You are on page 1of 24

International Journal of Greenhouse Gas Control 53 (2016) 401–424

Contents lists available at ScienceDirect

International Journal of Greenhouse Gas Control


journal homepage: www.elsevier.com/locate/ijggc

Thermodynamic modeling of liquid–liquid phase change solvents for


CO2 capture
Muhammad Waseem Arshad, Nicolas von Solms, Kaj Thomsen ∗
Technical University of Denmark, Department of Chemical and Biochemical Engineering, Center for Energy Resources Engineering (CERE), Søltofts Plads
Building 229, DK-2800 Kongens Lyngby, Denmark

a r t i c l e i n f o a b s t r a c t

Article history: A thermodynamic model based on Extended UNIQUAC framework has been developed in this work
Received 20 January 2016 for the de-mixing liquid–liquid phase change solvents, DEEA (2-(diethylamino)ethanol) and MAPA
Received in revised form 6 August 2016 (3-(methylamino)propylamine). Parameter estimation was performed for two ternary systems, H2 O-
Accepted 15 August 2016
DEEA-CO2 and H2 O-MAPA-CO2 , and a quaternary system, H2 O-DEEA-MAPA-CO2 (phase change system),
Available online 31 August 2016
by using different types of experimental data (equilibrium and thermal) consisting of pure amine vapor
pressure, vapor-liquid equilibrium, solid-liquid equilibrium, liquid–liquid equilibrium, excess enthalpy,
Keywords:
and heat of absorption of CO2 in aqueous amine solutions. 94 model parameters and 6 thermodynamic
Thermodynamic modeling
Extended UNIQUAC
properties were fitted to approximately 1500 experimental data. The developed model accurately rep-
Phase change solvents resents the equilibrium and thermal data for the studied systems with a single unique set of parameters.
DEEA (2-(diethylamino)ethanol) The model parameters are valid in the temperature range from −25 to 200 ◦ C, CO2 partial pressure from 0
MAPA (3-(methylamino)propylamine) to 945 kPa, and concentration of DEEA, MAPA, and CO2 up to 131, 23 and 33 mol(kg H2 O)−1 respectively.
CO2 capture The model calculated speciation are also presented for the studied systems. The model developed in this
Liquid–liquid equilibrium work can be used for process simulation of CO2 capture with aqueous blends of DEEA/MAPA.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction per tonne CO2 after optimization (Knudsen et al., 2009). Some novel
solvents investigated in European projects CASTOR and CESAR have
Post-combustion absorption-desorption process for CO2 cap- shown good results. These solvents can capture 90% CO2 from flue
ture with amine based solvents is the most mature and industrially gas and reduce the energy penalty down to 3.1 GJ per tonne CO2
available technology today (Tontiwachwuthikul and Idem, 2013). (Knudsen et al., 2011). Recently, some innovative solvent systems
At the moment, this is the only technology that has been demon- have shown great potential. The thermomorphic biphasic solvent
strated commercially for CO2 capture, for example, (1) Technology (TBS) systems have shown a solvent regeneration duty of 2.5 GJ per
Center Mongstad (TCM) in Norway, the largest demonstration test tonne CO2 (Zhang et al., 2011). The DMXTM (solvent) process by
facility for CO2 capture technologies to date with a designed capac- IFPEN has claimed to reduce the energy demand down to 2.1 GJ
ity of capturing 100,000 tonne CO2 per year (Gassnona Annual per tonne CO2 (Raynal et al., 2011a,b). Liebenthal et al. (2013) has
Report, 2012), and (2) SaskPower in Canada, the world’s first full studied a phase change solvent system of aqueous DEEA/MAPA for
scale coal-fired post-combustion CCS facility at Boundary Dam CO2 capture. This solvent system has shown a reboiler duty of 2.4 GJ
power station (Unit #3) capable of capturing one million tonnes per tonne CO2 at a desorber pressure of 400 kPa. The heat duty
of CO2 per year (SaskPower Annual Report, 2014). Nevertheless, can further be reduced to 2.2 GJ per tonne CO2 at lower reboiler
high energy demand of the process is still the main challenge of pressures. Most of these emerging solvent systems are in their early
this technology (Liang et al., 2015a). stage of development (lab or pilot scale) and have yet to prove their
Conventional single solvents such as monoethanolamine (MEA), commercial viability.
30% MEA, uses a large amount of energy to capture CO2 , approx. There are immense efforts in quest of optimal amine based
4.1 GJ per tonne CO2 , but the energy penalty can be reduced to 3.7 GJ solvent systems for CO2 capture (Liang et al., 2015b). The “game-
changer” (solvent system) must have fast reaction kinetics, high
CO2 loading capacity, and very importantly low solvent regener-
∗ Corresponding author.
ation energy. In addition, it should also have high cyclic capacity,
E-mail addresses: mwa114@gmail.com (M.W. Arshad), nvs@kt.dtu.dk
low chemical and thermal degradation, low corrosiveness, and be
(N. von Solms), kth@kt.dtu.dk (K. Thomsen). environmentally benign. Keeping these characteristics in focus, two

http://dx.doi.org/10.1016/j.ijggc.2016.08.014
1750-5836/© 2016 Elsevier Ltd. All rights reserved.
402 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

amine based solvents, DEEA (2-(diethylamino)ethanol) and MAPA The Gamma-Phi ( − ) approach is used in this work for
(3-(methylamino)propylamine), are of interest in this study. Aque- the phase equilibrium calculations. This means that the Extended
ous blends of DEEA/MAPA (at certain concentrations) give biphasic UNIQUAC thermodynamic model is used to calculate the aqueous-
liquid–liquid phase split (phase change solvents) when reacted phase activity coefficient () and Soave-Redlich-Kwong (SRK) cubic
with CO2 . The details of these solvent systems can be seen in our equation of state (EoS) is used to determine the vapor-phase fugac-
previous work (Arshad et al., 2013a,b,c; Arshad et al., 2014a,b). The ity coefficient (). The model only requires UNIQUAC volume (r)
main concept in these phase change solvent systems is that the and surface area (q) parameters of each species for calculations of
upper light liquid phase, lean in CO2 and rich in DEEA, is sent back the combinatorial (or entropic) term and binary interaction energy
to the absorber without regeneration and the lower heavy liquid parameters for each pair of species for calculations of the resid-
phase, rich in CO2 and MAPA, is sent to the stripper for regeneration. ual (or enthalpic) term. The Debye-Hückel term has no adjustable
The recirculation of CO2 lean liquid phase to the absorber without parameters and represents only the electrostatic interactions of the
regeneration (together with the regenerated CO2 rich phase) can ionic species in the aqueous phase. Similarly, the SRK equation has
maintain a large driving force for CO2 absorption in the absorber. no interaction parameters to be fitted for the gas phase. The SRK
On the other hand, the heavy liquid phase rich in CO2 loading with EoS is applied with classical mixing rules (quadratic mixing rule for
a lower circulation rate to the desorber can achieve a relatively the a parameter and linear mixing rule for the b parameter).
lower sensible heat and heat of vaporization during regeneration.
This high driving force for CO2 absorption in the absorber and low 2. Chemical potential and activity coefficient
regeneration energy in the stripper can result in high efficiency and
low cost of the absorption-desorption process (Arshad, 2014). The criterion for the vapor-liquid equilibrium (VLE), solid-liquid
Thermodynamic models are essential for the process sim- equilibrium (SLE) and liquid–liquid equilibrium (LLE) is that the
ulation and optimization of the CO2 capture processes. When chemical potential of each component is same in all the phases.
CO2 (acid gas) is dissolved in the aqueous amine solutions, it The chemical potential of component i (i ) is the summation of
forms electrolyte solutions. Two activity coefficient based mod- standard state chemical potential (0i ) and a term that describes
els, e-NRTL (electrolyte Non-Random Two-Liquid) and Extended the concentration and activity coefficient of that component i as
UNIQUAC, are the most commonly used local composition mod- given by
els applied for thermodynamic modeling of the electrolyte systems
containing H2 O-Amine/Alkanolamine-Acid gas. The Extended UNI- i = 0i + RT ln ai = 0i + RT ln (xi i ) (1)
QUAC thermodynamic framework has been adopted in this work.
where, ai is the activity, xi is the mole fraction, and i is the activity
Previously, the Extended UNIQUAC model has been applied suc-
coefficient of component i. R is the gas constant (8.314 J mol−1 K−1 )
cessfully to a variety of aqueous amine and carbonate systems.
and T is the temperature in Kelvin. The activity coefficient describes
Thomsen and Rasmussen (1999) implemented the Extended UNI-
the non-ideal behavior of the component i and determined by
QUAC model to describe the vapor-liquid equilibria (VLE) and
the Extended UNIQUAC model. In this work, the symmetrical con-
solid-liquid equilibria (SLE) in the H2 O-NH3 -CO2 system. Darde
vention is adopted for water (solvent) and the unsymmetrical
et al. (2010) further upgraded the model to a wide range of con-
convention is adopted for all remaining components (solutes).
centration, temperature and pressure, and later Darde et al. (2012)
The chemical potential of water (w ), according to the symmet-
used the upgraded model in Aspen Plus for process simulation of
rical convention, can be written as
CO2 capture with aqueous ammonia. Gaspar et al. (2014) also used
this upgraded model in Aspen Plus to simulate a novel aqueous w = 0w + RT ln aw = 0w + RT ln (xw w ) (2)
ammonia process for CO2 capture. Fosbøl et al. (2009) developed
the model to determine the VLE, SLE, and excess enthalpy data where, 0w is the standard state chemical potential of water, aw
in a mixed solvent CO2 - Na2 CO3 -NaHCO3 - MEG(monoethylene is the activity of water, xw is the water mole fraction, and w is
glycol)-H2 O electrolyte system. Faramarzi et al. (2009) imple- the rational symmetrical activity coefficient of water. The activity
mented the Extended UNIQUAC framework to model the VLE coefficient of water in its pure component state is unity at all tem-
in H2 O-MEA-CO2 , H2 O-MDEA(methyldiethanolamine)-CO2 , and peratures, according to the symmetrical convention i.e. w = 1 for
H2 O-MEA-MDEA-CO2 systems and Sadegh et al. (2015a) further xw = 1.
improved the model by including more experimental data in the According to unsymmetrical convention, the activity coefficient
parameter estimation. Sadegh et al. (2015b) and Sadegh (2013) also of component i is defined as
implemented the model to H2 O-MDEA-H2 S and H2 O-MDEA-H2 S- i
i∗ = (3)
CH4 systems for acid gas removal from sour natural gas streams. i∞
Only two thermodynamic modeling studies were found in the
literature for aqueous DEEA and MAPA systems. Hartono et al. where, i∞ is the activity coefficient at infinite dilution and i∗ is
(2013) implemented the UNIQUAC framework to model the binary the rational unsymmetrical activity coefficient of solute i and has
and ternary VLE in H2 O-DEEA-MAPA system (unloaded systems). a value of unity at infinite dilution i.e. i∗ = 1 at infinite dilution.
Monteiro et al. (2013) reported the VLE and modeling results for The term “unsymmetrical” is used due to the fact that this activity
the H2 O-DEEA-CO2 system by using the e-NRTL model. No model- coefficient has a value of 1 at infinite dilution rather than in the
ing study was found in the literature for the CO2 loaded aqueous pure component state. The infinite dilution activity coefficient of
MAPA and aqueous DEEA/MAPA blends (phase change). A first com- a component in water is dependent on temperature and pressure.
plete modeling work is presented here for the H2 O-DEEA-CO2 , The rational unsymmetrical activity coefficient can be used in Eq.
H2 O-MAPA-CO2 , and H2 O-DEEA-MAPA-CO2 (phase change) sys- (1) to derive an expression for the chemical potential of component
tems using the Extended UNIQUAC thermodynamic model. The (ion) i based on the unsymmetrical convention.
type of experimental data used in this work are; pure amine vapor  
i = ∗i + RT ln xi i∗ (4)
pressure, VLE (both unloaded and CO2 loaded), SLE (freezing point
depression data), LLE, excess enthalpy, and differential heat of where, ∗i is the rational unsymmetrical standard state chemical
absorption of CO2 in aqueous amines. The model developed in this potential defined as
work can accurately describe the equilibrium and thermal data for  
∗i = 0i + RT ln i∞ (5)
the studied systems with a single unique set of parameters.
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 403

3. Equilibrium calculations convention for water (Eq. (2)) and unsymmetrical convention for
remaining species (Eq. (4)) to get following expression
3.1. Speciation equilibria    
∗ + ∗ + − ∗CO + 0H  a∗ a∗

HCO− (aq) H (aq) 2 (aq) 2 O(l) HCO− (aq) H+ (aq)
3
When CO2 in dissolved in aqueous amine solutions, several reac- − = ln 3
(15)
RT a∗CO aH2 O(l)
2 (aq)
tions take place depending on the type of amine. The following
speciation equilibria were considered for the thermodynamic mod- An expression similar to Eq. (15) can be obtained for all the
eling. equilibria given in Eqs. (6)–(12) and solved simultaneously for spe-
Dissociation of Water: ciation calculations. Eq. (15) can be written in a general form as

H2 O (l)  H+ (aq) + OH− (aq) (6) Gj0 


− = vi ln ai (16)
RT
Hydration of CO2 : i

CO2 (aq) +H2 O (l)  HCO− +


3 (aq) + H (aq) (7) where, Gj0 is the increment in standard state Gibbs energy for the
reaction j. ai is the activity of species i and vi is the stoichiometric
Bicarbonate dissociation:
coefficient of species i, negative for the reactants and positive for
HCO− 2− +
3 (aq)  CO3 (aq) + H (aq) (8) the products.

Protonation of DEEA: 3.2. Vapor-liquid equilibria


DEEA (aq) + H+ (aq)  DEEAH+ (aq) (9)
The following vapor-liquid equilibria were considered
Protonation of MAPA:
H2 O (l)  H2 O (g) (17)
MAPA (aq) + H+ (aq)  MAPAH+ (aq) (10)
CO2 (aq)  CO2 (g) (18)
Di-protonation of MAPA: DEEA (aq)  DEEA (g) (19)
+ + + +
MAPAH (aq) + H (aq)  HMAPAH (aq) (11) MAPA (aq)  MAPA (g) (20)
MAPA carbamate (prim): The equilibrium condition is that the chemical potential of com-
ponent i in the aqueous phase is identical to the chemical potential
MAPA (aq) + HCO− −
3 (aq)  MAPACOO (aq) + H2 O (l) (12)
of component i in the gas phase at a given temperature and pres-
When CO2 is dissolved in the aqueous tertiary amine solutions sure.
like DEEA, it promotes the hydration of CO2 to produce bicarbonate
i(aq) = i(g) (21)
as suggested by Donaldson and Nguyen (1980).
The chemical potential of component i in the gas phase at a
DEEA (aq) +H2 O (l) +CO2 (aq)  DEEAH+ (aq) +HCO−
3 (aq) (13)
temperature T and pressure P can be expressed as
It should be noted that the reaction given in Eq. (13) is used in  
0,ig
ˆi
yi P 
this work as two separate reactions given in Eqs. (7) and (9). i(g) (T, P) = i (T, P0 ) + RT ln (22)
P0
On the other hand MAPA is a diamine containing a primary
and a secondary amine functional group. It forms carbamate when 0,ig
where, i (T, P0 ) is the standard state chemical potential of pure
reacted with CO2 and there are possibilities of a large number
gas i at temperature T and pressure P0 = 1 bar, yi is the mole fraction
of chemical reactions and formation of several different species. ˆ i is the fugacity coefficient
of component i in the gas phase, and 
The complexity of the system is similar to that of aqueous 2-
of component i which describes the deviation from the ideal gas
((2-Aminoethyl)amino)ethanol (AEEA), an alkanolamine with both
behavior just like the activity coefficient (i ) which describes the
primary and secondary amine groups (Ma’mun et al., 2006). There
non-ideality in the solutions as described earlier.
may be eight possible MAPA species present in the aqueous
The chemical potential of component i in the aqueous phase can
phase: MAPA (aq), MAPAH+ , + HMAPAH+ , MAPACOO− (both pri-
be written as
mary and secondary carbamate), + HMAPACOO− (both primary and  
secondary protonated carbamate), and − OOCMAPACOO− . Ma’mun i(aq) (T, P) = ∗i(aq) + Vi(aq) (P − P0 ) + RT ln xi(aq) i(aq)

(23)
et al. (2006) reported that the formation of secondary carbamate
and dicarbamate of AEEA is very small in the whole CO2 loading Eqs. (22) and (23) can be equated according to the criterion for
range of 0–1 mol CO2 (mol AEEA)−1 . They also reported that the vapor-liquid equilibrium as given in Eq. (21) to get an expression
primary carbamate of AEEA is the dominating carbamate species given below
present in the aqueous phase. Therefore, only four species of MAPA 0,ig  
∗i(aq) − i (T, P0 ) Vi(aq) (P − P0 ) ˆ iP
(MAPA (aq), MAPAH+ , + HMAPAH+ , and MAPACOO− (prim)) were + = ln
yi 
(24)
considered for modeling in this work as shown in Eqs. (10)–(12). RT RT xi i∗ P0
For all the speciation equilibria given in Eqs. (6)–(12), the con-
The second term on left hand side of Eq. (24) is the pressure
dition for equilibrium is that the summation of chemical potential
correction to the aqueous phase chemical potential. This term can
of the reactants is equal to the summation of chemical potential of
be ignored at low pressures.
the products.
The expression for the calculation of vapor pressure of pure
 
i, Reactants = i, Products (14) components (only amines in this work, DEEA and MAPA) is given
as
i i  
G0 ˆ iP
yi 
For equilibrium calculations Eq. (14) is applied to, for exam- − = ln (25)
RT xi i∗ P0
ple, hydration of CO2 (Eq. (7)) by considering the symmetrical
404 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

The mole fraction of pure component i is same both in the liq- for the water-DEEA system can be calculated by using the Gibbs-
uid and the gas phases i.e. xi = yi . Also, the rational symmetrical Helmholtz equation.
activity coefficient of pure component i is unity i.e. i = 1 at xi = 1.  

Therefore by using the definition of rational unsymmetrical activity ∂ GE /RT    


HE ∂ ln w ∂ ln DEEA
=− = xw + xDEEA (31)
coefficient (Eq. (3)) for pure component, Eq. (25) can be reduced to ∂T RT 2 ∂T P,x ∂T P,x

P,x

G0 ˆ i ∞P

− = ln i
(26) where, DEEA is the symmetrical activity coefficient of DEEA which
RT P0 can be obtained from the unsymmetrical activity coefficient of
DEEA by using Eq. (3).
which is used in the calculation of vapor pressure of pure amines
(DEEA and MAPA). 4.2. Heat of absorption of CO2

3.3. Solid-liquid equilibrium The heat involved when CO2 is absorbed in the aqueous amine
solutions can be calculated from the energy balance of the absorp-
The solid-liquid equilibrium considered in this work is tion process.

n2 H2 − n1 H1 − nCO2 HCO2
H2 O (s)  H2 O (l) (27) Habs = (32)
nCO2
The criterion for the solid-liquid equilibrium is that the chemical
potential is identical in the two phases. The solid compound is in where, Habs is the heat of absorption per mole of CO2 absorbed. H1
its standard state. Therefore, the chemical potential of the solid and n1 are the molar enthalpy of formation and number of moles of
compound is equal to its standard state chemical potential. Since solution before CO2 absorption, H2 and n2 are the molar enthalpy
ice is the only solid phase formed in the experimental freezing point of formation and number of moles of solution after CO2 absorption,
depression data of both unloaded and CO2 loaded aqueous amine and HCO2 and nCO2 are the molar enthalpy of formation and number
solutions, Eq. (16) can be reduced to of moles of CO2 gas absorbed in the solution.
At constant pressure and composition, the total enthalpy of for-
G0 mation of an electrolyte solution can be calculated as
− = ln aw (28)
RT 
nH = ñw Hw + ñi Hi (33)
where, G0 = 0w − 0ice is the change in standard state chemical i
potential between the ice and liquid water and aw is the activity of
water. where, Hw and Hi are the partial molar enthalpy of water and com-
ponent (solute) i, respectively. ñw and ñi are the number of moles
of water and component i, which is the equilibrium composition
3.4. Liquid–liquid equilibrium
achieved after the establishment of speciation equilibrium.
The partial molar enthalpies of water and solute can be calcu-
An important characteristic of the aqueous DEEA-MAPA solu-
lated from the temperature derivatives of the chemical potentials at
tions is that they split into two liquid phases when reacted with
constant pressure and composition. Therefore, the differentiation
CO2 . The equilibrium condition is that the chemical potential of
of Eq. (2) (chemical potential of water) with respect to temperature
component i is identical in the two liquid phases, I and II.
at constant pressure and composition gives
Ii = IIi (29)  
 
 
∂ w /RT ∂ 0w /RT ∂ ln w
= + (34)
Since the same standard state is used for each independent com- ∂T ∂T ∂T P,x
P,x P,x
ponent in the two phases, the standard state chemical potential on
both sides of Eq. (29) will cancel out and reduces to the following Hw Hw0 ex
Hw
expression = + (35)
RT 2 RT 2 RT 2
aIi = aIIi (30) where, Hw 0 is the standard state molar enthalpy of formation and

Hwex is the partial molar excess enthalpy of water.


where, aIi and aIIi are the activity of component i in the liquid phases Similarly, the differentiation of chemical potential of solute (Eq.
I and II, respectively. When liquid–liquid equilibrium is considered, (4)) with respect to temperature at constant pressure and compo-
the components are not individual ions, but neutral species formed sition leads to
by the ions. Therefore, the criterion for the LLE calculations is that  
 
 
the activity of each independent component is the same in both ∂ i /RT ∂ ∗i /RT ∂ ln i∗
liquid phases. = + (36)
∂T ∂T ∂T P,x
P,x P,x

4. Thermal calculations Hi Hi∗ Hiex


= + (37)
RT 2 RT 2 RT 2
4.1. Excess ethalpy
where, Hi∗ is the standard state molar enthalpy of formation and Hiex
Thermal property data such as excess enthalpy are very use- is the partial molar excess enthalpy of solute i. By using Eqs. (35)
ful for determination of the UNIQUAC surface area parameter (q) and (37), the total enthalpy of formation of an electrolyte solution
because the contribution to the excess enthalpy is proportional at constant pressure and composition (Eq. (33)) becomes
to the q parameter. Experimental excess enthalpy data for only     
the water-DEEA system are available in the literature for mod-
0
nH = ñw Hw ex
+ Hw + ñi Hi∗ + Hiex (38)
eling (Mathonat et al., 1997). The symmetrical excess enthalpy i
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 405

The molar enthalpy of absorbed CO2 can be calculated as 6. Extended UNIQUAC model

T Extended UNIQUAC model, an activity coefficient model for


0 0
HCO2 = HCO + Cp,CO dT (39) electrolyte systems (Thomsen et al., 1996; Thomsen, 1997;
2 2
Thomsen and Rasmussen, 1999; Thomsen, 2009), is used in this
T0
work for thermodynamic modeling of liquid–liquid phase change
0
where, HCO is the standard state molar enthalpy of formation, solvent system. It consists of a short-range contribution from the
2
0 original UNIQUAC model (Abrams and Prausnitz, 1975; Maurer and
T0 = 298.15K is the standard state temperature, and Cp,CO is the
2 Prausnitz, 1978) combined with the Debye-Hückel long-range term
standard state heat capacity of CO2 .
which takes into account the electrostatic interactions of the ionic
species present.
5. Standard state properties Excess Excess Excess Excess
GExtended UNIQUAC
= GCombinatorial + GResidual + GDebye−H ückel
(46)
The standard state properties (Gibbs energy and enthalpy of for- The first two terms, the combinatorial (or entropic) and the residual
mation, and heat capacity) of different components and species (or enthalpic) terms, are identical to the terms used in the UNIQUAC
used in this work were taken from the NIST tables given at 25 ◦ C. The model. The combinatorial term requires two adjustable parameters
values of chemical potential (Gibbs energy) at temperature of inter- per species (volume and surface area parametes, r and q, respec-
est can be calculated from its values at 25 ◦ C (298.15 K) by using tively) and the residual term requires two adjustable parameters
Gibbs-Helmholtz equation. (u0ij and uTij ) per pair of binary species to calculate the interaction
 
d ln K d G0 /RT H 0 energy parameter uij from uij = u0ij + uTij (T − 298.15).
− = =− (40) The third term in the Extended UNIQUAC model is derived
dT dT RT 2
from the extended Debye-Hückel law. This term has no adjustable
Eq. (40) can be integrated from the standard state temperature parameter. It only represents the electrostatic interactions of the
(T0 = 298.15 K) to temperature T to get the following expression ionic species. The following expression of excess Gibbs energy is
used for the Debye-Hückel contribution to the model
T
H 0 2
ln KT − ln KT0 = dT (41) E
GDH 4A  √ √ b2 I
RT 2 = −xw Mw ln 1 + b I − b I + (47)
T0 RT b3 2

where, KT and KT0 are the equilibrium constants at temperatures where, xw is the mole fraction and Mw (kg mol−1 ) is the molar mass
T and T0 = 298.15 K, respectively. Temperature derivative of the of water. The parameter b depends on the size of the ions. How-
enthalpy of formation of the process give the heat capacity of the ever, it is usually considered as a contant and given a value of 1.50
species involved in the process as (kg mol−1 )½ . I is the ionic strength of the electrolyte solution first
defined by Lewis and Randall (1921) as:
dH 0
Cp0 = (42) 1
dT I= mi Zi2 (48)
2
where, Cp0 is the increment in the standard state heat capacity. i

Three parameter temperature dependent heat capacity correlation where, mi is the molality and Zi is the charge number of the ionic
used in the model (Thomsen et al., 1996; Thomsen and Rasmussen, species i. A is the Debye-Hückel parameter with a value of 1.1717
1999) inspired by Helgeson et al. (1981) is given as (kg mol−1 )½ at 25 ◦ C.
∗ ci  3/2
Cp,i = ai + bi T + (43) 1/2 e2
T − T,i A = (2 NA d0 ) (49)
4 ε0 εr kT
A constant value of 200 K is given to T,i for all the components.
By combining Eqs. (42) and (43) and integrating leads to where, NA is Avogadro’s number (6.023 × 1023 mol−1 ), d0 is the den-
sity of water in kg m−3 , e is the electron charge (1.60206 × 10−19 C),
   T −T 
 ε0 is the permittivity in vacuum (8.8542 × 10−12 C2 J−1 m−1 ), εr is
HT0 = HT0 + a (T − T0 ) + 0.5b T 2 − T02 + c ln
0 T0 − T the dielectric constant (or relative permittivity) of solvent (dimen-
(44) sionless) 78.4 for water at 298.15 K, k is the Boltzmann’s constant
(1.381 × 10−23 J K−1 ), and T is the temperature in Kelvin. Based on
the density of pure water and the dielectric constant of water, the
HT0 is the increment in standard state enthalpy at temperature Debye-Hückel parameter can be approximated in the temperature
T . a, b, and c are the increment in the heat capacity correlation range of 240 K < T < 540 K (Sander et al., 1986) by
parameters a, b, and c for the system considered. 
Inserting Eq. (44) in Eq. (41) and integrating lead to the following A = 1.131 + 1.335 × 10−3 (T − 273.15) + 1.164
form of the equilibrium constant  1/2
1    ×10−5 (T − 273.15)2 kgmol−1 (50)
1 T T0
R ln KT = R ln KT0 − HT0 − + a ln − −1
0 T T0 T0 T
 2
   The aqueous phase activity coefficients were calculated from
(T − T0 ) c T − T  T − T T the Extended UNIQUAC model and the vapor phase fugacity coef-
+0.5b + ln − ln (45)
T T T T0 − T T0 ficients were calculated from the SRK cubic equation of state.
The model only requires volume and surface area parameters
This equation can be used to calculate the composition of the per species and interaction energy parameters per pair of binary
solution at the temperature T if the activity coefficients are known species. The Debye-Hückel and the SRK terms have no adjustable
at this temperature. parameters.
406 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

7. Estimation of model parameters The parameter estimation was done by using two different rou-
tines: a modified Marquardt routine (Fletcher, 1971) and Nelder &
The model parameters were estimated by performing a least- Mead routine for function minimization (Nelder and Mead, 1965).
square minimization of the weighted sum of squared residuals (S) The modeling strategy consists of, firstly, the parameter estima-
as tion of the two sub-systems (H2 O-DEEA-CO2 and H2 O-MAPA-CO2 )

2
  P vap − Pexp
vap 2
separately followed by the H2 O-DEEA-MAPA-CO2 system where
 Pcalc − Pexp all the parameters were fitted to the experimental data simulta-
S= calc
vap +  
0.0125Pexp 0.06 Pexp + 0.01 neously. The parameter estimation for the two sub-systems were
P vap data VLE data
⎛  ⎞2
started with the calculation of vapor pressures of pure amines fol-

  H E − Hexp 2
G0 + RT vi ln ai lowed by the binary CO2 unloaded data (unloaded freezing point,
 ⎜ ⎟ E
excess enthalpy and unloaded VLE data) and then ternary CO2
+ ⎜ i ⎟ + calc (51)
⎝ 0.0015RT ⎠ 100Rx loaded data (loaded freezing point and VLE data and the heat
SLE data H E data of absorption data). Once a reasonable set of parameters were
  H Abs − Hexp 2    ln aII − ln aI 2
obtained for each of the H2 O-DEEA-CO2 and H2 O-MAPA-CO2 sub-
Abs
+ calc
+ i i systems, binary parameters across different species of DEEA and
500Rx 0.5 MAPA in the two sub-systems were estimated by using the ternary
H Abs data LLE data i
unloaded data (unloaded freezing point and VLE data). Then qua-
where, “calc” and “exp” represent the calculated values (by the ternary CO2 loaded data (VLE and heat of absorption data) were
model) and experimental data. The factors 0.0125, 0.06, 0.0015, introduced and all the pre-estimated parameters were refitted to
100, 500, and 0.5 are the weighting factors respectively used for the the experimental data simultaneously to get a new set of param-
pure amine vapor pressure, vapor-liquid equilibrium (VLE), solid- eters. Finally, the LLE data were introduced and a final set of
liquid equilibrium (SLE), excess enthalpy (H E ), heat of absorption parameters were estimated which can reproduce all data values
(H Abs ), and liquid–liquid equilibrium (LLE) data. These weighting and describe all data types (pure amine vapor pressure, excess
factors were optimized on the basis of experience in modeling the enthalpy, VLE, SLE, LLE and heat of absorption of CO2 ) with a single
H2 O-DEEA-MAPA-CO2 and its sub-systems. set of parameters.
The term P vap represents the pure amine vapor pressure (bar). The deviations between the model results and the experimental
The term P in the VLE data represents either the total pressure (bar) data are given in Eq. (52) as absolute average relative deviation
above the solutions or CO2 partial pressure data in bar. In the term (AARD)
for SLE data, G0 is the change in standard state Gibbs energy
(J mol−1 ) between solid (ice in this case) and liquid phases, vi is 1  ˚calc − ˚exp
AARD = | | × 100% (52)
stoichiometric coefficient of component “i”, ai is activity of compo- n ˚exp
n
nent “i”, R is gas constant (8.314 J mol−1 K−1 ) and T is temperature
in Kelvin. x = 1 K is included to make both excess enthalpy and heat where, ˚ is the type of data and n is the number of data points. “calc”
of absorption terms dimensionless. In the liquid–liquid equilibrium and “exp” represent the calculated data values (by the model) and
(LLE) data term, aIi and aIIi represent activity of component “i” in the experimental data respectively.
two liquid phases I and II. 94 model parameters and 6 thermodynamic properties were fit-
The weighting factor for vapor pressure data was chosen so that ted to approximately 1500 experimental data consisting of pure
1.25% difference between the calculated and experimental pres- amine vapor pressure, vapor-liquid equilibrium, solid-liquid equi-
sures would give a squared residual of 1. Similarly, the weighting librium, liquid–liquid equilibrium, excess enthalpy, and heat of
factor was assigned to relatively high pressure VLE data so that 6% absorption of CO2 . The 94 model parameters are 6 volume (r) and
difference between the calculated and experimental total pressures 6 surface area (q) parameters, 41 u0ij and 41 uTij binary parame-
or CO2 partial pressures would give a squared residual of 1. It can be
ters for calculating the interaction energy parameters uij = u0ij +
noticed that 0.01 bar is added to the experimental data in denom-
inator in the VLE term in order not to give high weight to the very uTij (T − 298.15). The 6 thermodynamic properties are the standard
low pressure data. The addition of this number represents that an state Gibbs energy of formation and standard state enthalpy of
experimental pressure of 0.01 bar would lead to a squared residual formation for the three species DEEA (l), MAPA (l), and MAPACOO− .
of 1 if the calculated pressure data deviate from the experimental The Extended UNIQUAC model parameters of the aqueous phase
pressures by 12%. The calculated pressures in the objective func- H+ and OH− were taken from Thomsen et al., 1996 and CO2 (aq),

tion (Eq. (51)) were the bubble point pressures. Only experimental CO2− 3 , and HCO3 from Thomsen and Rasmussen, 1999. The r and
total pressure data were used for the parameter estimations when q parameters, and the binary parameters u0ij and uTij for calculating
both total pressure and CO2 partial pressure data were available. the interaction energy parameters uij = u0ij + uTij (T − 298.15) for the
The CO2 partial pressure data were only used when total pressure six species DEEA (aq), DEEAH+ , MAPA (aq), MAPAH+ , + HMAPAH+ ,
data were not available. and MAPACOO− were determined in this work by fitting to the
The weighting factor for SLE data was chosen so that the solubil- experimental data.
ity indexes of 1.0015 and 0.9985 would give a square residual of 1. The estimated r and q parameters are given in Table 1. The esti-
Solubility index of a salt (ice in this work) is defined as  the ratio of mated binary parameters u0ij and uTij for each pair of species are
activity product of a salt by its solubility product, (SI = i avi i /Ksp ).
given respectively in Tables 2 and 3. Some of the u0ij parameters
The excess enthalpy and heat of absorption data were weighted
so that an absolute difference of 831.4 J (for excess enthalpy) and were given a very high value of 1010 with corresponding values of
4157 J (for heat of absorption) between the calculated and exper- 0 for the uTij parameters. These values indicate that those pairs of
imental data would lead to a squared residual of 1 for each term. species are most likely to have no interaction.
x = 1 K is included to make both excess enthalpy and heat of absorp- The values of standard state Gibbs energy and enthalpy of for-
tion terms dimensionless. The LLE data were weighted so that an mation are given in Table 4. These values were taken from the NIST
absolute difference of 0.5 between the natural logarithm of activity tables (NIST, 1990) for most of the components. For DEEA (g) and
of species “i” in the two liquid phases (I and II) at equilibrium would MAPA (g), these values were estimated by the Marrero and Gani
give a square residual of 1. method (Marrero and Gani, 2001) using the ProPred (property pre-
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 407

Table 1 Table 4
UNIQUAC volume (r) and surface area (q) parameters fitted to the experimental data. Standard state thermodynamic properties estimated from experimental data in this
Bold values are determined in this work. work. Bold values are determined in this work.

r q Gf0 / (kJ mol−1 ) Hf0 / (kJ mol−1 )


a a
H2 O 0.92 1.4 H2 O (l) −237.129a −285.83a
CO2 (aq) 5.741c 6.0806c H2 O (g) −228.572a −241.818a
MAPA (aq) 11.96 8.845 H2 O (s) −236.5376a −292.6244a
DEEA (aq) 3.225 3.641 CO2 (aq) −385.98a −413.8a
H+ 0.13779b 10−15 b CO2 (g) −394.359a −393.509a
MAPAH+ 4.499 9.664 H+ 0a 0a
+
HMAPAH+ 1.391 1.749 OH- −157.2481a −230.2433a
DEEAH+ 4.583 2.918 CO2- −527.81a −677.14a
3
OH- 9.3973b 8.8171b HCO-3 −586.77a −691.99a
CO2-
3 10.828c 10.769c DEEA (g) −29.36b −257.49b
HCO-3 8.0756 c 8.6806c MAPA (g) 142.73b −37b
MAPACOO- 9.922 6.499 DEEA (l) −44.903 −315.11
a MAPA (l) 129.9 −88.227
Abrams and Prausnitz (1974).
b DEEA (aq) −55.999 −333.560
Thomsen et al. (1996).
c MAPA (aq) 112.659 −141.062
Thomsen and Rasmussen (1999).
DEEAH+ −111.2663 −370.92
MAPAH+ 52.299 −176.392
+
HMAPAH+ 3.139 −220.142
diction) tool in ICAS (Integrated Computer Aided System) software MAPACOO- −255.71 −577.52
(CAPEC Software, 2014). a
NIST table (NIST, 1990).
The standard state Gibbs energy of formation for DEEA (l) and b
Determined from Marrero and Gani method by using ProPred (property predic-
MAPA (l) were calculated from their corresponding values in the tion) tool in ICAS (Integrated Computer Aided System) software (Marrero and Gani,
pure gas phase (DEEA (g) and MAPA (g)) by using Eq. (24). For 2001; CAPEC Software, 2014).
DEEA (aq) and MAPA (aq), the standard state Gibbs energy of for-
mation were estimated from their corresponding values in the pure
liquid phase (DEEA (l) and MAPA (l)) plus a contribution of activ- order, pKa = 9.7 at 25 ◦ C). The standard state Gibbs energy of for-
ity coefficient at infinite dilution. Littel et al. (1990) and Hamborg mation of DEEAH+ was then calculated from r G0 = f GDEEA(aq)
0 −
and Versteeg (2009) reported the dissociation constant (ln Ka ) of f G0 . Aronu et al. (2011) reported the first and second dis-
DEEAH+
protonated DEEA as a function of temperature. These dissocia- sociation constants for the mono and di protonated MAPA at
tion constant data were correlated to estimate the Gibbs energy different temperatures. These experimental data were correlated
of the dissociation reaction of protonated DEEA (Eq. (9) in reverse to calculate the Gibbs energy of reaction for the dissociation of

Table 2
u0ij = u0ji parameters for calculating UNIQUAC interaction energy parameters (uij = u0ij + uTij (T − 298.15)). Bold values are determined in this work.

H2 O DEEA MAPA CO2 H+ DEEAH+ MAPAH+ +


HMAPAH+ CO2-
3 HCO-3 OH- MAPACOO-

H2 O 0a
DEEA −4.3632 0
MAPA −125.7 106.25 0
CO2 41.0717a −311.22 287.39 40.5176a
H+ 105 a 1010 1010 1010 a 0a
DEEAH+ −150.15 473.38 460.86 −178.08 1010 0
MAPAH+ −455.55 −216.11 −13.017 113.99 1010 −751.1 0
+
HMAPAH+ 1256.2 2566.3 −64.801 552.76 1010 7.0431 −292.5 0
CO2-
3 361.388a 1581.9 612.11 1010 a 1010 a 1042.2 458.46 135.84 1458.344 a
HCO-3 577.050a 665.78 819.93 651.045a 1010 a 799.14 419.3 1888.7 800.008 a 771.04 a
OH- 600.495a 579.77 931.69 1010 a 1010 a 1010 1010 1010 1588.025 a 1010 a 1562.88 a
MAPACOO- −456.94 −250.18 379.78 1010 1010 −227.59 −646.89 2.9472 1121.9 342.24 592.19 1000
a
Thomsen and Rasmussen (1999).

Table 3
uTij = uTji parameters for calculating UNIQUAC interaction energy parameters (uij = u0ij + uTij (T − 298.15)). Bold values are determined in this work.

H2 O DEEA MAPA CO2 H+ DEEAH+ MAPAH+ +


HMAPAH+ CO2-
3 HCO-3 OH- MAPACOO-

H2 O 0a
DEEA 1 0
MAPA 0.533 0.9511 0
CO2 7.5184a 8.288 1.822 13.629a
H+ 0a 0 0 0a 0a
DEEAH+ 0.1918 1.816 −1.132 3.745 0 0
MAPAH+ −0.4654 −2.605 −8.025 7.793 0 −3.978 0
+
HMAPAH+ 7.318 −9.42 3.532 6.816 0 −1.494 9.693 0
CO2-
3 3.3516a 1.02 3.382 0 0a −3.486 1.476 4.128 −1.3448a
HCO-3 −0.38795a −0.3963 0.5863 2.773 a 0a 2.277 −0.1296 1.33 1.7241a −0.0198a
OH- 8.5455a 9.2 −0.2957 0a 0a 0 0 0 2.7496b 0a 5.6169a
MAPACOO- 0.01686 2.825 −2.191 0 0 −3.352 −5.564 2.22 0.4127 3.809 7.164 0
a
Thomsen and Rasmussen (1999).
b
The value is corrected from the previous value (2.5176, an editing mistake) as reported in Thomsen and Rasmussen (1999).
408 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

Table 5 2.5
Parameters for the standard state heat capacity (Eq. (43)), fitted to the experimental
data in this work. Bold values are determined in this work. Extended UNIQUAC

Vapor Pressure of DEEA / bar


−1 −1 −1 −2 −1 2 Steele et al., 2002
a/ (J mol K ) b/ (J mol K ) c/ (J mol )
Kapteina et al., 2005
a a a
H2 O (l) 58.3695 0.03896 523.88
Klepáčová et al., 2011
H2 O (g) 33.577b 0 0 1.5
Hartono et al., 2013
H2 O (s) 47.8996c 0 0
CO2 (aq) 243d 0d 0d
CO2 (g) 37.11b 0 0 1
H+ 0a 0a 0a
OH- 1418.2a −3.446a −51473.13a
CO2-
3 894.688d −2.8272d −21149.44d 0.5
HCO-3 −0.6771d 0.27375d −10089.51d
DEEA (g) 173e 0 0
MAPA (g) 136e 0 0 0
DEEA (l) 104.952f 0.5936f 0 0 50 100 150 200
MAPA (l) 147.111e 0.2546e 0
DEEA (aq) 210.9899 0.35675 0
T / oC
MAPA (aq) 485.4593 −0.49787 0
DEEAH+ −32.1278 0 0 Fig. 1. Vapor pressure of pure DEEA as a function of temperature.
MAPAH+ 178.5593 0 0
+
HMAPAH+ −114.9407 0 0
MAPACOO- 358.2 −0.1178 −10613.4 8. Results and discussion
a
Thomsen et al. (1996).
b
Wagman et al. (1982) (NIST). To model the liquid–liquid DEEA-MAPA system, the model
c
Fosbøl et al. (2009). parameters were estimated first for the two sub-systems
d
Thomsen and Rasmussen (1999). H2 O-DEEA-CO2 and H2 O-MAPA-CO2 followed by the H2 O-DEEA-
e
Determined from Joback and Reid method by using ProPred (property predic- MAPA-CO2 system which give liquid–liquid split. The results of the
tion) tool in ICAS (Integrated Computer Aided System) software (Joback and Reid,
1987; CAPEC Software, 2014).
two ternary sub-systems and the quaternary liquid–liquid systems
f
Correlated from the experimental data reported by Maham et al. (1997) and are given in the following sections. As described earlier, the param-
Rayer et al. (2012). eter estimation was done by starting with the simple system e.g.
pure amine vapor pressure followed by the addition of binary and
ternary data and finally the quaternary data. The data were sys-
tematically added and the relevant parameters were estimated.
protonated MAPA (Eq. (10) in reverse order, pKa = 10.6 at 25 ◦ C) However, the results are presented here according to the data type
and dissociation of di protonated MAPA (Eq. (11) in reverse order, for each system separately.
pKa = 8.6 at 25 ◦ C). Similarly, the standard state Gibbs energy of for-
mation of MAPAH+ and + HMAPAH+ were estimated respectively 8.1. H2 O-DEEA-CO2 system
0
from r G0 = f GMAPA(aq) − f G0 0 0
+ and r G = f G
MAPAH + − MAPAH
f G+0 . No information was found in the literature about the The results of pure DEEA vapor pressure at different tempera-
HMAPAH+
thermodynamic properties or equilibrium constant of the MAPA tures are given in Fig. 1. Experimental data of vapor pressure of
carbamate reaction. A guess values for the standard state Gibbs pure DEEA reported by several authors in the literature (Steele
energy and enthalpy of formation were used for the MAPA carba- et al., 2002; Kapteina et al., 2005; Klepáčová et al., 2011; Hartono
mate (MAPACOO− ) which were then altered during the parameter et al., 2013) are plotted along with the calculated curve. The model
estimation and fitted to all the experimental data in this work. describes the vapor pressure data very well. However, some devi-
The a, b, and c parameters for the standard state three- ations at high temperatures can be observed. The absolute average
parameters heat capacity correlation (Thomsen et al., 1996) relative deviations (AARD) between the model results and experi-
presented in Eq. (43) are given in Table 5. mental data are 6.5% for Steele et al. (2002); 20.3% for Kapteina et al.
The three parameters of the standard state heat capacity for (2005); and 5% each for Klepáčová et al. (2011) and Hartono et al.
the species DEEA (g), MAPA (g), and MAPA (l) were estimated from (2013). The AARD for all the vapor pressure data together from all
Joback and Reid method by using ICAS software (CAPEC Software, the literature sources is 9.9%.
2014; Joback and Reid, 1987). Maham et al. (1997) and Rayer Freezing point depression data have been reported by Arshad
et al. (2012) reported the molar heat capacity data of DEEA at dif- et al. (2013a) for both the unloaded and CO2 loaded aqueous DEEA
ferent temperatures. These data were correlated to estimate the solutions and these data were included in the parameter estimation
standard state heat capacity parameters for DEEA (l). The stan- in this work. Such data are very important to compute the water
dard state heat capacity parameters for MAPACOO− were estimated activity, a key parameter to estimate the amount of water evapo-
based on the assumption that the change in heat capacity of the ration in stripping section during the solvent regeneration in the
MAPA carbamate reaction (Eq. (12)) is constant (and equal to zero capture process. Freezing point depression data at different molal
i.e. CP = 0). This is based on the “Principle of Balance of Iden- concentrations of aqueous DEEA solutions (H2 O-DEEA) are plotted
tical Like Charges” as suggested by Murray and Cobble (1980). with the calculated curve as shown in Fig. 2. For H2 O-DEEA-CO2
According to this principle, the CP is almost constant if there system (CO2 loaded system), the freezing points are presented in
is a balance of ionic charge of similar species on both sides of Fig. 3 for four different DEEA solutions (12, 20, 30, and 33 mass%) at a
the equilibrium reaction. This principle was applied to the equi- varying CO2 concentration in the solutions. The model describes the
librium reaction in Eq. (12) to estimate the parameters of the freezing point data very well both in the unloaded (Fig. 2) and the
standard state heat capacity for the MAPA carbamate. The standard CO2 loaded (Fig. 3) DEEA solutions. However, there are some devia-
state heat capacity parameters of the remaining species were esti- tions in the CO2 loaded solutions at high DEEA concentrations of 30
mated by the model. The standard state heat capacity for DEEAH+ , and 33 mass%. These two curves showed a decrease in the freezing
MAPAH+ , and + HMAPAH+ were considered temperature indepen- points by increasing the CO2 concentrations up to a certain point
dent. and then the freezing points became almost constant. The model
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 409

0 calculated the trend lines very nicely with slightly higher calculated
freezing points throughout the curves and the slightly lower calcu-
-5 lated freezing points in the last couple of points in both the curves.
The AARD between the calculated and the experimental freezing
points are 4.2% and 3.5% respectively for the unloaded and the CO2
Freezing Point / oC

-10
loaded DEEA solutions.
Another important use of freezing point depression data is to
-15
compute the amount of chemical species present in the solution
(speciation) (Fosbøl et al., 2011). The extended UNIQUAC model
-20 calculated speciation in 12 and 33 mass% DEEA solutions are pre-
Extended UNIQUAC sented in Fig. 4. The mole fraction of different species (left ordinate)
-25 water-DEEA together with the pH of the solution (right ordinate) are plotted as a
function of CO2 concentration. It can be seen that, on dissolution of
-30 CO2 into the solution, the concentration of aqueous DEEA decreases
0 2 4 6 8 10 12 upon reaction with CO2 . This results in formation of protonated
DEEA Molality / mol DEEA.(kg H2O)-1 DEEA, carbonate, and bicarbonate species. This can be seen from
increment in concentration of these three species. It can also be
Fig. 2. Freezing point depression of aqueous DEEA system. noticed that the formation of carbonate ions first increases, reach-
Experimental data from Arshad et al. (2013a). ing a maximum, and then decreases back to the minimum. When
aqueous DEEA is fully consumed, the concentration of aqueous CO2
starts building up in the solution. This can be seen in 12 mass% DEEA
0 case. The speciation behavior of H2 O-DEEA-CO2 system is similar
Extended UNIQUAC to that of H2 O-AMP(2-amino-2-methyl-1-propanol)-CO2 system,
12 mass % DEEA qualitatively (Richner and Puxty, 2012). When CO2 , an acid gas,
-4 20 mass % DEEA is dissolved in an amine (base) solution, the pH of the solution is
30 mass % DEEA decreased and the solution become more acidic. This is indicated
Freezing Point / oC

33 mass % DEEA by the dashed lines in the figure.


-8
Vapor-liquid equilibrium data are essential for the design and
modeling of the unit operations in the capture process. Vapor-liquid
-12 equilibrium data for the CO2 loaded aqueous DEEA solutions (both
total pressure and CO2 partial pressure) were reported by Monteiro
et al. (2013) and Arshad et al. (2014a,b). The experimental total
-16 pressure data from both the literature sources along with the cal-
culated curves for the 5 M DEEA solutions are exhibited in Fig. 5 as
-20
a function of CO2 concentration at different temperatures. There is
0 1 2 3 4 5 a good agreement between the experimental data and the model
CO2 Molality / mol CO2.(kg H2O)-1 results. However, the model calculated total pressures are slightly
higher for the 313.15 K isotherm and slightly lower for the two
Fig. 3. Freezing point depression of H2 O-DEEA-CO2 systems at different composi- isotherms at 353.15 K and 393.15 K at high CO2 concentrations. The
tions of DEEA and CO2 . estimated AARD between the model results and the experimental
Experimental data from Arshad et al. (2013a).
data are 7.6% for Monteiro et al. (2013) and 10.4% for Arshad et al.

DEEA(aq) DEEAH+ DEEA(aq) DEEAH+


CO3-- HCO3- CO3-- HCO3-
CO2(aq) pH of soluon CO2(aq) pH of soluon
0.025 11.5 0.075 11.5

12 mass % DEEA 33 mass % DEEA

0.02 10.5 0.06


10.5
mole fracon, xi / (-)
mole fracon, xi / (-)

pH of soluon

pH of soluon

0.015 9.5 0.045

9.5

0.01 8.5 0.03

8.5
0.005 7.5 0.015

0 6.5 0 7.5
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 4. Model calculated speciation for 12 and 33 mass% DEEA solutions.


410 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

12 12
Extended UNIQUAC Extended UNIQUAC
5M DEEA 313.15 K - Arshad et al., 2014 2M DEEA
353.15 K - Monteiro et al., 2013
353.15 K - Arshad et al., 2014 10
10 373.15 K - Monteiro et al., 2013
393.15 K - Arshad et al., 2014
393.15 K - Monteiro et al., 2013
353.15 K - Monteiro et al., 2013
373.15 K - Monteiro et al., 2013 8
8 393.15 K - Monteiro et al., 2013

P total / bar
P total / bar

6
6

4
4
2

2
0
0 0.5 1 1.5 2 2.5 3
0 CO2 Molality / mol CO2.(kg H2O)-1
0 3 6 9 12 15
CO2 Molality / mol CO2.(kg H2O)-1 Fig. 7. Equilibrium total pressure in 2 M DEEA solutions as a function of CO2 com-
position at different temperatures.
Fig. 5. Equilibrium total pressure in 5 M DEEA solutions as a function of CO2 com-
position at different temperatures.
0.2
Experimental data from Arshad et al. (2014a) and Monteiro et al. (2013).
Extended UNIQUAC
2M DEEA
313.15 K - Monteiro et al., 2013
0.16 333.15 K - Monteiro et al., 2013
10
353.15 K - Monteiro et al., 2013
5M DEEA
0.12
P CO2 / bar

0.08
P CO2 / bar

0.1

Extended UNIQUAC
0.01 313.15 K - Arshad et al., 2014 0.04
353.15 K - Arshad et al., 2014
393.15 K - Arshad et al., 2014
0.001 313.15 K - Monteiro et al., 2013 0
333.15 K - Monteiro et al., 2013 0 0.5 1 1.5 2 2.5
353.15 K - Monteiro et al., 2013 CO2 Molality / mol CO2.(kg H2O)-1
0.0001
0 2 4 6 8 10 12 14 Fig. 8. Partial pressure of CO2 in 2 M DEEA solutions at different temperatures.
CO2 Molality / mol CO2.(kg H2O)-1
contents. This method gave high errors in the determined CO2
Fig. 6. Partial pressure of CO2 in 5 M DEEA solutions at different temperatures.
Experimental data from Arshad et al. (2014a) and Monteiro et al. (2013). amount when CO2 contents are low in the solutions. By excluding
the data below a CO2 loading of 0.02 mol CO2 (mol amine)−1 from
the regression, they reported a lower numbers of AARD of 8.1% for
(2014a). Similarly, Fig. 6 presents the model results of the CO2 par- the total pressure (compared to the 10.2% with all data included)
tial pressure together with the experimental data in the 5 M DEEA and 15.5% for CO2 partial pressure (compared to the 18.6% with
solutions as a function of the CO2 concentration. It can be seen that all data included). However, the deviations reported in our work
the model represents the experimental CO2 partial pressures well are obtained by including all the data in the parameter estimation.
with the estimated AARD of 19.1% for Monteiro et al. (2013) and It can be seen that the deviations in this work are slightly higher
22.4% for Arshad et al. (2014a). than those reported by Monteiro et al. (2013) But it should also be
The model results and the experimental data from Monteiro noted that the additional experimental VLE, heat of absorption and
et al. (2013) for the 2 M DEEA solutions are given in Fig. 7 for the freezing point data were used in this work and more importantly
total pressure and Fig. 8 for the partial pressure (solubility) of CO2 the model describes and reproduces all the data sets and data types
at three different temperatures. A good agreement between the for H2 O-DEEA-CO2 system together with H2 O-MAPA-CO2 and H2 O-
experimental data and the calculated curves can be seen with some DEEA-MAPA-CO2 systems with a single set of parameters.
deviations in the total pressures at high CO2 concentrations for Molar excess enthalpy and heat of absorption of CO2 in the aque-
the two isotherms at 353.15 K and 373.15 K (Fig. 7). The estimated ous DEEA solutions were the thermal property data available for
AARD are 10.1% and 14.1% respectively for the total pressure and the thermodynamic modeling. Fig. 9 presents the model results
CO2 partial pressure data. for the molar excess enthalpy in DEEA solutions together with the
Monteiro et al. (2013) presented the experimental VLE mea- experimental data from Mathonat et al. (1997) The model very well
surements and the thermodynamic modeling of the H2 O-DEEA-CO2 describes the excess enthalpy in the DEEA solutions with an AARD
system using the e-NRTL model. They reported that the large scat- of 7.2% between the experimental data and the calculated curve.
ter in the experimental data at low CO2 loadings have contributed Arshad et al. (2013b) reported the differential heat of absorption
greatly in the estimated deviations (AARD). This scatter in the data of CO2 in 5 M DEEA solutions at different temperatures and these
is inherited in the analytical method used for determining the CO2 data were included in the parameter estimation. The modeling
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 411

0 Freezing point depression results for the unloaded and CO2


loaded MAPA solutions are given in Figs. 13 and 14, respectively.
500 Extended UNIQUAC
The experimental freezing point data were available for the param-
eter estimation from Arshad et al. (2013a). Fig. 13 exhibits the
Mathonat et al., 1997
-1000 experimental and calculated freezing points for the aqueous MAPA
H E / J.mol-1

solutions with different concentrations. A very good agreement can


-1500
be observed between the experimental data and the model results
with an AARD of 3.4%. For the CO2 loaded aqueous MAPA solutions
(H2 O-MAPA-CO2 ), Fig. 14 presents the calculated and experimental
-2000
freezing points in different aqueous MAPA solutions (10, 20, and 27
mass%) at a varying CO2 concentrations. The model describes the
-2500 wave like curves very well for all the three CO2 loaded MAPA sys-
tems in the whole CO2 concentration range with an AARD of just
-3000 3.1%.
0 0.2 0.4 0.6 0.8 1 Fig. 15 presents the model calculated speciation in 10 and 27
mol fracon of DEEA mass% MAPA solutions. The mole fraction of different species (left
ordinate) together with the pH of the solution (right ordinate) are
Fig. 9. Excess enthalpy of DEEA as a function of mole fraction of DEEA.
plotted as a function of CO2 concentration. It can be observed that
dissolution of CO2 in aqueous MAPA solutions results in formation
results are shown in Fig. 10. The model calculated the differential of MAPA carbamate, di-protonated MAPA, and carbonate species.
heats of absorption of CO2 in 5 M DEEA solutions reasonably well. The formation of protonated MAPA is very low and its significant
The estimated deviation (AARD) is 11.6% between the experimental amount is present in 10 mass% MAPA case. An interesting speci-
and the model calculated results. ation behavior can also be seen after CO2 concentration of 1 and
The modeling results for the differential heat of absorption of 4 mol(kg H2 O)−1 respectively for 10 and 27 mass% MAPA solutions.
CO2 in 37 and 32 mass% DEEA solutions are presented in Fig. 11. At these concentrations, aqueous MAPA is almost fully consumed.
The experimental data were available from Kim (2009). It should There is a sudden drop in carbonate ions and an exponential
be noted that these data were not included in the parameter esti- increase in bicarbonate ions. This specific speciation behavior pos-
mation. A fairly good agreement between the experimental data sibly explains the wave like curves in the freezing point depression
and the model results can be seen in all the isotherms for both of CO2 loaded MAPA solutions (Fig. 14).
DEEA concentrations (37 and 32 mass%). The first experimental data The total pressure data for the H2 O-MAPA system at differ-
point seems to be an outlier in the 40 ◦ C isotherm for the 37 mass% ent temperatures were included in the parameter estimation. The
DEEA solution. At low CO2 concentrations, a large scatter can also be model results are presented in Fig. 16 along with the experimental
observed in the experimental data at 80 ◦ C for both the DEEA con- data available from Kim et al. (2008) and Monteiro et al. (2011).
centrations. The estimated AARD are 17.9% and 21.5% respectively The total pressure data are presented as a function of MAPA con-
for the 37 mass% and 32 mass% DEEA solutions. centration in the liquid phase at different temperatures. It can be
seen that the model represents the three isotherms at 313.15 K,
8.2. H2 O-MAPA-CO2 system 333.15 K and 353.15 K very well. However, some deviations can be
observed in the isotherm at 373.15 K at high MAPA concentrations.
The experimental vapor pressures of pure MAPA as a func- It can also be noted that the experimental total pressures for the
tion of temperature have been available from the literature (Kim 373.15 K isotherm are only available from Kim et al. (2008) and
et al., 2008; Verevkin and Chernyak, 2012; Hartono et al., 2013) their accuracy cannot be cross checked with other sources due to
and included in the parameter estimation. The model results are lack of data for the H2 O-MAPA system. The estimated deviations
plotted with the experimental vapor pressure data as shown in (AARD) between the model results and the experimental data are
Fig. 12. It can be seen that the calculated data are in very good agree- 4.4% and 3.7% respectively for Kim et al. (2008) and Monteiro et al.
ment with the experimental data. The deviations (AARD) between (2011).
the experimental data and the model results are 5.4% for all the The experimental equilibrium total pressure in 2 M MAPA solu-
data together and individually 2.1% for Kim et al. (2008); 15.2% for tions at different temperatures from Arshad et al. (2014a) and the
Verevkin and Chernyak (2012); and 0.9% for Hartono et al. (2013). model calculated values are plotted in Fig. 17. A fairly good agree-

-80 -100
40 oC 80 oC
-80
∆H abs / kJ.(mol CO2)-1

∆H abs / kJ.(mol CO2)-1

-60

-60
-40
-40
Extended UNIQUAC Extended UNIQUAC
-20 5M DEEA 5M DEEA
-20

0 0
0 3 6 9 12 15 0 2 4 6 8
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 10. Differential heat of absorption of CO2 in 5 M DEEA solutions at 40 and 80 ◦ C.


Experimental data from Arshad et al. (2013b).
412 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

-120 -80

40 oC 40 oC
-100

∆H abs / kJ.(mol CO2)-1

∆H abs / kJ.(mol CO2)-1


-60
-80

-60 -40

-40
Extended UNIQUAC -20 Extended UNIQUAC
-20 37 mass % DEEA 32 mass % DEEA

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-120 -120
80 oC
-100 -100 80 oC

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1

-80 -80

-60 -60

-40 -40
Extended UNIQUAC Extended UNIQUAC
-20 37 mass % DEEA -20 32 mass % DEEA

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 11. Differential heat of absorption of CO2 in 37 mass% (left side images) and 32 mass% (right side images) DEEA solutions at 40 and 80 ◦ C.
Experimental data from Kim (2009) were not used in the parameter estimation.

1.2 0

Extended UNIQUAC
1 Kim et al., 2008 -5
Vapor Pressure of MAPA / bar

Verevkin and Chernyak, 2012


Freezing Point / oC

Hartono et al., 2013


0.8 -10

0.6 -15

0.4 -20
Extended UNIQUAC
water-MAPA
0.2 -25

0 -30
0 20 40 60 80 100 120 140 160 0 1 2 3 4 5 6
T / oC MAPA Molality / mol MAPA.(kg H2O)-1

Fig. 12. Vapor pressure of pure MAPA as a function of temperature. Fig. 13. Freezing point depression in aqueous MAPA system.
Experimental data from Arshad et al. (2013a).

ment between the experimental and the calculated values can be tal data for the 393.15 K isotherm in the whole CO2 concentration
seen at 313.15 K. The model also calculated the total pressures fairly range which contributed the major part to the estimated devia-
good for 353.15 K isotherm at low pressures but deviate at pres- tion (AARD) of 55.1% for Arshad et al. (2014a). Similarly, the model
sures above 3 bar. Similar kind of deviations can also be observed calculated values of CO2 solubility have good agreement with the
at high pressures for the isotherm at 393.15 K. The ARRD between experimental data at 333.15 K from Pinto et al. (2014a). However,
the experimental and the model calculated total pressures is 16.1%. the model results show deviations for the two isotherms at 313.15 K
The Extended UNIQUAC model calculated solubility (partial and 353.15 K both at low and high CO2 partial pressures. Some
pressure) of CO2 in 2 M MAPA solutions and the experimental data scatter in the experimental data can also be observed in both the
from Arshad et al. (2014a) and Pinto et al. (2014a) are compared in isotherms at 313.15 K and 353.15 K. The estimated AARD value for
Fig. 18 at the temperature ranging between 313.15 K and 393.15 K the data set is 37.2%.
(40–120 ◦ C). The model results are quite fair in the whole CO2 con- The experimental total pressure and CO2 partial pressure data
centration range for the 313.15 K and 353.15 K isotherms data from as a function of CO2 concentration for 1 M MAPA solutions from
Arshad et al. (2014a). However, the model calculated results of Arshad et al. (2014a) are compared with the modeling results in
CO2 partial pressures are too high compared with the experimen- Figs. 19 and 20, respectively. In Fig. 19, the model describes the
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 413

0 1.2
Extended UNIQUAC
Extended UNIQUAC
313.15 K - Kim et al., 2008
10 mass % MAPA 1 333.15 K - Kim et al., 2008
-4 20 mass % MAPA
353.15 K - Kim et al., 2008
27 mass % MAPA 0.8 373.15 K - Kim et al., 2008
Freezing Point / oC

333.15 K - Monteiro et al., 2011

P total / bar
-8 353.15 K - Monteiro et al., 2011
0.6

-12 0.4

0.2
-16

0
0 50 100 150 200
-20
0 1 2 3 4 5 6 MAPA Molality (Liquid Phase) / mol MAPA.(kg H2O)-1

CO2 Molality / mol CO2.(kg H2O)-1 Fig. 16. Total pressure of H2 O-MAPA system at different temperatures.

Fig. 14. Freezing point depression of H2 O-MAPA-CO2 systems at different compo-


sitions of MAPA and CO2 . 8
Experimental data from Arshad et al. (2013a). Extended UNIQUAC
7 313.15 K - Arshad et al., 2014 2M MAPA
353.15 K - Arshad et al., 2014
6
total pressures in 1 M MAPA solutions fairly well with a few high 393.15 K - Arshad et al., 2014
calculated total pressure data points for the 313.15 K and 393.15 K P total / bar 5
isotherms and a few low calculated total pressure data points for
the 353.15 K isotherm at high CO2 concentrations. Similarly, Fig. 20 4
exhibits a reasonably fair agreement between the experimental
values and model results with some deviations in all the three 3
isotherms. The estimated deviations (AARD) are 15.2% and 39%
2
respectively for the total pressure and CO2 partial pressure data.
The experimental total pressure in 5 M MAPA solutions from 1
Pinto et al., 2014a are plotted with the modeling results in Fig. 21.
The modeling results look reasonable for the data at 353.15 K and 0
393.15 K with some deviations from the data at 373.15 K in the 0 1 2 3 4 5
whole CO2 concentration range. It can be observed that the exper- CO2 Molality / mol CO2.(kg H2O)-1
imental data points of the isotherms at different temperatures are
crossing over each other, reflecting inaccuracy in the experimen- Fig. 17. Equilibrium total pressure in 2 M MAPA solutions as a function of CO2 com-
position at different temperatures.
tal data. The estimated AARD between the model results and the
Experimental data from Arshad et al. (2014a).
experimental data is 35.2%.

MAPA(aq) MAPAH+ MAPA(aq) MAPAH+


MAPAH2++ CO3-- MAPAH2++ CO3--
HCO3- MAPACOO- HCO3- MAPACOO-
CO2(aq) pH of soluon CO2(aq) pH of soluon
0.03 12 0.07 13
10 mass % MAPA 27 mass % MAPA

0.025 0.06

11 12
mole fracon, xi / (-)
mole fracon, xi / (-)

0.05
pH of soluon

0.02
pH of soluon

0.04
0.015 10 11
0.03

0.01
0.02
9 10
0.005
0.01

0 8 0 9
0 0.5 1 1.5 2 0 1 2 3 4 5 6
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 15. Model calculated speciation for 10 and 27 mass% MAPA solutions.
414 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

10 10

2M MAPA 5M MAPA
1 8
Extended UNIQUAC
353.15 K - Pinto et al., 2014a
P CO2 / bar

P total / bar
0.1 6 373.15 K - Pinto et al., 2014a
393.15 K - Pinto et al., 2014a

0.01 Extended UNIQUAC 4


313.15 K - Arshad et al., 2014
353.15 K - Arshad et al., 2014
393.15 K - Arshad et al., 2014 2
0.001
313.15 K - Pinto et al., 2014a
333.15 K - Pinto et al., 2014a
353.15 K - Pinto et al., 2014a
0
0.0001
0 2 4 6 8 10 12 14
0 1 2 3 4 5
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 18. Partial pressure of CO2 in 2 M MAPA solutions at different temperatures. Fig. 21. Equilibrium total pressure in 5 M MAPA solutions as a function of CO2
Experimental data from Arshad et al. (2014a) and Pinto et al. (2014a). composition at different temperatures.

7 0.2
1M MAPA
6
5M MAPA
Extended UNIQUAC
313.15 K - Arshad et al., 2014 0.16
5 353.15 K - Arshad et al., 2014
Extended UNIQUAC
393.15 K - Arshad et al., 2014
P CO2 / bar

313.15 K - Pinto et al., 2014a


P total / bar

4 0.12
333.15 K - Pinto et al., 2014a
3
0.08
2

1 0.04

0
0 0.5 1 1.5 2 2.5
0
CO2 Molality / mol CO2.(kg H2O)-1 0 2 4 6 8 10 12
CO2 Molality / mol CO2.(kg H2O)-1
Fig. 19. Equilibrium total pressure in 1 M MAPA solutions as a function of CO2 com-
position at different temperatures.
Fig. 22. Partial pressure of CO2 in 5 M MAPA solutions at different temperatures.
Experimental data from Arshad et al. (2014a).

10 0.014

1M MAPA 5M MAPA
0.012

1 0.01
P CO2 / bar
P CO2 / bar

Extended UNIQUAC
0.008
313.15 K - Pinto et al., 2014a
0.1
0.006

Extended UNIQUAC 0.004


0.01
313.15 K - Arshad et al., 2014
353.15 K - Arshad et al., 2014 0.002
393.15 K - Arshad et al., 2014
0.001 0
0 0.5 1 1.5 2 2.5 0 2 4 6 8 10 12
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 20. Partial pressure of CO2 in 1 M MAPA solutions at different temperatures. Fig. 23. Magnified view of partial pressure of CO2 in 5 M MAPA solutions at 313.15 K
Experimental data from Arshad et al. (2014a). up to 0.014 bar.
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 415

-100 -100
40 oC 40 oC
-80 -80

∆H abs / kJ.(mol CO2)-1

∆H abs / kJ.(mol CO2)-1


-60 -60

-40 -40

Extended UNIQUAC Extended UNIQUAC


-20 -20
2M MAPA 1M MAPA

0 0
0 1 2 3 4 5 0 0.5 1 1.5 2 2.5
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-120 -120
80 oC 80 oC
-100 -100

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1

-80 -80

-60 -60

-40 Extended UNIQUAC -40 Extended UNIQUAC


2M MAPA 1M MAPA
-20 -20

0 0
0 1 2 3 4 0 0.5 1 1.5 2
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-250 -160
120 oC 120 oC
-200
∆H abs / kJ.(mol CO2)-1
∆H abs / kJ.(mol CO2)-1

-120

-150
-80
-100
Extended UNIQUAC
-40
-50 Extended UNIQUAC 1M MAPA
2M MAPA

0 0
0 1 2 3 0 0.5 1 1.5
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 24. Differential heat of absorption of CO2 in 2 M (left side images) and 1 M (right side images) MAPA solutions at 40, 80, and 120 ◦ C.
Experimental data from Arshad et al. (2013b).

Pinto et al. (2014a) reported the solubility of CO2 in 5 M MAPA data, and showed large deviations at low CO2 concentrations for
solutions at different temperatures. These experimental data are the 120 ◦ C isotherm. Similar kind of results can also be seen for
plotted with the calculated curves in Fig. 22. A good agreement 1 M MAPA solutions with slightly large deviations compared to
between the modeling results and the experimental data can be the 2 M MAPA results at respective temperatures. The estimated
seen at low pressures for both the isotherms at 313.15 K 333.15 K. deviations (AARD) between the experimental data and the mod-
However, some deviations can also be observed at high pressure. A eling results are 11.7% and 15.6% respectively for the 2 M and 1 M
magnified view of the results at low CO2 partial pressures of up to MAPA solutions with major contributions from the data at high
0.014 bar for the 313.15 K isotherm are shown in Fig. 23 which also temperatures.
show some scatter in the experimental data. The reported experimental data for differential heats of absorp-
Differential heats of absorption of CO2 in the aqueous MAPA tion of CO2 in 8 mass% and 3 mass% MAPA solutions from Kim (2009)
solutions were the only thermal property data available for ther- and the modeling results are presented in Fig. 25. The model very
modynamic modeling of the H2 O-MAPA-CO2 system. The modeling well describes the 8 mass% MAPA data at 40 ◦ C and slightly lower
results for the 2 M and 1 M MAPA solutions along with the experi- calculated results for the 3 mass% MAPA data at the same temper-
mental data from Arshad et al. (2013b) are given in Fig. 24 (left side ature. However, there is a large scatter in the experimental data
images for 2 M MAPA and right side images for 1 M MAPA). The at 80 ◦ C for both the MAPA systems (8 mass% and 3 mass%) and
model calculated the differential heat of absorption of CO2 in 2 M poorly represented by the model. The estimated deviations (AARD)
MAPA solutions very well at 40 ◦ C. However, the model calculated are 17.9% and 31.6% respectively for the 8 mass% and 3 mass%
results are slightly lower at 80 ◦ C compared to the experimental MAPA solutions with largest contribution coming from the scat-
416 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

-100 -100

40 oC 40 oC
-80 -80

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1
-60 -60

-40 -40

Extended UNIQUAC Extended UNIQUAC


-20 -20
8 mass % MAPA 3 mass % MAPA

0 0
0 0.4 0.8 1.2 1.6 2 0 0.2 0.4 0.6 0.8 1
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-160 -200
80 oC 80 oC
-160
-120
∆H abs / kJ.(mol CO2)-1

∆H abs / kJ.(mol CO2)-1


-120
-80
-80

-40 Extended UNIQUAC


-40 Extended UNIQUAC
8 mass % MAPA
3 mass % MAPA
0 0
0 0.3 0.6 0.9 1.2 1.5 1.8 0 0.2 0.4 0.6 0.8
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 25. Differential heat of absorption of CO2 in 8 mass% (left side images) and 3 mass% (right side images) MAPA solutions at 40 and 80 ◦ C.
Experimental data from Kim (2009) were not used in the parameter estimation.

tered experimental data at 80 ◦ C. It should be noted that these data 0


were not included in the parameter estimation.
It can be noticed that the modeling of the H2 O-MAPA-CO2 sys-
tem showed high deviations in the VLE and heat of absorption -5
results. One possible reason may be the large scatter and low accu-
Freezing Point / oC

racy of some of the experimental data (e.g., VLE data of 5 M MAPA


system). Excluding those experimental data from the parameter -10
estimation was not an option because these were the only available
data for modeling of the system. Other reason may be lacking of one
-15 Extended UNIQUAC
of the MAPA species in the modeling. This species may the proto-
DEEA/MAPA 5:1
nated carbamate of MAPA (+ HMAPACOO− ). As mentioned earlier DEEA/MAPA 3:1
that the complexity of the MAPA system is similar to that of the DEEA/MAPA 1:1
-20
aqueous AEEA as reported by Ma’mun et al. (2006). They reported DEEA/MAPA 1:3
the formation of protonated carbamate species of AEEA at high DEEA/MAPA 1:5
CO2 loadings. Due to unavailability of the standard state thermo- -25
dynamic property data, it was difficult to include such a species in 0 1 2 3 4 5 6 7 8
the parameter estimation. However, attempts were made in this Amine Molality / mol Amine.(kg H2O)-1
work to model the MAPA system by including the protonated pri-
mary carbamate of MAPA with guess values for the standard state Fig. 26. Freezing point depression in H2 O-DEEA-MAPA system.
Gibbs energy and enthalpy of formation. These values were altered Experimental data from Arshad et al. (2013a).

during the parameter estimation to fit to the experimental data.


But the results did not show any improvements in the model. This
An overview of all the results, for all the systems studied in this
might be due to inappropriate values for the standard state Gibbs
work and for all the data types, has been presented in Tables A1–A6
energy and enthalpy of formation for the protonated primary carba-
in Appendix A given at the end of the paper.
mate of MAPA. Therefore, it was decided to exclude + HMAPACOO−
from the parameter estimation. It should be noted here that the
standard state thermodynamic properties for the MAPA primary 8.3. H2 O-DEEA-MAPA-CO2 system (liquid–liquid phase change
carbamate (MAPACOO− ) were also not available in the literature system)
and were guessed initially and then these values were fitted to
the experimental data. In case of MAPA primary carbamate, the The parameter estimation for the blended aqueous DEEA-MAPA
results showed significant improvements and thus retained in the systems was started by calculating the freezing points and VLE
model. (total pressure) for the H2 O-DEEA-MAPA solutions. The calculated
freezing point data are illustrated in Fig. 26 together with the exper-
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 417

7
Extended UNIQUAC
1 5M DEEA + 2M MAPA
6 313.15 K - Arshad et al., 2014
353.15 K - Arshad et al., 2014

5 393.15 K - Arshad et al., 2014


P total / bar

P CO2 / bar
4

0.1
3

2
Extended UNIQUAC 323.15 K - Hartono et al., 2013
333.15 K - Hartono et al., 2013 353.15 K - Hartono et al., 2013
1
373.15 K - Hartono et al., 2013 383.15 K - Hartono et al., 2013

0.01
0
0 10 20 30 40 50 60 70
0 5 10 15 20 25 30 35
Amine Molality (Liquid Phase) / mol Amine.(kg H2O)-1
CO2 Molality / mol CO2.(kg H2O)-1
Fig. 27. Total pressure of H2 O-DEEA-MAPA system at different temperatures.
Fig. 29. Partial pressure CO2 in 5 M DEEA + 2 M MAPA solutions as a function of CO2
composition at three different temperatures.
Experimental data from Arshad et al. (2014a).
8
Extended UNIQUAC
5M DEEA + 2M MAPA
7 313.15 K - Arshad et al., 2014
8
353.15 K - Arshad et al., 2014
6 5M DEEA + 1M MAPA Extended UNIQUAC
393.15 K - Arshad et al., 2014 7 313.15 K - Arshad et al., 2014
5 353.15 K - Arshad et al., 2014
P total / bar

6
393.15 K - Arshad et al., 2014
4
5
P total / bar

3
4

2 3

1 2

0 1
0 5 10 15 20 25 30 35
CO2 Molality / mol CO2.(kg H2O)-1 0
0 5 10 15 20
Fig. 28. Equilibrium total pressure in 5 M DEEA + 2 M MAPA solutions as a function CO2 Molality / mol CO2.(kg H2O)-1
of CO2 composition at three different temperatures.
Experimental data from Arshad et al. (2014a). Fig. 30. Equilibrium total pressure in 5 M DEEA + 1 M MAPA solutions as a function
of CO2 composition at three different temperatures.
Experimental data from Arshad et al. (2014a).

imental data from Arshad et al. (2013a). The freezing point data are
plotted as a function of total amine concentration in the solutions
for the 5:1, 3:1, 1:1, 1:3, and 1:5 molar ratios of DEEA/MAPA. A be seen in Fig. 29. However, the model very well calculated the CO2
good agreement between the modeling results and the experimen- partial pressures at 313.15 K.
tal data can be observed with an estimated AARD of 5.2%. Fig. 27 Similarly, for the 5D1M system, the total pressure results are
presents a comparison between the calculated and experimental shown in Fig. 30 and the CO2 partial pressure results are illustrated
(Hartono et al., 2013) total pressure data as a function of amine in Fig. 31 along with the experimental data at three temperatures.
molality in the liquid phase at different temperatures. The model The model describes the total pressure data very well in the whole
describes the total pressure data fairly well with an estimated AARD CO2 concentration range for all the three isotherms (Fig. 30) with an
of 13.6%. estimated AARD of 8.2%. Similar nice results can also be observed in
The experimental VLE data (both total pressure and CO2 partial Fig. 31 for the CO2 partial pressure data with a few high calculated
pressure) for the 5 M DEEA + 2 M MAPA (5D2 M) and 5 M DEEA + 1 M data points for the 393.15 K isotherm at high CO2 concentrations.
MAPA (5D1 M) systems were reported by Arshad et al. (2014a). The The estimated AARD between the calculated and the experimental
results of 5D2 M system are presented in Fig. 28 for the total pres- data set is 16.4%, twice the AARD of total pressure data.
sure and Fig. 29 for the CO2 partial pressure. A fairly good agreement The extended UNIQUAC model calculated speciation (from VLE
between the calculated and experimental total pressures can be data) in 5D2M and 5D1M systems at 40 ◦ C are presented in Fig. 32.
seen in Fig. 28 (AARD of 11.2%) with some deviations in the results The mole fraction of different species (left ordinate) together with
at 353.15 K and 393.15 K. It can be noticed that these deviations the pH of the solution (right ordinate) are plotted as a function
at 353.15 K and 393.15 K are similar to that of 2 M MAPA results of CO2 concentration. It can be noticed that MAPA (fast reaction
(Fig. 17) which were probably reproduced also in the 5D2 M sys- kinetics) reacts first with CO2 to form mainly MAPA carbamate and
tem. These deviations are also reflected in the calculated CO2 partial di-protonated MAPA species in both cases (5D2M and 5D1M). Once
pressures for the same isotherm (at 353.15 K and 393.15 K) as can MAPA is fully consumed, DEEA takes over to form protonated DEEA
418 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

7 ous DEEA is consumed and the amount of CO2 physically absorbed


5M DEEA + 1M MAPA Extended UNIQUAC in the reaction mixture is also very low.
6 313.15 K - Arshad et al., 2014 As mentioned earlier, when CO2 is dissolved in an aqueous solu-
353.15 K - Arshad et al., 2014 tion of DEEA, it promotes hydration of CO2 to produce bicarbonate
5 393.15 K - Arshad et al., 2014 and protonated DEEA, according to Donaldson and Nguyen (1980)
(Eq. (13)). The 5D2M system contains 82.6 mass% total amine (63.5%
P CO2 / bar

4 DEEA and 19.1% MAPA) and only 17.4 mass% H2 O. When aqueous
DEEA starts reacting after aqueous MAPA, there is a scarcity of H2 O
3 in the reaction mixture, thus low hydration of CO2 . This possibly
explains the large amount of unreacted aqueous DEEA, decreased
2 slope of the bicarbonate ions curve, relatively less amount of proto-
nated DEEA, and considerable amount of physically absorbed CO2
1 in the reaction mixture for 5D2M system. On the other hand, 5D1M
system contains 71.3 mass% total amine (62% DEEA and 9.3% MAPA)
0 and 28.7 mass% H2 O. In 5D1M system, the amount of MAPA is
0 5 10 15 20
halved and the amount of water is 69% more compared to 5D2M
CO2 Molality / mol CO2.(kg H2O)-1 system. This means that more water is available for hydration of
CO2 when aqueous DEEA starts reacting after aqueous MAPA in
Fig. 31. Partial pressure CO2 in 5 M DEEA + 1 M MAPA solutions as a function of CO2
composition at three different temperatures. 5D1M system compared to 5D2M system. This can be observed
Experimental data from Arshad et al. (2014a). in 5D1M system with a little amount of unreacted aqueous DEEA,
large amount of bicarbonate and protonated DEEA species, and a
negligible amount of physically absorbed CO2 .
Fig. 33 illustrates the calculated differential heats of absorption
and bicarbonate species. The formation of mono-protonated MAPA
of CO2 in 5D2M (left side images) and 5D1M (right side images)
and carbonate species is negligible in both cases.
systems together with the experimental data from Arshad et al.
In the 5D2M case, the formation of MAPA carbamate first
(2013b) for the temperature range of 40–120 ◦ C. The experimen-
increases to a maximum (at this point aqueous MAPA is almost
tal and the calculated values present a good agreement for the
fully consumed) and then starts decreasing with an increase in CO2
5D2M system in the whole range of temperature and CO2 compo-
concentration. Formation of bicarbonate (quite steep) and proto-
sition with an estimated AARD of 6.1%. However, the 5D1M system
nated DEEA species can be observed when DEEA starts reacting. It
shows slightly higher deviations in all the three isotherms com-
can also be noticed that a large amount of unreacted aqueous DEEA
pared to that of the 5D2M systems. The estimated AARD for the
is present and a considerable amount of CO2 is physically absorbed
5D1M system is 8.3%.
in the reaction mixture. In the 5D1M case, on the other hand, the
The experimental differential heat of absorption data available
formation of MAPA carbamate first increases to a maximum and
from Kim (2009) for the aqueous blends of DEEA-MAPA are also
then become almost constant (slightly decreases, then increases a
compared with the calculated values as shown in Fig. 34 for the
little and decreases again). A steep increase in formation of both
37 mass% DEEA + 3 mass% MAPA (left side images) and 32 mass%
bicarbonate and protonated DEEA species can be observed when
DEEA + 8 mass% MAPA (right side images) systems at 40, 80, and
DEEA starts reacting. It can also be noticed that most of the aque-

DEEA(aq) DEEAH+ DEEA(aq) DEEAH+


MAPA(aq) MAPAH+ MAPA(aq) MAPAH+
MAPAH2++ CO3-- MAPAH2++ CO3--
HCO3- MAPACOO- HCO3- MAPACOO-
CO2(aq) pH of soluon CO2(aq) pH of soluon
0.35 12 0.25 12
5M DEEA + 2M MAPA (40 oC) 5M DEEA + 1M MAPA (40 oC)

0.3
0.2
11 11
0.25
mole fracon, xi / (-)

mole fracon, xi / (-)

pH of soluon
pH of soluon

0.15
0.2
10 10
0.15
0.1

0.1
9 9
0.05
0.05

0 8 0 8
0 3 6 9 12 15 18 21 24 27 30 33 0 2 4 6 8 10 12 14 16 18 20
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 32. Model calculated speciation for 5 M DEEA + 2 M MAPA (left) and 5 M DEEA + 1 M MAPA (right) solutions at 40 ◦ C.
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 419

-120 -120

40 oC
-100 -100 40 oC

∆H abs / kJ.(mol CO2)-1

∆H abs / kJ.(mol CO2)-1


-80 -80

-60 -60

-40 -40
Extended UNIQUAC Extended UNIQUAC
-20 5M DEEA + 2M MAPA -20 5M DEEA + 1M MAPA

0 0
0 10 20 30 40 0 5 10 15 20 25
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-120 -120
80 oC 80 oC
-100 -100

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1

-80 -80

-60 -60

-40 -40
Extended UNIQUAC Extended UNIQUAC
-20 5M DEEA + 2M MAPA -20 5M DEEA + 1M MAPA

0 0
0 5 10 15 20 25 0 3 6 9 12
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-120 -120
120 oC 120 oC
-100 -100
∆H abs / kJ.(mol CO2)-1
∆H abs / kJ.(mol CO2)-1

-80 -80

-60 -60

-40 Extended UNIQUAC -40 Extended UNIQUAC


5M DEEA + 2M MAPA 5M DEEA + 1M MAPA
-20 -20

0 0
0 3 6 9 12 15 0 1 2 3 4 5 6
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 33. Differential heat of absorption of CO2 in 5 M DEEA + 2 M MAPA (left side images) and 5 M DEEA + 1 M MAPA (right side images) solutions at 40, 80, and 120 ◦ C.
Experimental data from Arshad et al. (2013b).

120 ◦ C. The results look satisfactory for both the systems at 40 and curves and tie lines) are plotted in comparison with the experimen-
80 ◦ C. However, large deviations can be observed for both the sys- tal data. Fig. 35 shows the calculated values at 40 ◦ C for the 6 molal
tems at 120 ◦ C. The estimated AARD between the model results and MAPA solutions with a constant CO2 concentration of 4.6 molal
the experimental data is 18.1% for the 37 mass% DEEA + 3 mass% and varying DEEA concentrations. To keep the calculations simple
MAPA system and 11.2% for the 32 mass% DEEA + 8 mass% MAPA and better visualization of the graphs, the model calculations were
system. It should be noted that these experimental data were not based on slightly different concentration conditions compared to
included in the parameter estimation. the experimental data i.e., the MAPA and CO2 concentrations were
As mentioned earlier, one of the main features of 5 M DEEA + 2 M not constant in the experimental data but an almost averaged value
MAPA mixture is that it gives liquid–liquid split upon CO2 absorp- was used for the model calculations. This simplification allowed
tion. The upper phase, lean in CO2 and rich in DEEA, is sent to the the model to calculate the smooth binodal curves and the uniform
absorber without regeneration and only the CO2 rich lower phase tie lines which, otherwise, was not possible. Similarly, the model
(also rich in MAPA) is regenerated in the stripper. It is essential that results at 60 ◦ C for the 7.5 molal MAPA solutions with a constant
the model can predict the LLE in the mixed aqueous DEEA-MAPA CO2 concentration of 4.5 molal are illustrated in Fig. 36 and for the
system. Only 32 LLE data points were available for the parameter 7.5 molal MAPA solutions with a constant CO2 concentration of 3.75
estimation from Pinto et al. (2014b) at three different tempera- at 80 ◦ C in Fig. 37 together with the experimental data scattering
tures. Figs. 35–37 illustrate the LLE results in H2 O-DEEA-MAPA-CO2 around the binodal curves.
solutions at 40 ◦ C, 60 ◦ C, and 80 ◦ C. The calculated results (binodal
420
-100 -100

40 oC 40 oC
-80 -80

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1
-60 -60

-40 -40

-20 Extended UNIQUAC -20 Extended UNIQUAC


37 mass % DEEA + 3 mass % MAPA 32 mass % DEEA + 8 mass % MAPA

M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424
0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-120 -100
80 oC
80 oC
-100
-80

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1

-80
-60
-60
-40
-40

Extended UNIQUAC -20 Extended UNIQUAC


-20
37 mass % DEEA + 3 mass % MAPA 32 mass % DEEA + 8 mass % MAPA
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

-200 -160

120 oC 120 oC
-160

∆H abs / kJ.(mol CO2)-1


∆H abs / kJ.(mol CO2)-1

-120

-120
-80
-80

-40
-40 Extended UNIQUAC Extended UNIQUAC
37 mass % DEEA + 3 mass % MAPA 32 mass % DEEA + 8 mass % MAPA
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2 2.5 3
CO2 Molality / mol CO2.(kg H2O)-1 CO2 Molality / mol CO2.(kg H2O)-1

Fig. 34. Differential heat of absorption of CO2 in 37 mass% DEEA + 3 mass% MAPA (left side images) and 32 mass% DEEA + 8 mass% MAPA (right side images) solutions at 40, 80, and 120 ◦ C.
Experimental data from Kim (2009) were not used in the parameter estimation.
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 421

25 9. Conclusions
oC Upper Phase
40
Lower Phase The Extended UNIQUAC framework has been implemented
MAPA Molality / mol.(kg H2O)-1

20 Binodal curve, Extended UNIQUAC in this work to describe the thermodynamics of de-mixing
Tie lines, Extended UNIQUAC liquid–liquid aqueous DEEA-MAPA solvent systems for CO2 cap-
ture. The two sub-systems, H2 O-DEEA-CO2 and H2 O-MAPA-CO2 ,
15 were modeled first followed by the H2 O-DEEA-MAPA-CO2 system
which gives liquid–liquid phase split. Different types of experi-
mental equilibrium and thermal data were used for the parameter
10 estimation. Ninety four model parameters and six thermodynamic
properties were fitted to approximately 1500 experimental data
consisting of pure amine vapor pressure, vapor-liquid equilib-
5
rium, solid-liquid equilibrium, liquid–liquid equilibrium, excess
enthalpy, and heat of absorption of CO2 in aqueous amine solutions.
Out of 94 fitted model parameters, volume (r) and surface area (q)
0
parameters have 6 parameters each and the binary parameters,
0 20 40 60 80 100 120 140
u0ij and uTij , have 41 parameters each for calculating the interac-
DEEA Molality / mol.(kg H2O)-1
tion energy parameters, uij = u0ij + uTij (T − 298.15). The six fitted
Fig. 35. Modeling results of liquid–liquid equilibrium in H2 O-DEEA-MAPA-CO2 thermodynamic properties are standard state Gibbs energy of for-
solutions at 40 ◦ C with experimental data from Pinto et al. (2014b). Calculated val- mation and standard state enthalpy of formation for the three
ues are for 6 molal MAPA solutions with a constant CO2 concentration of 4.6 molal
species DEEA (l), MAPA (l), and MAPACOO− . The model can accu-
and varying DEEA molality.
rately represent the equilibrium and thermal data for the studied
25
systems with a single unique set of parameters. The model param-
eters are valid in the temperature range from −25 to 200 ◦ C, CO2
Upper Phase
60 oC Lower Phase
MAPA Molality / mol.(kg H2O)-1

partial pressure from 0 to 945 kPa, and concentration of DEEA,


20 Binodal curve, Extended UNIQUAC
MAPA, and CO2 up to 131, 23 and 33 mol (kg H2 O)−1 respec-
Tie lines, Extended UNIQUAC
tively. The model calculated speciation are also presented for the
studied systems. The model developed in this work can be used
15
for process simulation of CO2 capture with aqueous blends of
DEEA/MAPA.
10
Acknowledgements

5 Financial support from European Commission’s 7th Framework


Program (Grant Agreement No. 241393) through the iCap project
is greatly acknowledged. Special thanks to Prof. Erling H. Stenby
0 (DTU – Chemistry, Denmark) and Assoc. Prof. Philip L. Fosbøl (DTU
0 20 40 60 80 100 120 140 – Chemical Engineering, Denmark) for their efforts in establishing
the project funding and making DTU a part of the iCap project.
DEEA Molality / mol.(kg H2O)-1

Fig. 36. Modeling results of liquid–liquid equilibrium in H2 O-DEEA-MAPA-CO2 Appendix A.


solutions at 60 ◦ C with experimental data from Pinto et al. (2014b). Calculated val-
ues are for 7.5 molal MAPA solutions with a constant CO2 concentration of 4.5 molal To check the overall performance of the model, an overview of
and varying DEEA molality. all the results is given in Tables A1–A6 . The results are presented
20 here according to data type for all the studied systems.
Upper Phase
80 oC Table A1
Lower Phase
MAPA Molality / mol.(kg H2O)-1

16 Binodal curve, Extended UNIQUAC Results for the vapor pressure of pure amines.
Tie lines, Extended UNIQUAC T/ ◦ C Source N AARD%

DEEA
12
59.35–195.63 Steele et al. (2002) 22 6.5
5.05–45.15 Kapteina et al. (2005) 31 20.3
60.05–176.35 Klepáčová et al. (2011) 13 5
8 64.91–154.98 Hartono et al. (2013) 38 5
MAPA
53.66–139.35 Kim et al. (2008) 11 2.1
3.05–30.05 Verevkin and Chernyak (2012) 12 15.2
4 55.99–139.02 Hartono et al. (2013) 18 0.9

0
0 20 40 60 80 100 Table A2
Results for the excess enthalpy of DEEA.
DEEA Molality / mol.(kg H2O)-1
T/ ◦ C mole% DEEA Source N AARD%
Fig. 37. Modeling results of liquid–liquid equilibrium in H2 O-DEEA-MAPA-CO2
25 4.72–90.11 Mathonat et al. (1997) 18 7.2
solutions at 80 ◦ C with experimental data from Pinto et al. (2014b). Calculated values
are for 7.5 molal MAPA solutions with a constant CO2 concentration of 3.75 molal
and varying DEEA molality.
422 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

Table A3
Results for the freezing point depression of binary and ternary water-amine-CO2 systems.

T/ ◦ C mass% DEEA mass% MAPA mass% CO2 Source N AARD%

H2 O-DEEA
−0.44 to −24.6 2.54–55.01 – – Arshad et al. (2013a) 23 4.2
H2 O-MAPA
−0.62 to −23.87 – 2.59–32.5 – Arshad et al. (2013a) 16 3.4
H2 O-DEEA-MAPA
−0.52 to −21.24 0.54–39.18 0.33–25.68 – Arshad et al. (2013a) 44 5.2
H2 O-DEEA-CO2
−2.37 to −18.03 12, 20, 30, 33 – 0.23–11.0 Arshad et al. (2013a) 40 3.6
H2 O-MAPA-CO2
−2.77 to −15.42 – 10, 20, 27 0.36–15.12 Arshad et al. (2013a) 32 3.1

Table A4
Results for the total pressure of binary and ternary water-amine systems.

mass% DEEA mass% MAPA T/ ◦ C P/ bar Source N AARD%

H2 O-DEEA
0.056 − 70.94 – 50–95 0.094–0.87 Hartono et al. (2013) 44 11
H2 O-MAPA
– 1.21–80.05 40–100 0.06–1.004 Kim et al. (2008) 62 4.4
– 10.29–93.66 60–80 0.06–0.461 Monteiro et al. (2011) 48 3.7
H2 O-DEEA-MAPA
18.92 − 84.07 19.23–68.1 50–110 0.071–1.11 Hartono et al. (2013) 38 13.7

Table A5
Results for the VLE (total pressure and CO2 partial pressure) of ternary and quaternary water-amine-CO2 systems.

mass% DEEA mass% MAPA CO2 / mol(kg H2 O)−1 T/ ◦ C Source P/ bar N AARD%

H2 O-DEEA-CO2
61.09 (5 M) – 0.204–13.902 40–120 Arshad et al. (2014a) PTotal = 0.123–5.887 91 10.4
PCO2 = 0.006– 5.771 91 22.4
61.09 (5 M) – 0.066–9.013 40–120 Monteiro et al. (2013) PTotal = 1.76–10.353 26 7.6
PCO2 = 0.00063–0.1923 34 19.1
23.66 (2 M) – 0.054–2.69 40–120 Monteiro et al. (2013) PTotal = 1.157–9.771 26 10.1
PCO2 = 0.00029–0.17342 33 14.1
H2 O-MAPA-CO2
– 8.9 (1 M) 0.383–2.078 40–120 Arshad et al. (2014a) PTotal = 0.097–5.978 51 15.2
PCO2 = 0.003–5.345 51 39
– 17.88 (2 M) 0.979–3.912 40–120 Arshad et al. (2014a) PTotal = 0.101–5.582 45 16.1
PCO2 = 0.003–4.577 45 55.1
– 17.88 (2 M) 0.911–2.867 40–80 Pinto et al. (2014a) PCO2 = 0.00023–0.19977 59 37.2
– 45.18 (5 M) 1.169–10.744 80–120 Pinto et al. (2014a) PTotal = 1.573–9.192 24 35.2
H2 O-DEEA-MAPA-CO2
63.533 (5 M) 19.116 (2 M) 1.16–32.496 40–120 Arshad et al. (2014a) PTotal = 0.112–5.892 62 11.2
PCO2 = 0.009–5.15 62 29.3
62.025 (5 M) 9.331 (1 M) 0.663–19.984 40–120 Arshad et al. (2014a) PTotal = 0.096–5.829 68 8.2
PCO2 = 0.004–5.219 68 16.4

Table A6
Results for the differential heat of absorption of CO2 in ternary and quaternary water-amine-CO2 systems.

mass% DEEA mass% MAPA CO2 / mol(kg H2 O)−1 T/ ◦ C Source N AARD%

H2 O-DEEA-CO2
61.09 (5 M) – 0.204–13.902 40–80 Arshad et al. (2013b) 72 11.6
32 – 0.317–4.963 40–80 Kim (2009) 38 21.5
37 – 0.326–5.708 40–80 Kim (2009) 49 17.9
H2 O-MAPA-CO2
– 8.9 (1 M) 0.383–2.078 40–120 Arshad et al. (2013b) 66 15.6
– 17.88 (2 M) 0.979–3.912 40–120 Arshad et al. (2013b) 69 11.7
– 3 0.069–0.881 40–80 Kim (2009) 30 31.6
– 8 0.046–1.94 40–80 Kim (2009) 41 17.9
H2 O-DEEA-MAPA-CO2
63.533 (5 M) 19.116 (2 M) 1.16–32.496 40–120 Arshad et al. (2013b) 82 6.1
62.025 (5 M) 9.331 (1 M) 0.663–19.984 40–120 Arshad et al. (2013b) 77 8.3
37 3 0.257–6.027 40–120 Kim (2009) 57 18.1
32 8 0.34–5.527 40–120 Kim (2009) 41 11.2

References Aronu, U.E., Hartono, A., Svendsen, H.F., 2011. Kinetics of carbon dioxide absorption
into aqueous amine amino acid salt: 3-(methylamino)propylamine/sarcosine
Abrams, D.S., Prausnitz, J.M., 1974. Statistical thermodynamics of liquid mixtures: solution. Chem. Eng. Sci. 66, 6109–6119.
a new expression for the excess Gibbs energy of partly or completely miscible Arshad, M.W., Fosbøl, P.L., von Solms, N., Thomsen, K., 2013a. Freezing point
systems. AIChE J. 21 (1), 116–128. depressions of phase change CO2 solvents. J. Chem. Eng. Data 58 (7),
1918–1926.
M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424 423

Arshad, M.W., Fosbøl, P.L., von Solms, N., Svendsen, H.F., Thomsen, K., 2013b. Heat Liebenthal, U., Pinto, D.D.D., Monteiro, J.G.M.-S., Svendsen, H.F., Kather, A., 2013.
of absorption of CO2 in phase change solvents: 2-(diethylamino)ethanol and Overall process analysis and optimisation for CO2 capture from coal fired
3-(methylamino)propylamine. J. Chem. Eng. Data 58 (7), 1974–1988. power plants based on phase change solvents forming two liquid phases.
Arshad, M.W., von Solms, N., Svendsen, H.F., Thomsen, K., 2013c. Heat of Energy Procedia 37, 1844–1854.
absorption of CO2 in aqueous solutions of DEEA, MAPA and their mixture. Littel, R.J., Bos, M., Knoop, G.J., 1990. Dissociation constants of some alkanolamines
Energy Procedia 37, 1532–1542. at 293, 303, 318, and 333 K. J. Chem. Eng. Data 35, 276–277.
Arshad, M.W., Svendsen, H.F., Fosbøl, P.L., von Solms, N., Thomsen, K., 2014a. Ma’mun, S., Jakobsen, J.P., Svendsen, H.F., 2006. Experimental and modeling study
Equilibrium total pressure and CO2 solubility in binary and ternary aqueous of the solubility of carbon dioxide in aqueous 30 mass%
solutions of 2-(diethylamino)ethanol (DEEA) and 2-((2-aminoethyl)amino)ethanol solution. Ind. Eng. Chem. Res. 45 (8),
3-(methylamino)propylamine (MAPA). J. Chem. Eng. Data 59 (3), 764–774. 2505–2512.
Arshad, M.W., Fosbøl, P.L., von Solms, N., Svendsen, H.F., Thomsen, K., 2014b. Maham, Y., Helper, L.G., Mather, A.E., Hakin, A.W., Marriott, R.A., 1997. Molar heat
Equilibrium solubility of CO2 in alkanolamines. Energy Procedia 51, 217–223. capacities of alkanolamines from 299.1 to 397.8 K. Group additivity and
Arshad, M.W., 2014. Measuring and Thermodynamic Modeling of De-mixing CO2 molecular connectivity analyses. J. Chem. Soc. Faraday Trans. 93 (9),
Capture Systems Ph.D. Thesis. Technical University of Denmark, Kongens 1747–1750.
Lyngby, Denmark. Marrero, J., Gani, R., 2001. Group-contribution based estimation of pure
CAPEC Software, 2014. ProPred (pure component Property Prediction) tool in ICAS component properties. Fluid Phase Equilib. 183/184, 183–208.
(Integrated Computer Aided System). http://www.capec.kt.dtu.dk/Software/ Mathonat, C., Maham, Y., Mather, A.E., Hepler, L.G., 1997. Excess molar enthalpies
CAPEC-Software. of (water + monoalkanolamine) mixtures at 298.15 K and 308: 15 K. J. Chem.
Darde, V., van Well, W.J.M., Stenby, E.H., Thomsen, K., 2010. Modeling of carbon Eng. Data 42, 993–995.
dioxide absorption by aqueous ammonia solutions using the Extended Maurer, G., Prausnitz, J.M., 1978. On the derivation and extension of the UNIQUAC
UNIQUAC model. Ind. Eng. Chem. Res. 49, 12663–12674. equation. Fluid Phase Equilib. 2, 91–99.
Darde, V., Maribo-Mogensen, B., van Well, W.J.M., Stenby, E.H., Thomsen, K., 2012. Monteiro, J.G.M.-S., Pinto, D.D.D., Hartono, A., Svendsen, H.F., 2011. Unpublished
Process simulation of CO2 capture with aqueous ammonia using the Extended results of VLE of aqueous solutions of MAPA.
UNIQUAC model. Int. J. Greenh. Gas Control 10, 74–87. Monteiro, J.G.M.-S., Pinto, D.D.D., Zaidy, S.A.H., Hartono, A., Svendsen, H.F., 2013.
Donaldson, T.L., Nguyen, Y.N., 1980. Carbon dioxide reaction kinetics and transport VLE data and modelling of aqueous N,N-diethyethanolamine (DEEA) solutions.
in aqueous amine membranes. Ind. Eng. Chem. Fundam. 19, 260–266. Int. J. Greenh. Gas Control 19, 432–440.
Faramarzi, L., Kontogeorgis, G.M., Thomsen, K., Stenby, E.H., 2009. Extended Murray Jr., R.C., Cobble, J.W., 1980. Chemical equilibria in aqueous systems at high
UNIQUAC model for thermodynamic modeling of CO2 absorption in aqueous temperatures. Annual Meeting − International Water Conference, 295–310.
alkanolamine solutions. Fluid Phase Equilib. 282, 121–132. NIST, 1990. Chemical Thermodynamics Database Vesion 1.1. U.S. Department of
Fletcher, R., 1971. A Modified Marquardt subroutine for non-linear least squares, Commerce, National Institute of Standards and Technology, Gaithersburg
Harwell Report, http://www.hsl.rl.ac.uk/ Direct link: http://www.hsl.rl.ac.uk/ Maryland, pp. 20899.
archive/specs/va27.pdf. Nelder, Mead, 1965. StatLib-Applied Statistics Algorithms, Nelder-Mead simplex
Fosbøl, P.L., Thomsen, K., Stenby, E.H., 2009. Modeling of the mixed solvent minimization routine (AS 47), Link No. 47 at http://lib.stat.cmu.edu/apstat/.
electrolyte system CO2 -Na2 CO3 -NaHCO3 -monoethylene glycol-water. Ind. Pinto, D.D.D., Bruder, P., Monteiro, J.G.M-S., Jens, C., Foss, C., Hartono, A., Svendsen,
Eng. Chem. Res. 48 (4), 4565–4578. H.F., 2014. Unpublished VLE data and modelling of aqueous
Fosbøl, P.L., Randi, N., Arshad, M.W., Tecle, Z., Thomsen, K., 2011. Aqueous N-methyl-1,3-diaminopropane (MAPA) solutions.
solubility of piperazine and 2-amino-2-methyl-1-propanol plus their mixtures Pinto, D.D.D., Zaidy, S.A.H., Hartono, A., Svendsen, H.F., 2014b. Evaluation of a
using an improved freezing-point depression method. J. Chem. Eng. Data 56, phase change solvent for CO2 capture: absorption and desorption tests. Int. J.
5088–5093. Greenh. Gas Control 28, 318–327.
Gaspar, J., Arshad, M.W., Blaker, E.A., Langseth, B., Hansen, T., Thomsen, K., von Rayer, A.V., Henni, A., Tontiwachwuthikul, P., 2012. Molar heat capacities of
Solms, N., Fosbøl, P.L., 2014. A low energy aqueous ammonia CO2 capture solvents used in CO2 capture: a group additivity and molecular connectivity
process. Energy Procedia 63, 614–623. analysis. Can. J. Chem. Eng. 90, 367–376.
Gassnona Annual Report, 2012. http://www.gassnova.no/en/Documents/ Raynal, L., Bouillon, P.-A., Gomez, A., Broutin, P., 2011a. From MEA to demixing
AnnualReport2012.pdf. solvents and future steps, a roadmap for lowering the cost of post-combustion
Hamborg, E.S., Versteeg, G.F., 2009. Dissociation constants and thermodynamic carbon capture. Chem. Eng. J. 171, 742–752.
properties of amines and alkanolamines from (293 to 353) K. J. Chem. Eng. Raynal, L., Pascal, A., Bouillon, P.-A., Gomez, A., le Febvre de Nailly, M., Jacquin, M.,
Data 54, 1318–1328. Kittel, J., di Lella, A., Mougin, P., Trapy, J., 2011b. The DMXTM process: an
Hartono, A., Saleem, F., Arshad, M.W., Usman, M., Svendsen, H.F., 2013. Binary and original solution for lowering the cost of post-combustion carbon capture.
ternary VLE of the 2-(diethylamino)-ethanol Energy Procedia 4, 779–786.
(DEEA)/3-(methylamino)-propylamine (MAPA)/water system. Chem. Eng. Sci. Richner, G., Puxty, G., 2012. Assessing the chemical speciation during CO2
101, 401–411. absorption by aqueous amines using in situ FTIR. Ind. Eng. Chem. Res. 51,
Helgeson, H.C., Kirkham, D.H., Flowers, G.C., 1981. Theoretical prediction of the 14317–14324.
thermodynamic behavior of aqueous electrolytes at high pressures and Sadegh, N., Stenby, E.H., Thomsen, K., 2015a. Thermodynamic modeling of CO2
temperatures: IV. Calculation of activity coefficients, osmotic coefficients, and absorption in aqueous N-methyldiethanolamine using Extended UNIQUAC
apparent molal and standard and relative partial molal properties to 600 ◦ C model. Fuel 144, 295–306.
and 5 kbar. Am. J. Sci. 281, 1249–1516. Sadegh, N., Stenby, E.H., Thomsen, K., 2015b. Thermodynamic modeling of
Joback, K.G., Reid, R.C., 1987. Estimation of pure-component properties from hydrogen sulfide absorption by aqueous N-methyldiethanolamine using the
group-contributions. Chem. Eng. Commun. 57, 233–243. Extended UNIQUAC model. Fluid Phase Equilib. 392, 24–32.
Kapteina, S., Slowik, K., Verevkin, S.P., Heinstz, A., 2005. Vapor pressures and Sadegh, N., 2013. Acid Gas Removal from Natural Gas with Alkanolamines: A
vaporization of a series of ethnaolamines. J. Chem. Eng. Data 50, 398–402. Modeling and Experimental Study. Technical University of Denmark, DK-2800
Kim, I., Svendsen, H.F., Børresen, E., 2008. Ebulliometric determination of Lyngby, Denmark.
vapor-liquid equilibria for pure water, monoethanolamine, Sander, B., Rasmussen, P., Fredenslund, A., 1986. Calculation of vapour-liquid
N-methyldiethanolamine, 3-(methylamino)-propylamine, and their binary and equilibria in nitric acid-water-nitrate salt systems using an Extended
ternary solutions. J. Chem. Eng. Data 53 (11), 2521–2531. UNIQUAC equation. Chem. Eng. Sci. 41 (5), 1185–1195.
Kim, I., 2009. Heat of Reaction and VLE of Post Combustion CO2 Absorbents. PhD 2014. SaskPower Annual Report. http://www.saskpower.com/wp-content/
Thesis. Norwegian University of Science and Technology, Trondheim, Norway. uploads/2014-SaskPower-Annual-Report.pdf.
Klepáčová, K., Huttenhuis, P.J.G., Derks, P.W.J., Versteeg, G.F., 2011. Vapor pressures Steele, W.V., Chirico, R.D., Knipmeyer, S.E., Nguyen, A., 2002. Measurements of
of several commercially used alkanolamines. J. Chem. Eng. Data 56, 2242–2248. vapor pressure, heat capacity, and density along the saturation line for
Knudsen, J.N., Jensen, J.N., Vilhelmsen, P.-J., Biede, O., 2009. Experience with CO2 cyclopropane acid, N,N-diethylethanoalamine, 2,3-dihydrofuran,
capture from coal flue gas in pilot-scale: testing of different amine solvents. 5-hexen-2-one, perfluorobutanoic acid, and 2-phenylpropionaldehyde. J.
Energy Procedia 1, 783–790. Chem. Eng. Data 47, 715–724.
Knudsen, J.N., Andersen, J., Jensen, J.N., Biede, O., 2011. Evaluation of process Thomsen, K., Rasmussen, P., 1999. Modeling of vapor-liquid-solid equilibrium in
upgrades and novel solvents for the post combustion CO2 capture process in gas-aqueous electrolyte systems. Chem. Eng. Sci. 54, 1787–1802.
pilot-scale. Energy Procedia 4, 1558–1565. Thomsen, K., Rasmussen, P., Gani, R., 1996. Correlation and prediction of thermal
Lewis, G.N., Randall, M., 1921. The activity coefficient of strong electrolytes. J. Am. properties and phase behaviour for a class of aqueous electrolyte systems.
Chem. Soc. 43, 1112–1154. Chem. Eng. Sci. 51, 3675–3683.
Liang, Z., Fu, K., Idem, R., Tontiwachwuthikul, P., 2015a. Review on current Thomsen, K., 1997. Aqueous Electrolytes: Model Parameters and Process
advances, future challenges and consideration issues for post-combustion CO2 Simulation. Ph.D. Thesis. Technical University of Denmark, Kongens Lyngby,
capture using amine-based absorbents. Chin. J. Chem. Eng., http://dx.doi.org/ Denmark.
10.1016/j.cjche.2015.06.013 (in press). Thomsen, K., 2009. Electrolyte Solutions: Thermodynamics, Crystallization,
Liang, Z., Rongwong, W., Liu, H., Fu, K., Gao, H., Cao, F., Zhang, R., Sema, T., Henni, A., Separation Methods. Technical University of Denmark, DTU Chemical
Sumon, K., Nath, D., Gelowitz, D., Srisang, W., Saiwan, C., Benamor, A., Al-Marri, Engineering, Kongens Lyngby, Denmark.
M., Shi, H., Supap, T., Chan, C., Zhou, Q., Abu-Zahra, M., Wilson, M., Olson, W., Tontiwachwuthikul, P., Idem, R., 2013. Recent Progress and New Developments in
Idem, R., Tontiwachwuthikul, P., 2015b. Recent progress and new Post-Combustion Carbon-Capture Technology with Reactive Solvents. Future
developments in post-combustion carbon-capture technology with amine
based solvents. Int. J. Greenh. Gas Control 40, 26–54.
424 M.W. Arshad et al. / International Journal of Greenhouse Gas Control 53 (2016) 401–424

Science Book Series. Future Science Ltd., http://dx.doi.org/10.4155/ properties. Selected values for inorganic and C1 and C2 organic substances in
9781909453340. SI units. J. Phys. Chem. Ref. Data 11 (Suppl. 2).
Verevkin, S.P., Chernyak, Y., 2012. Vapor pressure and enthalpy of vaporization of Zhang, J., Misch, R., Tan, Y., Agar, D.W., 2011. Novel thermomorphic biphasic amine
aliphatic propanediamines. J. Chem. Thermodyn. 47, 328–334. solvents for CO2 absorption and low-temperature extractive regeneration.
Wagman, D.D., Evans, W.H., Parker, V.B., Schumm, R.H., Halow, I., Bailey, S.M., Chem. Eng. Technol. 34 (9), 1481–1489.
Kenneth, L.C., Nuttall, R.L., 1982. The NBS tables of chemical thermodynamic

You might also like