 to 393 K and Pressures up to 7.6 MPa: New Experimental Data
enthalpies). and Model Extension. Ind. Eng. Chem. Res. 2001, 40, 696–706.
(7) Chen, C.-C. Toward Development of Activity Coefficient Models
for Process and Product Design of Complex Chemical Systems. Fluid Phase
Equilib. 2006, 241, 103–112.
(8) Austgen, D. M.; Rochelle, G. T.; Chen, C.-C. 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. 1991, 30, 543–555.
(9) Posey, M. L. Thermodynamic Model for Acid Gas Loaded Aqueous
Alkanolamine Solutions, Ph.D. Thesis, University of Texas at Austin, Austin,
TX, 1996.
(10) Song, Y.; Chen, C.-C. Symmetric Electrolyte Nonrandom Two-
Liquid Activity Coefficient Model. Ind. Eng. Chem. Res. 2009, 48, 7788–
7797.
(11) Chen, C.-C.; Evans, L. B. A Local Composition Model for the
Excess Gibbs Energy of Aqueous Electrolyte Systems. AIChE J. 1986, 32,
444–454.
(12) Chen, C.-C.; Britt, H. I.; Boston, J. F.; Evans, L. B. Local
Figure 21. Comparison of the experimental data for heat capacity of the Composition Model for Excess Gibbs Energy of Electrolyte Systems. Part
MDEA-H2O-CO2 system and the model results at T ) 298 K. Symbols I: Single Solvent, Single Completely Dissociated Electrolyte Systems. AIChE
represent experimental data from Weiland et al.67 ((O) 60 wt % MDEA, J. 1982, 28, 588–596.
(4) 50 wt % MDEA, (0) 40 wt % MDEA, and (]) 30 wt % MDEA); lines (13) Ermatchkov, V.; Kamps, Á. P.-S.; Maurer, G. Solubility of Carbon
represent model results. Dioxide in Aqueous Solutions of N-Methyldiethanolamine in the Low Gas
Loading Region. Ind. Eng. Chem. Res. 2006, 45, 6081–6091.
model.11 Figures 14-16 show that the two versions of the (14) Pitzer, K. S. Thermodynamics of Electrolytes. I. Theoretical Basis
model yield practically identical results at low MDEA and General Equations. J. Phys. Chem. 1973, 77, 268–277.
174 Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011

(15) Arcis, H.; Rodier, L.; Karine, B.-B.; Coxam, J.-Y. Modeling of (39) DiGuilio, R. M.; Lee, R. J.; Schaeffer, S. T.; Brasher, L. L.; Teja,
(Vapor + Liquid) Equilibrium and Enthalpy of Solution of Carbon Dioxide A. S. Densities and Viscosities of the Ethanolamines. J. Chem. Eng. Data
(CO2) in Aqueous Methyldiethanolamine (MDEA) Solutions. J. Chem. 1992, 37, 239–242.
Thermodyn. 2009, 41, 783–789. (40) Xu, S.; Qing, S.; Zhen, Z.; Zhang, C.; Carroll, J. Vapor Pressure
(16) Faramarzi, L.; Kontogeorgis, G. M.; Thomsen, K.; Stenby, E. H. Measurements of Aqueous N-Methyldiethanolamine Solutions. Fluid Phase
Extended UNIQUAC Model for Thermodynamic Modeling of CO2 Absorp- Equilib. 1991, 67, 197–201.
tion in Aqueous Alkanolamine Solutions. Fluid Phase Equilib. 2009, 282, (41) Voutsas, E.; Vrachnos, A.; Magoulas, K. Measurement and
121–132. Thermodynamic Modeling of the Phase Equilibrium of Aqueous
(17) Thomsen, K.; Rasmussen, P. Modeling of Vapor-Liquid-Solid N-Methyldiethanolamine Solutions. Fluid Phase Equilib. 2004, 224, 193–
Equilibrium in Gas-Aqueous Electrolyte Systems. Chem. Eng. Sci. 1999, 197.
54, 1787–1802. (42) Kim, I.; Svendsen, H. F.; Børresen, E. Ebulliometric Determination
(18) Böttinger, W.; Maiwald, M.; Hasse, H. Online NMR Spectroscopic of Vapor-Liquid Equilibria for Pure Water, Monoethanolamine, N-
Study of Species Distribution in MEA-H2O-CO2 and DEA-H2O-CO2. Methyldiethanolamine, 3-(Methylamino)-propylamine, and Their Binary and
Fluid Phase Equilib. 2008, 263, 131–143. Ternary Solutions. J. Chem. Eng. Data 2008, 53, 2521–2531.
(19) Hilliard, M. A Predictive Thermodynamic Model for an Aqueous (43) Maham, Y.; Mather, A. E.; Mathonat, C. Excess properties of
Blend of Potassium Carbonate, Piperazine, and Monoethanolamine for (alkyldiethanolamine + H2O) mixtures at temperatures from (298.15 to
Carbon Dioxide, Ph.D. Dissertation, University of Texas at Austin, Austin, 338.15) K. J. Chem. Thermodyn. 2000, 32, 229–236.
TX, 2008. (44) Chiu, L.-F.; Li, M.-H. Heat Capacity of Alkanolamine Aqueous
(20) Gross, J.; Sadowski, G. Perturbed-Chain SAFT: An Equation of Solutions. J. Chem. Eng. Data 1999, 44, 1396–1401.
State Based on a Perturbation Theory for Chain Molecules. Ind. Eng. Chem. (45) Jou, F. Y.; Mather, A. E.; Otto, F. D. Solubility of H2S and CO2 in
Res. 2001, 40, 1244–1260. Aqueous Methyldiethanolamine Solution. Ind. Eng. Chem. Process Des.
(21) Gross, J.; Sadowski, G. Application of the Perturbed-Chain SAFT DeV. 1982, 21, 539–544.
Equation of State to Associating Systems. Ind. Eng. Chem. Res. 2002, 41, (46) Chakma, A.; Meisen, A. Solubility of CO2 in Aqueous Methyldi-
5510–5515. ethanolamine and N,N-Bis(hydroxyethl)piperazine Solutions. Ind. Eng.
(22) Aspen Physical Property System, V7.2; Aspen Technology, Inc.: Chem. Res. 1987, 26, 2461–2466.
Burlington, MA, 2010. (47) Maddox, R. N.; Bhairi, A. H.; Diers, J. R.; Thomas, P. A.
(23) Van Ness, H. C.; Abbott, M. M. Vapor-Liquid Equilibrium: Part Equilibrium Solubility of Carbon Dioxide or Hydrogen Sulfide in Aqueous
VI. Standard State Fugacities for Supercritical Components. AIChE J. 1979, Solutions of Monoethanolamine, Diglycolamine, Diethanolamine and Me-
25, 645–653. thyldiethanolamine. GPA Res. Rep. 1987, 1–47.
(24) Brelvi, S. W.; O’Connell, J. P. Corresponding States Correlations (48) MacGregor, R. J.; Mather, A. E. Equilibrium Solubility of H2S and
for Liquid Compressibility and Partial Molar Volumes of Gases at Infinite CO2 and Their Mixtures in a Mixed Solvent. Can. J. Chem. Eng. 1991, 69,
Dilution in Liquids. AIChE J. 1972, 18, 1239–1243. 1357–1366.
(25) Yan, Y.-Z.; Chen, C.-C. Thermodynamic Modeling of CO2 Solubil- (49) Jou, F. Y.; Carroll, J. J.; Mather, A. E.; Otto, F. D. The Solubility
ity in Aqueous Solutions of NaCl and Na2SO4. Submitted to J. Supercrit. of Carbon Dioxide and Hydrogen Sulfide in a 35 wt % Aqueous Solution
Fluids. Methyldiethanolamine. Can. J. Chem. Eng. 1993, 71, 264–268.
(26) Clarke, J. K. A. Kinetics of Absorption of Carbon Dioxide in (50) Dawodu, O. F.; Meisen, A. Solubility of Carbon Dioxide in Aqueous
Monoethanolamine Solutions at Short Contact Times. Ind. Eng. Chem. Mixtures of Alkanolamines. J. Chem. Eng. Data 1994, 39, 548–552.
Fundam. 1964, 3, 239–245. (51) Liu, H.; Xu, G.; Zhang, C.; Wu, Y. Solubilities of Carbon Dioxide
(27) Wang, Y. W.; Xu, S.; Otto, F. D.; Mather, A. E. Solubility of N2O in Aqueous Activated Methyldiethanolamine Solutions. Huadong Ligong
in Alkanolamines and in Mixed Solvents. Chem. Eng. J. 1992, 48, 31–40. Daxue Xuebao 1999, 25, 242–246.
(28) Versteeg, G. F.; Van Swaaij, W. P. M. Solubility and Diffusivity (52) Mathonat, C.; Majer, V.; Mather, A. E.; Grolier, J.-P. E. Enthalpies
of Acid Gases (CO2, N2O) in Aqueous Alkanolamine Solutions. J. Chem. of Absorption and Solubility of CO2 in Aqueous Solutions of Methyldi-
Eng. Data 1988, 33, 29–34. ethanolamine. Fluid Phase Equilib. 1997, 140, 171–182.
(29) Wagman, D. D.; Evans, W. H.; Parker, V. B.; Schumm, R. H.; (53) Rho, S.-W.; Yoo, K.-P.; Lee, J. S.; Nam, S. C.; Son, J. E.; Min,
Halow, I.; Bailey, S. M.; Churney, K. L.; Nuttall, R. L. The NBS tables of B.-M. Solubility of CO2 in Aqueous Methyldiethanolamine Solutions.
chemical thermodynamic properties. Selected values for inorganic and C1 J. Chem. Eng. Data 1997, 42, 1161–1164.
and C2 organic substances in SI units. J. Phys. Chem. Ref. Data 1982, 11 (54) Baek, J.-I.; Yoon, J.-H. Solubility of Carbon Dioxide in Aqueous
(Supplement No. 2), pp 2-38 and 2-83. Solutions of 2-Amino-2-Methyl-1,3-Propanediol. J. Chem. Eng. Data 1998,
(30) Kamps, Á. P.-S.; Maurer, G. Dissociation Constant of N-Meth- 43, 635–637.
yldiethanolamine in Aqueous Solution at Temperatures from 278 to 368 (55) Rogers, W. J.; Bullin, J. A.; Davison, R. R. FTIR Measurements
K. J. Chem. Eng. Data 1996, 41, 1505–1513. of Acid-Gas-Methyldiethanolamine Systems. AIChE J. 1998, 44, 2423–
(31) Criss, C. M.; Cobble, J. W. The Thermodynamic Properties of High 2430.
Temperature Aqueous Solutions. V. The Calculation of Ionic Heat Capacities (56) Silkenbäumer, D.; Rumpf, B.; Lichtenthaler, R. N. Solubility of
up to 200°. Entropies and Heat Capacities above 200°. J. Am. Chem. Soc. Carbon Dioxide in Aqueous Solutions of 2-Amino-2-methyl-1-propanol and
1964, 86, 5390–5393. N-Methyldiethanolamine and Their Mixtures in the Temperature Range from
(32) Von Niederhausern, D. M.; Wilson, G. M.; Giles, N. F. Critical 313 to 353 K and Pressures up to 2.7 MPa. Ind. Eng. Chem. Res. 1998, 37,
Point and Vapor Pressure Measurements for 17 Compounds by a Low 3133–3141.
Residence Time Flow Method. J. Chem. Eng. Data 2006, 51, 1990–1995. (57) Xu, G.-W.; Zhang, C.-F.; Qin, S.-J.; Gao, W.-H.; Liu, H.-B. Gas-
(33) Daubert, T. E.; Hutchison, G. Vapor Pressure of 18 Pure Industrial Liquid Equilibrium in a CO2-MDEA-H2O System and the Effect on
Chemicals. AIChE Symp. Ser. 1990, 86, 93–114. Piperazine on It. Ind. Eng. Chem. Res. 1998, 37, 1473–1477.
(34) Noll, O.; Valtz, A.; Richon, D.; Getachew-Sawaya, T.; Mokbel, I.; (58) Lemoine, B.; Li, Y.-G.; Cadours, R.; Bouallou, C.; Richon, D.
Jose, J. Vapor Pressures and Liquid Densities of N-Methylethanolamine, Partial Vapor Pressure of CO2 and H2S over Aqueous Methyldiethanolamine
Diethanolamine, and N-Methyldiethanolamine. ELDATA: Int. Electron. J. Solutions. Fluid Phase Equilib. 2000, 172, 261–277.
Phys.-Chem. Data 1998, 4, 105–120. (59) Bishnoi, S.; Rochelle, G. T. Thermodynamics of Piperazine-
(35) Maham, Y.; Mather, A. E.; Hepler, L. G. Excess Molar Enthalpies Methyldiethanolamine-Water-Carbon Dioxide. Ind. Eng. Chem. Res. 2002,
of (Water + Alkanolamine) Systems and Some Thermodynamic Calcula- 41, 604–612.
tions. J. Chem. Eng. Data 1997, 42, 988–992. (60) Kamps, Á. P.-S.; Rumpf, B.; Maurer, G.; Anoufrikov, Y.; Kuranov,
(36) Chen, Y.-J.; Shih, T.-W.; Li, M.-H. Heat Capacity of Aqueous G.; Smirnova., N. A. Solubility of CO2 in H2O + N-Methyldiethanolamine
Mixtures of Monoethanolamine with N-Methyldiethanolamine. J. Chem. + (H2SO4 or Na2SO4). AIChE J. 2002, 48, 168–177.
Eng. Data 2001, 46, 51–55. (61) Ali, B. S.; Aroua, M. K. Effect of Piperazine on CO2 loading in
(37) Zhang, K.; Hawrylak, B.; Palepu, R.; Tremaine, P. R. Thermody- Aqueous Solutions of MDEA at Low Pressure. Int. J. Thermophys. 2004,
namics of Aqueous Amines: Excess Molar Heat Capacities, Volumes, and 25, 1863–1870.
Expansibilities of (Water + Methyldiethanolamine (MDEA)) and (Water (62) Sidi-Boumedine, R.; Horstmann, S.; Fischer, K.; Provost, E.; Fürst,
+ 2-Amino-2-methyl-1-propanol (AMP)). J. Chem. Thermodyn. 2002, 34, W.; Gmehling, J. Experimental Determination of Carbon Dioxide Solubility
679–710. Data in Aqueous Alkanolamine Solutions. Fluid Phase Equilib. 2004, 218,
(38) AI-Ghawas, H. A.; Hagewiesche, D. P.; Ruiz-Ibanez, G.; Sandall, 85–94.
O. C. Physicochemical Properties Important for Carbon Dioxide Absorption (63) Ma’mun, S.; Nilsen, R.; Svendsen, H. F.; Juliussen, O. Solubility of
in Aqueous Methyldiethanolamine. J. Chem. Eng. Data 1989, 34, 385– Carbon Dioxide in 30 mass% Monoethanolamine and 50 mass% Methyldi-
391. ethanolamine Solutions. J. Chem. Eng. Data 2005, 50, 630–634.
Ind. Eng. Chem. Res., Vol. 50, No. 1, 2011 175
(64) Dicko, M.; Coquelet, C.; Jarne, C.; Northrop, S.; Richon, D. Acid (67) Weiland, R. H.; Dingman, J. C.; Cronin, D. B. Heat capacity of
Gases Partial Pressures above a 50 wt% Aqueous Methyldiethanolamine Aqueous Monoethanolamine, Diethanolamine, N-Methyldiethanolamine, and
Solution: Experimental Work and Modeling. Fluid Phase Equilib. 2010, N-Methyldiethanolamine-Based Blends with Carbon Dioxide. J. Chem. Eng.
289, 99–109. Data 1997, 42, 1004–1006.
(65) Mathonat, C. Calorimetrie de mélange, a ecoulement, a temperatures (68) Jakobsen, J. P.; Krane, J.; Svendsen, H. F. Liquid-Phase Composi-
et pressions elevees. Application a l’etude de l’elimination du dioxide de tion Determination in CO2-H2O-Alkanolamine Systems: An NMR Study.
carbone a l’aide de solutions aqueuses d’alcanolamines, Universite Blaise Ind. Eng. Chem. Res. 2005, 44, 9894–9903.
Pascal, Paris, 1995, p 265.
(66) Carson, J. K.; Marsh, K. N.; Mather, A. E. Enthalpy of Solution of ReceiVed for reView March 19, 2010
Carbon Dioxide in (Water + Monoethanolamine, or Diethanolamine, or ReVised manuscript receiVed June 25, 2010
N-Methyldiethanolamine) and (Water + Monoethanolamine + N-Meth- Accepted July 12, 2010
yldiethanolamine) at T ) 298.15 K. J. Chem. Thermodyn. 2000, 32, 1285–
1296. IE1006855

You might also like