You are on page 1of 15

Process Safety and Environmental Protection 141 (2020) 95–109

Contents lists available at ScienceDirect

Process Safety and Environmental Protection


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

Modeling and validation of carbon dioxide absorption in aqueous


solution of piperazine + methyldiethanolamine by PC-SAFT and
E-NRTL models in a packed bed pilot plant: Study of kinetics and
thermodynamics
Arash Esmaeili a , Zhibang Liu a , Yang Xiang a,∗ , Jimmy Yun b,c , Lei Shao a,∗
a
Research Center of the Ministry of Education for High Gravity Engineering and Technology, Beijing University of Chemical Technology, Beijing 100029,
China
b
College of Chemical and Pharmaceutical Engineering, Hebei University of Science and Technology, Shijiazhuang, Hebei Province 050018, China
c
School of Chemical Engineering, The University of New South Wales, Sydney, NSW 2052, Australia

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

Article history: A pilot plant with a closed cycle of the absorption/desorption process has been taken into account for the
Received 7 February 2020 simulation of carbon dioxide (CO2 ) capture by piperazine (PZ) + methyldiethanolamine (MDEA) solution
Received in revised form 4 May 2020 using Aspen Plus rate-based model with all design and operational parameters such as the hydraulic
Accepted 4 May 2020
specifications of absorber and stripper as well as inlet flue gas conditions which are present in a commer-
Available online 13 May 2020
cial gas-fired burner for heating houses. Metal FLEXIPAC 250Y has been considered as the packing type to
apply the model to the proposed correlations on Aspen Plus for calculation of flooding and pressure drop.
Keywords:
In order to simulate the process, a new property package has been developed by electrolyte non-random
Absorption
Aspen plus two-liquid (E-NRTL) model and perturbed-chain statistical associating fluid theory (PC-SAFT) equation
Carbon dioxide of state which are used for the calculation of activity and fugacity coefficients respectively. The model
PC-SAFT has been compared with some experimental data e.g. CO2 absorption efficiency and CO2 loading and
Piperazine demonstrated good agreement with them.
Simulation © 2020 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction non-random two-liquid (E-NRTL) model. Kamps et al., 2003 mod-


eled the experimental data by modified Pitzer equation to calculate
The increase of carbon dioxide (CO2 ) content, as one of most activity coefficients. Ermatchkov and Maurer (2011) utilized Pitzer
important greenhouse gases, from industrial and domestic sources equation for activity coefficients and applied Virial equation of state
into the atmosphere is an unavoidable issue in human life. Various (EOS) for fugacity coefficients calculation. Najibi and Maleki (2013)
technologies are being used for CO2 capture, and one of proved performed the estimation for liquid and vapor phases by extended
processes is amine absorption based on countercurrent contact of Debye-Huckel model and Peng-Robinson equation of state. Bishnoi
CO2 and an amine solution in absorber for which literature has been and Rochelle (2002a), Bishnoi and Rochelle (2002), Huang et al.
published with a wide range of solvents and operational conditions. (2015) and Hemmati et al. (2019) used E-NRTL model and Redlich-
Among all studies on the blended amine solutions, some Kwong-Soave (SRK) equation of state for the liquid and vapor
researchers have worked on CO2 capture by piperazine (PZ) + properties. As a new approach, Yan and Chen (2010) have worked
methyldiethanolamine (MDEA) solution and calculated CO2 sol- on CO2 solubility in aqueous solutions of NaCl and Na2 SO4 based
ubility either as experimentally e.g. Zhang et al. (2003) and Ali on E-NRTL model and perturbed-chain statistical associating fluid
and Aroua (2004) or by various models and equations of state to theory (PC-SAFT) equation of state. Aqueous solutions of MDEA are
estimate activity and fugacity coefficients and CO2 vapor pressure. used for the selective removal of H2 S in the presence of CO2 . Due to
Dash et al. (2011) and Mudhasakul et al. (2013) used electrolyte required low reboiler heat duty, high stability and capacity, MDEA
is a key amine absorbent for CO2 capture, the activators like PZ are
commonly added to enhance the rate of CO2 absorption in aqueous
∗ Corresponding authors. MDEA solution (Ermatchkov and Maurer, 2011).
E-mail addresses: xiangy@mail.buct.edu.cn (Y. Xiang), shaol@mail.buct.edu.cn In this study, the model used by Yan and Chen (2010) is applied
(L. Shao). for the calculation of CO2 solubility in the blended solution of PZ

https://doi.org/10.1016/j.psep.2020.05.010
0957-5820/© 2020 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
96 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

Nomenclature l liquid phase


m molar, molecule
Symbols Used s solvent
A Debye-Huckel parameter v vapor phase
Cp heat capacity (J kmol−1 K−1 ) w water
ds solvent density (kg m−3 )
∗E
Gm molar excess Gibbs free energy (J kmol−1 ) Abbreviations
G (T ) Gibbs free energy (J kmol−1 ) E − NRTL electrolyte non-random two-liquid

Gj (T ) reference-state Gibbs free energy change for reac- MDEA methyldiethanolamine
tion j (J kmol−1 ) PC − SAFT perturbed-chain statistical associating fluid the-
f G Gibbs free energy of formation (J kmol−1 ) ory
Hij Henry’s law constant of solute i in solvent j (kPa) PZ piperazine
H f H
f H enthalpy of formation (J kmol−1 )
Habs vap Hs (T ) + MDEA because E-NRTL model is the most versatile and suitable
vap Hs (T ) Ix electrolyte property method for aqueous and mixed solvent sys-
Ix ionic strength of solvent tems as required data for pure components and binary interaction
Kj equilibrium constant of reaction j parameters of common species such as CO2 , H2 O and hydrocar-
k Boltzman constant (1.38065×10−23 J K−1 ) bons were stored in Aspen Plus databank (Mudhasakul et al., 2013),
kAB effective association volume and PC-SAFT equation of state is able to predict vapor proper-
M molecular weight (kg kmol−1 ) ties for its capability even in high pressures. The essence of this
m segment number approach (SAFT) is to use a reference fluid that incorporates both
NA Avogadro’s number (6.02205×1023 mol−1 ) the chain length (molecular size and shape) and molecular associ-
n number of moles ation, in place of the much simpler hard-sphere reference fluid.
P ref reference pressure (kPa) The molecularly based EOS not only provides a useful thermo-
P total pressure (kPa) dynamic basis for deriving chemical potentials or fugacities that
P0 saturation pressure (kPa) are required for phase equilibrium simulations but also allows
p0,l
j
vapor pressure of solution components (kPa) for separating and quantifying the effects of molecular structure
Qe electron charge (1.60219×10−19 C) and interactions on bulk properties and phase behavior (Chen
R gas constant (8.314 J mol−1 K−1 ) et al. (2012)). Five key parameters are used to apply PC-SAFT
T temperature (K) equation of state; segment number (m), segment energy (ε/kB ),
v partial molar volume (m3 kmol−1 ) segment size (), association energy (εAB /kB ) and effective asso-
vij reaction stoichiometric coefficient of component i ciation volume (kAB ) which have been outlined by Shahriari et al.
in jth reaction (2012), Khoshsima and Shahriari (2017), Khoshsima and Shahriari
x mole fraction in liquid phase (2018). Despite the good capability of the Peng-Robinson EOS
y mole fraction in vapor phase in correlating vapor-liquid equilibrium (VLE) of binary mixtures,
zi charge number of ion i its limitation in reproducing some properties has been discussed
in the literature while PC-SAFT model is capable of reproduc-
Greek letters ing experimental data with the same deviations for ternary and
˛ non-randomness factor quaternary mixtures and pure hydrocarbons (Nascimento et al.,
 activity coefficient 2019). Sodeifian et al. (2018) have studied different EOS to evaluate
ε dielectric constant drug solubility in supercritical CO2 , among which PC-SAFT repre-
ε/kB segment energy (K) sented more reliable data as CO2 has quadrupole momentum and
εAB /kB association energy (K) makes it special among other compounds. Rodriguez and Beckman
 CO2 absorption efficiency (%) (2020) have derived SAFT equation of state to design the non-
 closest approach parameter idealities and molecular interactions at different levels and PC-SAFT
 segment size (Å ) has been successfully used to model various biodiesel compo-
binary energy interaction parameter nents (alcohols, CO2 and glycerol) alone or in binary and ternary
 fugacity coefficient mixtures.
For the simulation of CO2 absorption, it is highly recommended
Subscripts and Superscripts that the rate-based model is used instead of equilibrium model.
∞ infinite dilution The characteristic that distinguishes this model from the equi-
a, a’ anion librium model is the accuracy of results in such a way that the
abs Am rate-based model considers the following parameters in order
Am amine to predict the performance of CO2 absorption process and the
aq aqueous phase required energy of solvent regeneration: mass transfer correla-
BO Brelvi-O’Connell tion, effective interfacial area, holdup, flooding, thermodynamic
c, c ’ cation and kinetic model, heat transfer correlation, physicochemical prop-
E, ex excess erties (density, viscosity, diffusivity, surface tension, and etc.) and
f formation flow model. The rate-based model is a combination of kinetic and
i species index thermodynamic parameters of reaction. In addition, the model is
ig ideal gas an extension of RadFrac model in Aspen Plus which includes the
ig → l departure from ideal gas to liquid calculation of the reaction rate in film, mass transfer between
j reaction, species index phases, electrolyte chemistry and hydrodynamics because mass
and heat transfer correlations are utilized in this model in place
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 97

Table 1
Parameters of PC-SAFT equation of state for solute and solution components (Zhang and Chen, 2011a; Gross and Sadowski, 2002; Razavi et al., 2019).

Components
Parameters Name in Aspen Plus
CO2 H2 O MDEA PZ

Segment number, m PCSFTM 2.569 1.065 3.304 2.652


Segment energy, ε/kB (K) PCSFTU 1.521 × 102 3.665 × 102 2.374 × 102 2.741 × 102
Segment size,  (Å ) PCSFTV 2.563 3.000 3.597 3.713
Association energy, εAB /kB (K) PCSFAU 0.000 2.500 × 103 3.710 × 103 1.923 × 103
Effective association volume, kAB (Å ) PCSFAV 0.000 3.486 × 10−2 6.645 × 10−2 7.275 × 10−2

Table 2
The coefficients of chemical equilibrium constants used in E-NRTL model.

Parameter
Equilibrium Constant
a b c d T (K) Reference

K1 1.329 × 102 −1.344 × 104 −2.247 × 10 0.000 273−498 Derks et al. (2005), Cullinane and Rochelle
(2005), Jakobsen et al. (2005)
K2 2.314 × 102 −1.209 × 104 −3.678 × 10 0.000 273−498 Derks et al. (2005), Cullinane and Rochelle
(2005), Jakobsen et al. (2005)
K3 2.160 × 10 2
−1.243 × 10 4
−3.548 × 10 0.000 273−498 Derks et al. (2005), Cullinane and Rochelle
(2005), Jakobsen et al. (2005)
K4 5.143 × 102 −3.491 × 104 −7.460 × 10 0.000 273−323 Dash et al. (2011), Dash and Bandyopadhyay
(2016)
K5 4.665 × 102 1.614 × 103 −9.754 × 10 2.471 × 10−1 273−343 Dash et al. (2011), Dash and Bandyopadhyay
(2016)
K6 6.822 −6.066 × 103 −2.290 3.600 × 10−3 273−343 Dash et al. (2011), Dash and Bandyopadhyay
(2016)
−3
K7 −1.156 × 10 1.769 × 10 3
−1.467 2.400 × 10 273−343 Dash et al. (2011), Dash and Bandyopadhyay
(2016)
K8 −3.686 −6.754 × 103 0.000 0.000 273−328 Dash and Bandyopadhyay (2016)

of equilibrium and efficiency factors to predict column perfor- 2.1. Aqueous-phase chemical equilibrium
mance.
In fact, the equilibrium approach is inappropriate for reactive As mentioned earlier, the model used to predict thermodynamic
absorption while the rate-based model uses the mass transfer and properties is the combination of E-NRTL to explain H2 O-CO2 -PZ-
interfacial area correlations to predict separation between phases MDEA chemical properties and activity coefficients in the liquid
with a high degree of accuracy as demonstrated by Onda et al. phase in addition to PC-SAFT equation of state which is applied
(1968), Bravo and Fair (1982) and Billet and Schultes (1993), even to describe vapor phase properties and fugacity coefficients. The
though it suffers from the challenge of convergence and its sen- PC-SAFT equation of state is applied for pure associating compo-
sibility to changes in the input because its governing equations nents and mixtures of solvents and gases as well as vapor-liquid
are highly nonlinear (Mudhasakul et al., 2013; Hemmati et al., and liquid-liquid equilibrium of binary mixtures of associating
2019). substances. The PC-SAFT equation of state requires five pure-
The theory used for this model is the two-film theory, based component parameters for properties prediction, two of which
on which CO2 absorption is implemented from the gas phase into characterize the association (Gross and Sadowski, 2002). Pure com-
the liquid phase in such a way that the fluid is divided into two ponent parameters of PC-SAFT equation of state for CO2 and the
parts, bulk and film. CO2 diffuses through gas film and is absorbed constituents of solution can be found in Table 1.
in the liquid phase; moreover, absorption reactions CO2 with PZ + The chemical equilibrium reactions for electrolyte systems are
MDEA are so fast and can be considered as the pseudo-first-order instantaneous and involve only proton transfer. Such reactions in
reactions which are the only contributing factors in the liquid phase conjunction with E-NRTL are used in all streams and unit operations
as Hemmati et al. (2019) mentioned. of the model as follows (Mudhasakul et al., 2013; Zhang and Chen,
2011a):

2. Chemical and phase equilibrium 2H2 O ↔ K1 H3 O+ + OH− (R1)

In order to simulate and model CO2 absorption in amine CO2 + 2H2 O ↔ K2 HCO3− + H3 O+ (R2)
solutions, CO2 solubility should be estimated by both chemical +
HCO3− + H2 O ↔ K3 CO2−
3 + H3 O (R3)
equilibrium and physical solubility in water and amine solvents
(Zhang and Chen, 2011a). In fact, thermodynamic equilibrium is PZ + H3 O+ ↔ K4 PZH+ + H2 O (R4)
expressed by combining the conditions for VLE with the condi- + − − +
tions for chemical equilibrium in the liquid phase (Ermatchkov PZH COO + H2 O ↔ K5 PZCOO + H3 O (R5)
and Maurer, 2011). If one of the compounds of CO2 , PZ or MDEA PZ + HCO3− ↔ K6 PZCOO + H2 O−
(R6)
is dissolved in water, with the exception of very dilute solutions,
− −
chemical reaction can be neglected as Kamps et al., 2003 men- PZCOO + HCO3− ↔ K7 PZ(COO )2 + H2 O (R7)
tioned. Because of chemical reactions of CO2 with aqueous solution + +
MDEA + H3 O ↔ K8 MDEAH + H2 O (R8)
(PZ + MDEA) and due to presence of PZ carbamate and carbonate
ions in the liquid phase, the liquid phase is strongly non-ideal (Dash The other chemical reaction set for H2 O-CO2 -PZ-MDEA is ded-
and Bandyopadhyay, 2016). icated separately to the absorber because the reactions which
98 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

occur there are both equilibrium and rate-controlled. The estimated Table 3
The kinetic parameters of rate-controlled reactions with PZ + MDEA solution in the
reaction mechanism consists of two steps including zwitterion
absorber (Hemmati et al., 2019).
(HPZCOO− ) formation between CO2 and PZ during the absorption
process followed by zwitterion deprotonation through the pro- Type of Electrolyte No. of reaction k Ea (cal/mol)
duction of PZ carbamate and protonated base (Hemmati et al., − R9 1.33 × 1017 1.325 × 104
HCO3
2019). In this absorption process, PZ, MDEA, OH− and H2 O are the R10 6.63 × 1016 2.565 × 104
main base elements to provide the base-catalyzed deprotonation R11 1.70 × 1010 3.190 × 102
PZCOO−
R12 3.40 × 1023 1.416 × 104
of zwitterion into the aqueous solution. In absorber, all reactions
R13 1.04 × 1014 8.038 × 103
are assumed to be in chemical equilibrium except those of CO2 PZ(COO− )2
R14 3.20 × 1020 8.692 × 103
with OH− , CO2 with PZ, CO2 with PZCOO− and CO2 with MDEA R15 6.85 × 1010 9.029 × 103
MDEAH+ R16 6.62 × 1017 2.213 × 104
which are kinetically controlled because of their finite reaction
rates. These kinetic reactions, as power law expression, control the
absorption rate and enhance mass transfer from the gas phase to
The ideal gas Gibbs free energy of the solvent is calculated from
the liquid phase in the absorber (Mudhasakul et al., 2013; Kamps
the ideal gas Gibbs free energy of formation of the solvent at 298.15
et al., 2003Hemmati et al., 2019;). ig
K (f Gs,298.15 , DGFORM in Aspen Plus), the ideal gas enthalpy of
CO2 + OH− −→ HCO3− (R9) formation of the solvent at 298.15 K (f Hs,298.15 , DHFORM in Aspen
ig

− ig
HCO3− −→ CO2 + OH (R10) Plus) and ideal gas heat capacity of the solvent (Cp,s , CPIG in Aspen
Plus). The value of these parameters accompanied with the critical
PZ + CO2 + H2 O −→ PZCOO− + H3 O+ (R11) conditions of solvent components are shown in Table 4.
PZCOO− + H3 O+ −→ PZ + CO2 + H2 O (R12)  T
ig ig ig
− − +
Gs (T) = f Hs,298.15 + Cp,s dT − T
PZCOO + CO2 + H2 O −→ PZ(COO )2 + H3 O (R13) 298.15
    
PZ COO− + H3 O+ −→ PZCOO− + CO2 + H2 O (R14) ig
f Hs,298.15
ig
− f Gs,298.15 T ig
Cp,s
2
× + dT (3)
MDEA + CO2 + H2 O −→ MDEAH+ + HCO− (R15) 298.15 298.15
T
3

MDEAH+ + HCO−
3 −→ MDEA + CO2 + H2 O (R16)
The ideal gas heat capacity of solution components and solute
ig
One single mole of piperazine theoretically might absorb two (Cp , J/kmol.K) can be calculated using Aspen ideal gas heat capac-
moles of CO2 because of its cyclic secondary diamine struc- ity polynomial (see Fig. A2) and the coefficients are illustrated in
ture and may possibly enhance the carbamate formation (Ali Table 5 as follows (Yan and Chen, 2010):
Khan et al., 2017). At low CO2 loading, the dominant reaction ig
Cp = C1 + C2 T + C3 T2 + C4 T3 + C5 T4 + C6 T5 for C7 ≤ T ≤ C8 (4)
products are piperazine carbamate (PZCOO− ) and protonated
piperazine (PZH+ ). However, PZCOO− and piperazine dicarbamate ig
Cp = C9 + C10 TC11 for T < C7 (5)
(PZ(COO− )2 ) are gradually converted to protonated piperazine
carbamate (HPZCOO− ) at higher CO2 loadings. Therefore, the For molecular solute of CO2 , the Gibbs free energy in aqueous
dominant reaction product is HPZCOO− at high solution loading. phase infinite dilution is calculated from Henry’s law:
Although piperazine dicarbamate is present, it accounts for only a

∞,aq ig Hi,w
small amount of piperazine (Dash et al., 2011). Gi (T) = f Gi (T) + RT ln (6)
Each of the above equilibrium chemical reaction (R1-R8) is char- Pref
acterized by the thermodynamic equilibrium constant which is ∞,aq
where Gi (T ) is the mole fraction-scale aqueous-phase infinite
calculated from the change of Gibbs free energies and has temper- ig
dilution Gibbs free energy of solute i at temperature T, f Gi (T ) is
ature dependency defined in mole-fraction scale (Dash et al., 2011;
the ideal gas Gibbs free energy of formation of solute i at temper-
Zhang and Chen, 2011a; Rayer et al., 2014):
ature T, Hi,w is the Henry’s constant of solute i in water and P ref is
Gj (T)

 b the reference pressure. The pure component parameters of solutes
lnKj = − = vij ln(xi i ) = a + + C lnT + dT (1) are presented in Table 6.
RT T
i For ionic species, the Gibbs free energy in aqueous-phase infi-
◦ nite dilution (mole fraction scale) is calculated from the following
where Kj is the equilibrium constant of reaction j, Gj (T ) is the
correlation (7), which includes the molality scale Gibbs free energy
reference-state Gibbs free energy change for reaction j at temper- of formation of solute i in aqueous-phase infinite dilution at 298.15
ature T, vij is the reaction stoichiometric coefficient of component ∞,aq
K (f Gi,298.15 , DGAQFM in Aspen Plus), the enthalpy of formation of
i in j th reaction, xi is the mole fraction of the component i and ∞,aq
solute i in aqueous-phase infinite dilution at 298.15 K (f Hi,298.15 ,
i is the activity coefficient of the component i. The mole-fraction
scale equilibrium constants of chemical reactions R1-R8 are given DHAQFM in Aspen Plus) and the heat capacity of solute i in aqueous-
∞,aq
phase infinite dilution (Cp,i , CPAQ in Aspen Plus). The last term,
in Table 2 and kinetics parameters of reactions R9-R16 are shown  1000 
in Table 3. RT ln Mw
, is added to convert molality concentration scale of
For the aqueous phase reactions, the reference states chosen ∞,aq ∞,aq
f Gi,298.15 to mole-fraction scale of Gi . These parameters are
are pure liquid of the solvents (PZ, MDEA and water) and aqueous indicated in Tables 7–10.
phase infinite dilution for the solutes (ionic and molecular). The
Gibbs free energy of a solvent (see Fig. A1) is estimated by the ideal  T

ig ∞,aq ∞,aq ∞,aq


gas Gibbs free energy of the solvent at temperature T, Gs (T ), and Gi (T) = f Hi,298.15 + Cp,i dT − T
298.15
Gibbs free energy departure from ideal gas to liquid at temperature
ig→l  ∞,aq ∞,aq
 T ∞,aq
T, Gs (T ), which are expressed as (Zhang and Chen, 2011a): f Hi,298.15 − f Gi,298.15 Cp,i 1000
× + dT + RTln (7)
ig ig→l
298.15 T Mw
Gs (T) = Gs (T) + Gs (T) (2) 298.15
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 99

Table 4
The pure component parameters of solution constituents used in this model.

Parameter
Component
ig ig
OMEGA PC (kPa) TC (K) VC (m3 /kmol) ZC f Hs,298.15 (J/kmol) f Gs,298.15 (J/kmol)

PZ (Derks et al., 2005; Hilliard, 2008) 3.191 × 10−1 5.800 × 103 6.610 × 102 3.100 × 10−1 3.230 × 10−1 1.640 × 107 1.700 × 108
MDEA (Zhang and Chen, 2011a, b) 1.165 3.880 × 103 6.750 × 102 3.680 × 10−1 2.560 × 10−1 −3.800 × 108 −1.690 × 108
H2 O (Yan and Chen, 2010; Rayer et al., 2014) 3.440 × 10−1 2.205 × 104 6.473 × 102 5.590 × 10−2 2.290 × 10−1 −2.420 × 108 −2.287 × 108

Table 5
The coefficients of ideal gas heat capacity (J/kmol.K).

Component
Coefficient
H2 O CO2 PZ MDEA

C1 3.374 × 104 1.979 × 104 9.849 × 104 2.730 × 104


C2 −7.018 7.343 × 10 7.300 × 102 5.408 × 102
C3 2.729 × 10−2 −5.602 × 10-2 0.000 0.000
C4 −1.660 × 10-5 1.720 × 10−5 0.000 0.000
C5 4.300 × 10−9 0.000 0.000 0.000
C6 −4.170 × 10-13 0.000 0.000 0.000
C7 2.000 × 102 3.000 × 102 2.780 × 102 2.780 × 102
C8 3.000 × 103 1.088 × 103 3.970 × 102 3.970 × 102
C9 3.325 × 104 2.910 × 104 0.000 0.000
C10 1.900 × 10−20 7.187 × 10−1 0.000 0.000
C11 9.284 1.637 0.000 0.000
Reference Yan and Chen (2010), Zhang and Chen (2011a) Yan and Chen (2010), Zhang and Chen (2011a) Aspen databank Zhang and Chen (2011a)

Table 6
The pure component parameters of solutes used in this model (Dash and Bandyopadhyay, 2016; Rayer et al., 2014).

Parameter
Component
ig ig
OMEGA PC (bar) TC (K) VC (m3 /kmol) ZC f Hs,298.15 (J/kmol) f Gs,298.15 (J/kmol)

CO2 2.250 × 10−1 7.380 × 10 3.042 × 102 9.390 × 10−2 2.740 × 10−1 −3.935 × 108 −3.943 × 108
N2 4.000 × 10−2 3.400 × 10 1.261 × 102 9.000 × 10−2 – 0.000 –
O2 1.900 × 10−2 5.080 × 10 1.551 × 102 7.320 × 10−2 – 0.000 –

Table 7
The Gibbs free energy and enthalpy of formation of PZ and MDEA ionic species (Huang et al., 2015; Zhang and Chen, 2011a; Hilliard, 2008).

Electrolyte Species
Parameter Name in Aspen Plus
PZH + PZH +2 PZCOO− PZ(COO− )2 HPZCOO− MDEAH +
∞,aq
f Gi,298.15 × 10−8 (J/kmol) DGAQFM 1.024 0.919 −2.164 −5.766 −2.734 −2.595
∞,aq
f Hi,298.15 × 10−8 (J/kmol) DHAQFM −0.915 −1.226 −4.820 −8.606 −5.224 −5.109

Table 8
The Gibbs free energy and enthalpy of formation of water and CO2 ionic species (Kamps et al., 2003; Hilliard, 2008; Zhang and Chen, 2011b).

Electrolyte Species
Parameter Name in Aspen Plus
H3 O+ OH − HCO3 − CO32−
∞,aq
f Gi,298.15 × 10−8 (J/kmol) DGAQFM −2.371 −1.572 −5.867 −5.278
∞,aq
f Hi,298.15 × 10−8 (J/kmol) DHAQFM −2.858 −2.299 −6.919 −6.771

Table 9
The heat capacity of PZ and MDEA ionic species in aqueous-phase infinite dilution.

Electrolyte Species
Parameter
PZH + PZH +2 PZCOO− PZ(COO− )2 HPZCOO− MDEAH + Reference
∞,aq
Cp,i (J/kmol.K) 1.939 × 10 5
0.000 3.024 × 10 5
7.204 × 10 4
2.377 × 10 4
3.320 × 105 Aspen Databank

Table 10
The heat capacity of water and CO2 ionic species in aqueous-phase infinite dilution.

Electrolyte Species
Parameter
H3 O+ OH − HCO3 − CO32− Reference
∞,aq
Cp,i (J/kmol.K) 7.529 × 10 4
−1.485 × 10 5
−2.926 × 10 4
−3.971 × 105 Zhang and Chen (2011b)
100 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

Table 11 critical volume in the Aspen databank (see Table A1). The char-
The calculated properties based on PC-SAFT equation of state.
acteristic volume are expressed as temperature dependent in the
Name in Aspen Plus
Parameter following correlation:
PHIVMX Fugacity coefficient of vapor mixture
PHIV Fugacity coefficient of vapor pure component
VBO
i = v∞
i
= v1,i + v2,i T (12)
DHVMX Molar enthalpy departure of vapor mixture
DHV Molar enthalpy departure of vapor pure component Whereas CO2 reacts with amine solutions, the physical solubility
DGVMX Molar Gibbs free energy departure of vapor mixture of CO2 in amine solutions cannot be measured directly. As a result,
DGV Molar Gibbs free energy departure of vapor pure component a non-reacting gas such as NO2 can be used as a surrogate to CO2
DSVMX Molar entropy departure of vapor mixture
in estimating the physical solubility of CO2 in these solvents. The
DSV Molar entropy departure of vapor pure component
VVMX Molar volume of vapor mixture
N2 O analogy has been frequently used to estimate the solubility of
VV Molar volume of vapor pure component CO2 in amine solutions. The N2 O analogy for the solubility of CO2 in
amine solution has the following relations (Samanta et al., 2007):

2.2. Vapor-liquid (physical) equilibrium HCO2


HCO2 −AM = HN2 O−AM × ( ) (13)
HN2 O in water
Based on the knowledge of thermodynamics, when the equi-
librium between the vapor and liquid phases is established, the Versteeg and van Swaaij (1988) proposed two convenient equa-
fugacity of the liquid and vapor must be equal (Kamps et al., 2003; tions based on the available data of solubility of N2 O and CO2 in
Xia et al., 2003). Therefore, the activity coefficients can be calcu- water which are as follows:
lated. The extended Raoult’s law is used to express the vapor-liquid mol 2284
equilibrium for solvents (water, PZ, MDEA), while the extended HN2 O−H2 O = 1.17 × 10−7 × exp ( ) (14)
m3 .Pa T
Henry’s law is applied for the estimation of activity coefficients
in vapor-liquid equilibrium of solute (carbon dioxide) (Dash et al., mol 2044
HCO2 −H2 O = 3.54 × 10−7 × exp ( ) (15)
2011): m3 .Pa T

vi (P − P0i ) The components CO2 , CO, N2 , O2 and H2 are considered as Henry
iv .yi . P = xi .i .P0i . exp (8) components (solutes) for which Henry’s law is applied and Henry’s
RT
  constant are specified for these components with respect to PZ,
v∞
i
(P − P0j ) MDEA and water (see Table A2). The above-mentioned equations
iv .yi . P = xi .i* .Hij (T, P). exp (9) are used to correlate the Henry’s constants of solute in pure solvent
RT
j at system temperature and the solvent vapor pressure (p0,l
j
) which
i can be described by the following expression (Yan and Chen, 2010):
= i∞ (10)
i* bij
lnHij T, p0,l
j
= aij + + cij lnT (K) + dij T(K) (16)
where i is expressed as solvent in Raoult’s law and solute in Henry’s T (K)
law; iv is the vapor phase fugacity coefficient; yi is mole fraction in
The Henry’s constant of solute i in pure solvent j at system
vapor phase; P is total pressure (kPa); xi is mole fraction in liquid
temperature and pressure (Hi,j (T, P)) can be corrected with the
phase; i is activity coefficient; P o is saturation pressure (kPa); vi
Poynting term for pressure. At low pressures, the Poynting correc-
is partial molar volume (m3 /kmol); v∞ i
is partial molar volume at
tion is approximately unity and can be neglected (Zhang and Chen,
aqueous-phase infinite dilution (m3 /kmol); Hij is Henry’s law con- 2011a; Bishnoi and Rochelle, 2000).
stant of solute i in solvent j (kPa). As mentioned before, PC-SAFT
  
equation of state is used to estimate the vapor phase properties P
1
and the following parameters in Table 11 are calculated in Aspen Hi,j (T, P) = Hij T, p0,l
j
exp v∞
i
dp (17)
RT 0,l
Plus. p
j
Bishnoi and Rochelle (2002a) presented the initial concentration
of PZ in the solvent is limited by its solubility in water and MDEA The extended Antoine equation (PLXANT in Aspen Plus) is uti-
solutions in such a way that the solubility below 40 ◦ C (313 K) is lized to calculate vapor pressure of solution components which is
limited by formation of the PZ hexahydrate which melts at 316 K. expressed as (Yan and Chen, 2010):
Therefore the minimum temperature for PZ absorption should be
C2i
40 ◦ C which has been considered in this study. lnpo,l = C1i + + C4i T + C5i lnT + C6i TC7i for C8i ≤ T ≤ C9i (18)
j T + C3i
The infinite dilution activity coefficient (i∞ ) of PZ in water is
calculated from predicted activity coefficient data obtained using where C1i to C9i are the correlation parameters. The extended
modified UNIFAC model. The nonlinear temperature dependence Antoine equation parameters for PZ were obtained from Aspen Plus
of this parameter is represented as (Dash et al., 2011): databank, and these parameters for CO2 and solution components
25, 464 are illustrated in Table 12. The value for ions are set to −1.0 × 1020

lnPZ = 433.15 − − 61.3 lnT(K) (11) (Onda et al., 1968) as they are non-volatile.
T (K)
The activity coefficient of any species (solute or solvent, ion or
The partial molar volume of solute i at aqueous-phase infinite molecule) is related to the molar excess Gibbs free energy and
dilution (v∞i
) is calculated by using the Brelvi-O’Connell model determined by the following thermodynamic correlation (Dash
(Bishnoi and Rochelle, 2000). It represents the universal functions et al., 2011; Hilliard, 2008):
relating the compressibility to the reduced density and the solute
partial molal volume to the reduced solvent density. As per these G*E  (nG*E
m / RT)
lni = m
= 冀 冁 (19)
empirical correlations, the compressibility of a liquid and the par- RT ni T,P, nj =
/ i
tial molal volume of a solute at infinite dilution in the solvent can be
estimated only by using the characteristic volume for the required where i, j = m, c and a, which represent molecules, cations and
solute and the solvent. Those of PZ and MDEA are defaulted to their anions respectively and the excess Gibbs free energy associated
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 101

Table 12
The extended Antoine equation parameters for CO2 and solution components (Dash et al., 2011; Zhang and Chen, 2011a; Hilliard, 2008).

Component
Parameter
H2 O CO2 PZ MDEA Ions

C1i 7.255 × 10 7.283 × 10 6.236 × 10 1.227 × 102 −1.000 × 1020


C2i −7.206 × 103 −3.403 × 103 −7.102 × 103 −1.325 × 104 0.000
C3i 0.000 0.000 0.000 0.000 0.000
C4i 0.000 9.490 × 10−3 0.000 0.000 0.000
C5i −7.138 −8.560 −5.670 −1.384 × 10 0.000
C6i 4.050 × 10−6 2.910 × 10−16 1.720 × 10−6 3.200 × 10−6 0.000
C7i 2.000 6.000 2.000 2.000 0.000
C8i 3.192 × 102 2.730 × 102 2.730 × 102 2.930 × 102 0.000
C9i 6.473 × 102 3.042 × 102 6.610 × 102 7.420 × 102 2.000 × 103

with the E-NRTL model is given by the following equation (Rayer The subscripts m, c and a represent molecules, cations and anions
et al., 2014; Hilliard, 2008). respectively (Dash et al., 2011).

G*E G*E,PDH G*E,Born G*E,local  ⎛ ⎞


m
= m + m + m (20) Xj Gjm
RT RT RT RT jm

G*E,local
  ⎜ X ’ ⎟
Therefore, applying Eq. 20 to Eq. 19 yields, m
= Xm 
j
+ Xc ⎜ a ⎟
RT
Xk Gkm
⎝ X

a"
lni = lniPDH + lniBorn + lnilocal (21) m C a’

k a"

The first term in Eq. 20 accounts for the Pitzer-Debye-Huckel  ⎛ ⎞ 


Xj Gjc,a’ c Xj Gja,c’ a
(PDH) long-range ion-ion interactions (ionic forces) that occur at jc,a’ c ja,c’ a
 ⎜ X ’ ⎟
low ionic strengths and the reference state for ionic speciation is × 
j
+ Xa ⎜ c ⎟ × j (27)
the infinite dilute state of electrolyte in the mixed solvent (Dash Xk Gkc,a’ c
⎝ X
⎠ X Gka,c’ a
a c’ c" k
et al., 2011; Huang et al., 2015):
c"

k k

G*E,PDH  1000 0.5 4A Ix  


m 0.5
= − xk ln 1 + Ix (22) where Xj = xj .Cj (Cj = zj for ions and unity for molecules). The
RT Ms 
k other terms are defined as follows:

where Ms is molecular weight of the solution and  is the closest Xa Gca,m
approach parameter. Debye-Huckel parameter, A , as well as ionic a
strength of solvent, Ix , are given by the following equations (Dash Gcm =  = exp (−˛cm cm ) (28)
et al., 2011; Rayer et al., 2014): Xa’
 1.5
a’
1 2 NA ds 0.5
Q2e 
A = (23) Xc Gca,m
3 1000 εs kT

1  Gam =
c
 = exp (−˛am am ) (29)
Ix = xi z2i (24) Xc’
2
i c’

where NA is Avogadro’s number; ds is the solvent density (kg/m3 );


 
Gjc,a’ c = exp −˛jc,a’ c . jc,a’ c (30)
Qe is the electron charge; k is Boltzman constant and zi is charge
 
number of ion i. Gja,c’ a = exp −˛ja,c’ a . (31)
ja,c’ a
The Born correction term accounts for the change in reference
state given by the difference in the dielectric constants (εi ), and it Since the interaction energies are symmetric, it can be inferred
corrects the reference state of ionic species from infinite dilution in that (Chen et al., 1982):
the solvent to infinite dilution in water (Dash et al., 2011; Bishnoi
am = cm = ca,m (32)
and Rochelle, 2000):
  xi z2 mc = ma = m,ca (33)
G*E,Born Q2e 1 1
m
= − i
10−2 (25)
RT 2kT εs εw ri In the above-mentioned correlations, ˛ij is the non-randomness
i
factor and ij is the binary energy interaction parameter. The
The dielectric constant (CPDIEC in Aspen Plus) is related to the non-randomness factor for molecule-molecule pairs is considered
ability of a component to stabilize an ionic solution (see Table A3). as 0.2 for saturated hydrocarbon with polar non-associated liq-
As the dielectric constant increases, the tendency for ions to form uids and systems that exhibit liquid-liquid immiscibility, 0.3 for
and remain as ionic species also rises, and the dielectric constant is nonpolar substances, nonpolar non-associated liquids, 0.47 for self-
defined as follows: associated substances with nonpolar substances. Non-randomness
1 1 factor (GMENCN in Aspen Plus) is considered 0.2 for molecule-
εi = Ai + Bi − (26) electrolyte pairs (Cullinane and Rochelle, 2005; Xu, 2011).
T 298.15

The E-NRTL expression for short-range interaction between Xa ˛ca,m
molecules and electrolytes with considering the fact that the ref- a
erence state for ion is the infinite dilute state in water, could be ˛mc = ˛cm =  (34)
taken into account as most important interactions because con- Xa’
tributions of the two former terms are only analytical in nature. a’
102 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

Fig. 1. The summary of chemical and physical equilibrium calculations in CO2 absorption process by amine solutions.

Fig. 2. The schematic of the pilot plan used for CO2 capture by PZ + MDEA solution in Aspen Plus.
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 103


Xc ˛ca,m
c
˛ma = ˛am =  (35)
Xc’
c’

In order to calculate the liquid-phase activity coefficients,


the binary E-NRTL interaction parameters for molecule-molecule,
molecule-electrolyte and electrolyte-electrolyte binary systems
are required. Binary energy interaction parameters for molecule-
molecule, molecule-ion pair and ion pair-ion pair are expressed as
function of temperature (Cullinane and Rochelle, 2005; Xu, 2011):

a Molecule-molecule pair parameters

Bi,j
i,j = Ai,j + + Fi,j ln(T) + Gi,j T (36) Fig. 3. Variation of PZ mole fraction in the absorber with various inlet mass concen-
T
trations.
The default value of parameters A, B, F and G is zero; however,
the binary interaction parameters used in this model are indicated
in Table 13.
The diameter of absorber and stripper is 125 mm; the packing
b Electrolyte-molecule pair parameters height of absorber and stripper is 4.20 and 2.52 m respectively with
20 stages for each column; two stages in both columns are dedi-
Dca,m

(Tref − T) T
 cated to the washing section, and the height of the washing section
ca,m = Cca,m + + Eca,m + ln (37) of both columns is 0.42 m. Metal FLEXIPAC 250Y has been consid-
T T Tref
Dm,ca

(Tref − T) T
 ered as structured packing for all stages. Flue gas flow rate is 72.0
m,ca = Cm,ca + + Em,ca + ln (38) kg/hr, and solution flow rate is 100−250 kg/hr, while absorber and
T T Tref stripper pressures are 1.0 and 2.0 bar respectively. The average flue
The default value of parameters Cm,ca , Cca,m (GMENCC in Aspen gas mass composition is N2 : 73.7 %, O2 : 12.1 %, H2 O: 6.1 % and CO2 :
Plus) as well as Dm,ca and Dca,m (GMENCD in Aspen Plus) have been 8.1 %. Inlet temperature of flue gas and PZ + MDEA solution (30
reported in the literature (see Table A4) despite the fact that the wt%) varies 48–60 ◦ C. The make-up and cooling water temperature
default value of E (GMENCE in Aspen Plus) is considered as zero is considered approximately 5–15 ◦ C to keep pre-washer water at
(Dash et al., 2011; Bishnoi and Rochelle, 2000). The vapor-liquid the absorber inlet 45–50 ◦ C and cool the gas partially in the con-
equilibrium, heat capacity and excess enthalpy data are used to denser. The rich amine solution is pumped to the lean-rich heat
determine the NRTL interaction parameters in molecule-molecule exchanger and then into the desorber to strip CO2 ; in the meantime,
binary system while the VLE data, heat capacity, heat of absorption PZ and MDEA are injected to the lean amine solution as make-up to
and NMR spectroscopic data are applied for the E-NRTL interac- compensate the loss of solution. The 30 wt% blended solution of PZ
tion parameters in molecule-electrolyte ternary system (Zhang and + MDEA is introduced to the absorber to evaluate CO2 absorption
Chen, 2011a). efficiency and rich-CO2 loading as well as the required reboiler heat
duty for full CO2 recovery.
Due to lack of enough operational and hydraulic data of
c Electrolyte-electrolyte pair parameters
absorbers as well as operational conditions of fluids in the previous
Dca,c’ a  (T − T) T
 literature where CO2 capture has been undertaken by the blended
ref
ca,c’ a = Cca,c’ a + + Eca,c’ a + ln (39) solution of PZ + MDEA, the model has been validated by the pub-
T T Tref
lished data of Ali Ali Khan et al. (2017) in the same conditions of
Dca,ca’  (T − T) T
 their studied pilot plant.
ref
ca,ca’ = Cca,ca’ + + Eca,ca’ + ln (40) In this work, the following models have been used to predict
T T Tref
accurate physicochemical properties: the Clarke model for liq-
The ion-ion pair parameters are assumed to be insignificant with uid molar volume (VAQCLK in Aspen Plus), Jones-Dole electrolyte
a value of zero as the concentration of ions in an electrolyte solu- correction model for liquid viscosity (MUL2JONS in Aspen Plus),
tion approaches zero (Hilliard, 2008). Zhang and Chen (2011a) also Onsager-Samaras model for liquid surface tension (SIGDIP in Aspen
reported that all molecule-molecule and electrolyte-electrolyte Plus), Riedel electrolyte correction model for thermal conductivity
binary parameters are defaulted to zero and the default values of (KLDIP in Aspen Plus) and Nernst-Hartley model for binary diffu-
electrolyte-molecule binary parameters can be used, although the sivity (DL1NST in Aspen Plus). PC-SAFT equation of state, which
calculated thermodynamic properties of the electrolyte solutions has been applied for acid gases capture in MDEA (Zhang and
are controlled by accurate NRTL interaction parameters associated Chen, 2011a, b), is used to predict fugacity coefficients in vapor
with the major species in the system. The schematic of the required phase. The reactions of CO2 with PZ + MDEA are very fast and
data of CO2 absorption in amine solutions by Aspen Plus in terms take place in the liquid phase with no reaction in the gas phase.
of kinetics and thermodynamics can be seen in Fig. 1. Therefore, the mixed model is considered for vapor-liquid con-
tact with the Discrxn resistance model for liquid phase and the
3. Process modeling Film resistance model for the gas phase which is based on one
of the simplest mass transfer theories assuming that all the con-
A pilot plant applied for CO2 capture by monoethanolamine centrations change occurs only in the film region. Furthermore,
(MEA) shown in Fig. 2 which was investigated by Notz et al. (2012) Barvo-Fair model is used to calculate mass transfer coefficients of
and mentioned in Aspen Plus manual has been considered in this vapor and liquid as well as interfacial area because of less devi-
study with the similar operational and design conditions as follows: ation in calculation than the model proposed by Hemmati et al.
104 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

Table 13
The coefficients of molecule-molecule binary interaction parameters for E-NRTL model.

Component i Component j Aij Bij Aji Bji ˛ij Reference

H2 O MDEA 8.509 −1.573 × 103 −1.714 −2.618 × 102 2.000 × 10−1 Zhang and Chen (2011b)
H2 O CO2 0.000 0.000 0.000 0.000 2.000 × 10−1 Bishnoi and Rochelle,
2000; Bishnoi and
Rochelle, 2002a
H2 O HPZCOO− 2.933 × 10−1 0.000 7.480 × 10−2 0.000 3.000 × 10−1
CO2 HPZCOO− 0.000 0.000 0.000 0.000 3.000 × 10−1
HPZCOO− MDEA 0.000 0.000 0.000 0.000 3.000 × 10−1 Aspen Databank
PZ MDEA −3.094 0.000 −3.113 0.000 3.000 × 10−1
H2 O PZ 5.415 −2.105 × 103 1.359 −9.604 × 102 2.000 × 10−1

Fig. 4. a. Variation of PZ electrolyte species mole fraction in the absorber with inlet 2 wt% PZ. b. Variation of PZ electrolyte species mole fraction in the absorber with inlet 3
wt% PZ. c. Variation of PZ electrolyte species mole fraction in the absorber with inlet 4 wt% PZ. d. Variation of PZ electrolyte species mole fraction in the absorber with inlet
5 wt% PZ. e. Variation of PZ electrolyte species mole fraction in the absorber with inlet 6 wt% PZ.
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 105

Fig. 5. Variation of MDEA and CO2 ion species mole fraction in the absorber with
inlet 2 wt% PZ+28 wt% MDEA. Fig. 6. Variation of CO2 loading in the absorber stages with various PZ mass con-
centrations.

(2019). Holdup and flooding are determined by Stichlmair correla-


tions.

4. Results and discussion

4.1. The effects of variation in PZ + MDEA concentration

For the first step, CO2 absorption is evaluated with five different
PZ mass concentrations at 48 ◦ C for both flue gas and lean solu-
tion while flue gas and solution mass flow rate are assumed to be
72.0 and 250 kg/h respectively. It is evident in Fig. 3 that PZ con-
centration decreases smoothly from top to bottom of the absorber
where it reacts with CO2 and ion species of PZ are generated, while
CO2 concentration reduces from the bottom to top. PZ mass con-
centration of 2–6 % is generally adopted as a consideration of the
process efficiency and cost because of lower rich-CO2 loadings com-
Fig. 7. Variation of CO2 mass transfer rate in the absorber stages with various PZ
pared to higher PZ concentrations (Ali Khan et al., 2017). Derks
mass concentrations.
et al. (2008) declared the concentration of MDEA is usually kept
about 3−4 mol/dm3 and a maximum PZ concentration of about 1.0
mol/dm3 is applied. In our work, the mass ratio of 6 wt% PZ to 24 increasing PZ mass concentration. On the other side, PZ dicarba-
wt% MDEA meets the mole ratio of 1:3 for PZ/MDEA. The value of mate (PZ(COO− )2 ) has the minimum impact on CO2 capture and
PZ degradations on the stages in contact with CO2 has decreased accounts for 17.1%–13.6% of PZ with rise in PZ content. These results
from 71 % to 67 % for 1 wt% PZ to 6 wt% PZ. are in good agreement with reports in the previous literature e.g.
The results in Fig. 4 clearly show the equilibrium concentra- (Dash et al., 2011; Bishnoi and Rochelle, 2002a; Huang et al., 2015)
tion profile for PZ electrolyte species versus absorber mass transfer for CO2 absorption by PZ + MDEA.
stages (3–20). The concentration of all PZ electrolyte species has The simulation of CO2 absorption in this study indicates that
increased with rising PZ mass concentration. The efficiency of CO2 the concentration of bicarbonate (HCO3 − ), carbonate (CO2− 3 ) and
absorption is significantly enhanced with the increase in PZ mass protonated MDEA (MDEAH+ ) stay almost constant in different PZ +
concentration as an activator in aqueous solution of MDEA, from MDEA ratios; therefore, the variation of these ion species is shown
66.8 % for 2 wt% PZ to 76.5 % for 6 wt% PZ. The simulation results in Fig. 5 only for the solution of 2 wt% PZ +28 wt% MDEA. In all five
illustrate that PZ carbamate (PZCOO− ) and protonated PZ (PZH+ ) are different concentrations of solution, the mole fraction of carbonate,
two effective electrolyte species at lower CO2 loadings and higher bicarbonate and protonated MDEA has decreased versus rising PZ
stages as Dash et al. (2011) mentioned, and their allocation (the mass concentration. As illustrated, MDEA compared with PZ has
ratio of each electrolyte to the summation of all PZ electrolytes at less reaction rate with CO2 , so two of PZ electrolyte species (PZH+
the bottom of absorber) for CO2 capture in this model has increased and PZCOO− ) have higher concentration than MDEAH+ at the upper
from 24.9%–25.4% as well as 25.8%–31.9% respectively as PZ mass stages; however, MDEAH+ concentration passes theirs along the
concentration rises from 2 wt% to 6 wt%. In addition, PZ carba- absorber as inlet concentration of MDEA is much higher than PZ in
mate and protonated PZ track each other at low CO2 loadings and all cases of this study.
low PZ concentrations as Bishnoi and Rochelle (2002a) already Moreover, most of CO2 is absorbed from the middle to the bot-
reported. However, protonated PZ keeps away PZ carbamate grad- tom of column owing to higher driving force between the flue
ually toward the bottom of absorber and their difference becomes gas and absorbent while most of physical CO2 solubility occurs
more significant as PZ concentration increases. Furthermore, pro- from the top to the middle of column (see Fig. A3). Figs. 6 and 7
tonated PZ carbamate (HPZCOO− ) plays more substantial role at show CO2 loading and mass transfer rate versus the absorber stages
the lower stages and higher CO2 loadings particularly in the lower for which interfacial area (Fig. 8) and liquid holdup (Fig. 9) have
PZ concentrations (2 wt% and 3 wt% PZ) in such a way that its been determined by Stichlmair correlations. It can be seen from
mole fraction dominates PZH+ and PZCOO− ; however, the dedi- Figs. 6 and 7 that CO2 loading and mass transfer rate have increased
cation of HPZCOO− to absorb CO2 declines from 32.2%–29.1% with with not only rising PZ mass concentration but at the lower stages of
106 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

Table 14
Variation of the parameters vs. PZ mass concentration (wt%) with solution flow rate of 200 kg/hr.

PZ conc. Mole frac. of CO2 loading in Activity Fugacity Vapor pressure Vapor pressure Max. stage Section
(wt%) CO2 in off-gas rich solution coefficient of coefficient of of CO2 in gas of CO2 in rich liquid holdup pressure drop
CO2 in liquid CO2 in vapor phase at solution (× 103 (lit) (mbar)
phase phase bottom of kPa)
column (kPa)

1 0.033 0.351 1.091 0.997 5.142 1.280 0.140 8.611


2 0.024 0.385 1.099 0.997 5.056 1.107 0.141 8.841
3 0.019 0.401 1.096 0.997 4.990 0.945 0.141 8.971
4 0.016 0.410 1.093 0.997 4.925 0.787 0.142 9.067
5 0.013 0.416 1.092 0.997 4.864 0.655 0.142 9.146
6 0.011 0.419 1.092 0.997 4.805 0.545 0.142 9.213

Fig. 8. Variation of interfacial area in the absorber stages with various PZ mass Fig. 10. CO2 absorption efficiency vs. PZ mass concentration in various solution flow
concentrations. rates.

mum solution flow rate is considered as 250 kg/h because a higher


flow rate of PZ + MDEA does not converge the model and contrarily
lead to decrease in CO2 absorption efficiency. Flow rate of 250 kg/h
has lower absorption efficiency than 200 kg/h at higher PZ mass
concentrations (5 wt% and 6 wt%). Therefore, the flow rate of 200
kg/h and the solution of 6 wt% PZ + 24 wt% MDEA with the efficiency
of 79.6 % can be regarded as the most efficient flow rate and mass
concentration for the solution compared to the pilot plant with 30
wt% MEA and the efficiency of 75.8 %. For all PZ concentrations, the
required power for the reboiler is considered 30,318 kJ/h (8.42 kW)
to reach 100 % CO2 recovery in the stripper compared with 7.95 kW
used based on MEA solution.
   
yCO2 ,out 1 − yCO2 ,in
 = 1− × 100 (41)
1 − yCO2 ,out yCO2 ,in
Fig. 9. Variation of liquid holdup in the absorber stages with various PZ mass con-
centrations.
4.2. The most efficient conditions of PZ + MDEA solution

absorber where the CO2 content and gas-phase mass transfer coef- For the second step, the solution flow rate of 200 kg/h has been
ficient are much higher in such a way their values have reached chosen as the most appropriate flow rate for this pilot plant and
the maximum point of 0.412 and 8.46 mol/h respectively for the some parameters have been evaluated with the variation of PZ mass
solution of 6 wt%PZ + 24 wt%MDEA at the stage of 20. concentration while flue gas enters the absorber at 48 ◦ C and con-
Fig. 10 indicates the variation of CO2 absorption efficiency () stant flow rate and composition. Fig. 11 illustrates CO2 absorption
versus PZ mass concentration in different solution flow rates. The efficiency with the variation of PZ mass concentration in different
results show increasing PZ mass concentration has led to more CO2 solution temperatures. Although increasing PZ mass concentration
capture because of higher reactivity and solubility as well as inter- has brought on rising absorption efficiency as shown in Fig. 11,
facial area (as can be seen in Fig. 8). In addition, the efficiency has increasing solution temperature does not have much impact on the
enhanced from 52.1%–72.4% for 4 wt% PZ with rising solution flow efficiency in such a way that only 4.3 % and 2.0 % increase is observed
rate from 100 kg/h to 250 kg/h which has created higher velocity, with rising temperature from 48 ◦ C to 60 ◦ C for 1 wt% PZ and 6
Reynolds number and higher liquid-side mass transfer coefficient. wt% PZ solutions respectively. Enhancing solution temperature not
Liquid flow rate plays an important role in CO2 absorption in packed only leads to increase in liquid-side mass transfer coefficient and
column as one of most essential parameters for the calculation of reaction rate between CO2 and the solution but also results in accel-
absorption rate. However, on the review of the interfacial area of erated backward equilibrium reactions. Thus, 48 ◦ C is considered
packing and column diameter in the present pilot plant, the maxi- as the most effective solution temperature in this case study.
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 107

Fig. 14. Validation of CO2 loading in various PZ mass concentrations with Ali Khan
Fig. 11. CO2 absorption efficiency vs. PZ mass concentration in various solution et al. (2017) studies.
temperatures.
almost unchanged at 0.997 and 1.096 respectively with increas-
ing PZ mass concentration. Vapor pressure of CO2 has declined at
the bottom of absorber from 5.142 to 4.805 kPa with PZ mass con-
centration increasing from 1% to 6% owing to a better absorption
between concentrated PZ solution and inlet CO2 and generation of
more PZ carbamate while vapor pressure (mole fraction) of CO2
has increased from the top to the bottom of column for all PZ mass
concentrations (see Fig. A4). The model predicts that capacity of
concentrated PZ is much more than that of dilute aqueous PZ as
CO2 partial pressure decreases in the gas phase; in addition, CO2
partial pressure in the rich solvent has also reduced, which indi-
cates that more CO2 has reacted with PZ to produce the electrolyte
species. In fact, the strongest effect of PZ is to reduce CO2 partial
pressure as Bishnoi and Rochelle (2002a) reported.
Fig. 12. Outlet CO2 content vs. flue gas temperature in various solution tempera- In order to validate the model, the data of pilot plant used by
tures. Ali Khan et al. (2017) have been compared with this model in
Figs. 13 and 14. The simulated data of CO2 absorption efficiency
in various gas flow rates (5 × 10−3 - 8 × 10-3 m3 /min) and rich
CO2 loading with different inlet CO2 partial pressure (10−15 kPa)
are in good agreement with the experimental data which prove the
accuracy of model. The model predicts results with a 4.76 % average
absolute deviation (AAD) between the experimental and simulated
data. The upper data for each PZ mass concentration are the simu-
lated values and that of the lower part (legends in black color) show
the reported experimental data.

5. Conclusion

The simulation of CO2 capture by aqueous PZ + MDEA solution


from a specific flue gas in a pilot plant by E-NRTL model and PC-
SAFT equation of state has been conducted. The modeling of CO2
absorption by amine solutions using Aspen Plus requires the accu-
Fig. 13. Validation of CO2 absorption efficiency in various PZ mass concentrations rate data for a wide range of parameters in terms of chemical and
with Ali Khan et al. (2017) studies.
physical solubility. Therefore, all kinetics parameters, binary and
ternary interaction parameters, Henry constant as well as enthalpy
Fig. 12 also show the effect of flue gas temperature on CO2 cap- and Gibbs free energy of formation in addition to five PC-SAFT
ture in this model. Four different temperatures between 48 ◦ C to parameters for the key components of model have been considered
60 ◦ C have been assessed for 6 wt% PZ + 24 wt% MDEA solution in this study.
with the flow rate of 200 kg/hr. The variation of flue gas temper- Based on the results of this simulation, PZ carbamate and
ature does not affect CO2 absorption at a constant amine solution protonated PZ are two important piperazine electrolyte species
temperature because the main mass transfer resistance is in liq- particularly in low CO2 loadings while protonated PZ carbamate
uid film even if the flue gas temperature increase leads to a rise in plays a more crucial role in the higher CO2 loadings. Due to higher
gas-side mass transfer coefficient. This observation is in agreement driving force at the bottom of absorber, more mass transfer rate
with the studies for CO2 capture by Lin et al. (2010). and CO2 loading are achieved there and most of CO2 is absorbed
Some estimated parameters in the most efficient case as flue in that part. Additionally, rising solution flow rate has more sig-
gas and solution temperature of 48 ◦ C and solution flow rate of 200 nificant effect on CO2 capture efficiency than increasing solution
kg/h can be found in Table 14. Fugacity coefficient of CO2 in vapor temperature while the variation of flue gas temperature does not
phase and activity coefficient of CO2 in liquid phase have remained have much impact on CO2 capture in such a way that the most effi-
108 A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109

cient operational conditions have been found as solution of 6 wt%


PZ +24 wt% MDEA, flue gas and solution temperature of 48 ◦ C and
solution flow rate of 200 kg/hr. Furthermore, the model has been
validated by different experimental data and shows low deviation
with the published data. This model can predict the CO2 absorption
by PZ + MDEA solution accurately especially in the higher pressure
conditions owing to the application of PC-SAFT equation of state.

Declaration of Competing Interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Fig. A3. Variation of Henry’s constant of CO2 in the absorber with various PZ mass
Acknowledgement
concentrations at temperature of 48 ◦ C.

This work was financially supported by the National Natural


Science Foundation of China (No. 21676008).

Appendix A

Figs. A1 and A2 show the Gibbs free energy of PZ and MDEA as


solvents used in this model as well as ideal gas heat capacity of
CO2 and solution components at temperature of 48 ◦ C. In addition,
Fig. A3 demonstrates Henry’s constants of CO2 at temperature of

Fig. A4. Variation of CO2 mole fraction in the absorber with various PZ mass con-
centrations.

Table A1
The correlation parameters of characteristic volume in Brelvi-O’Connell model (Yan
and Chen, 2010; Zhang and Chen, 2011a).

Characteristic Volume (m3 /kmol)


Fig. A1. Variation of Gibbs free energy of PZ and MDEA in the absorber at tempera-
Component
ture of 48.◦ C Parameter
H2 O CO2 MDEA PZ

v1,i 4.640 × 10−2 1.750 × 10−1 3.690 × 10−1 3.100 × 10−1


v2,i 0.000 −3.380 × 10-4 0.000 0.000

48 ◦ C which are estimated by Aspen Plus based on the mentioned


correlations and input data. Fig. A4 depicts the variation of CO2 mole
fraction in the absorber with the flow rate of 72.0 and 250.0 kg/h
for gas and liquid streams respectively and the same conditions in
Section 3.
Table A1 shows the parameters of characteristic volume used
in Eq. (12), Table A2 indicates the parameters of Henry’s constant
in Eq. (16), Table A3 illustrates dielectric constants in Eq. (26)
Fig. A2. Variation of ideal gas heat capacity of CO2 and solvents in the absorber at
and Table A4 represents the default values of electrolyte-molecule
temperature of 48 ◦ C. ternary parameters for E-NRTL model based on Eqs. (37) and (38).

Table A2
The parameters of Henry’s constants for vapor-liquid equilibrium (Yan and Chen, 2010; Zhang and Chen, 2011a).

Component i CO2 CO2 CO2 N2 O2 PZ MDEA

Component j H2 O MDEA PZ H2 O H2 O H2 O H2 O
aij 1.006 × 102 1.989 × 10 1.006 × 102 1.803 × 102 1.579 × 102 1.372 × 103 1.370 × 102
bij −6.147 × 103 −1.072 × 103 −6.147 × 103 −1.518 × 104 −1.399 × 104 −6.390 × 104 −1.674 × 104
cij −1.019 × 10 0.000 −1.019 × 10 2.155 × 10 −1.839 × 10 −2.184 × 102 −1.758 × 10
dij −1.000 × 10-2 0.000 −1.000 × 10-2 −4.700 × 10-3 −5.200 × 10-3 2.190 × 10−1 0.000
T (K) 273−473 0−2000 273−473 273−346 274−348 386−472 298−403
A. Esmaeili et al. / Process Safety and Environmental Protection 141 (2020) 95–109 109

Table A3 Huang, J., Gong, M., Dong, X., Li, X., Wu, J., 2015. Study of CO2 solubility in aqueous
Dielectric constants of solvents (Dash and Bandyopadhyay, 2016). solutions of N-methyldiethanolamine + piperazine by electrolyte NRTL model.
Sci. China Chem. 59 (3), 360–369, http://dx.doi.org/10.1007/s11426-015-5508-
Parameters 5.
Component
Jakobsen, J.P., Krane, J., Svendsen, H.F., 2005. Liquid-phase composition determina-
Ai Bi
tion in CO2 -H2 O-Alkanolamine systems: an NMR study. Ind. Eng. Chem. Res. 44,
MDEA 2.199 × 10 8.992 × 103 9894–9903, http://dx.doi.org/10.1021/ie048813.
PZ 4.253 1.532 × 103 Kamps, A.P., Xia, J., Maurer, G., 2003. Solubility of CO2 in (H2 O+Piperazine) and
H2 O 7.865 × 10 3.199 × 104 in (H2O+MDEA+ Piperazine). AIChE J. 49 (10), 2662–2670, http://dx.doi.org/10.
1002/aic.690491019.
Khoshsima, A., Shahriari, R., 2017. Modeling study of the phase behavior of mixtures
Table A4 containing non-ionic glycol ether surfactant. J. Mol. Liq. 230, 529–541, http://
dx.doi.org/10.1016/j.molliq.2017.01.058.
The default value of electrolyte-molecule ternary parameters for E-NRTL model.
Khoshsima, A., Shahriari, R., 2018. Molecular modeling of systems related to the
Parameter Cm,ca Dm,ca Cca,m Dca,m biodiesel production using the PHSC equation of state. Fluid Phase Equilib. 458,
58–83, http://dx.doi.org/10.1016/j.fluid.2017.10.029.
(water-ion pair) 8.045 0.000 −4.072 0.000 Lin, C., Lin, Y., Tan, C., 2010. Evaluation of alkanolamine solutions for carbon dioxide
(CO2 -ion pair) 1.500 × 10 0.000 −8.000 0.000 removal in cross-flow rotating packed beds. J. Hazard. Mater. 175, 344–351,
(molecule-ion pair) 1.000 × 10 0.000 −2.000 0.000 http://dx.doi.org/10.1016/j.jhazmat.2009.10.009.
Mudhasakul, S., Ku, H., Douglas, P.L., 2013. A simulation model of a CO2 absorption
process with methyldiethanolamine solvent and piperazine as an activator. Int.
J. Greenhouse Gas Control. 15, 134–141, http://dx.doi.org/10.1016/j.ijggc.2013.
References
01.023.
Najibi, H., Maleki, N., 2013. Equilibrium solubility of carbon dioxide in
Ali, B.S., Aroua, M.K., 2004. Effect of piperazine on CO2 loading in aqueous solutions N-methyldiethanolamine + piperazine aqueous solution: experimental mea-
of MDEA at low pressure. Int. J. Thermophys. 25, 1863–1870, http://dx.doi.org/ surement and prediction. Fluid Phase Equilib. 354, 298–303, http://dx.doi.org/
10.1007/s10765-004-7740-7. 10.1016/j.fluid.2013.06.022.
Ali Khan, A., Halder, G.N., Saha, A.K., 2017. Experimental investigation on Nascimento, F.P., Paredes, M.L., Bernardes, A.P., Pessoa, F.L., 2019. Phase behavior
efficient carbon dioxide capture using piperazine (PZ) activated aqueous of CO2 /toluene, CO2 /n-decane and CO2 /toluene/n-decane: experimental mea-
methyldiethanolamine (MDEA) solution in packed column. Int. J. Greenhouse surements and thermodynamic modeling with SAFT-VR Mie equation of state. J.
Gas Control. 64, 163–173, http://dx.doi.org/10.1016/j.ijggc.2017.07.016. Supercrit. Fluids 154, 104634, http://dx.doi.org/10.1016/j.supflu.2019.104634.
Billet, R., Schultes, M., 1993. Predicting mass transfer in packed columns. Chem. Eng. Notz, R., Mangalapally, H.P., Hasse, H., 2012. Post combustion CO2 capture by reactive
Technol. 16, 1–9, http://dx.doi.org/10.1002/ceat.270160102. absorption: pilot plant description and results of systematic studies with MEA.
Bishnoi, S., Rochelle, G.T., 2000. Physical and chemical solubility of carbon dioxide Int. J. Greenhouse Gas Control 6, 84–112, http://dx.doi.org/10.1016/j.ijggc.2011.
in aqueous methyldiethanolamine. Fluid Phase Equilib. 168, 241–258, http://dx. 11.004.
doi.org/10.1016/S0378-3812(00)00303-4. Onda, K., Takeuchi, H., Okumoto, Y., 1968. Mass transfer coefficients between gas
Bishnoi, S., Rochelle, G.T., 2002a. Thermodynamics of piperazine / and liquid phases in packed columns. J. Chem. Eng. Jpn. 1 (1), 56–62, http://dx.
methyldiethanolamine / water / carbon dioxide. Ind. Eng. Chem. Res. 41, doi.org/10.1252/jcej.1.56.
604–612, http://dx.doi.org/10.1021/ie0103106. Rayer, A.V., Armugam, Y., Henni, A., Tontiwachwuthikul, P., 2014. High-pressure
Bishnoi, S., Rochelle, G.T., 2002b. Absorption of carbon dioxide in aqueous piper- solubility of carbon dioxide (CO2 ) in aqueous 1-methyl piperazine solution. J.
azine / methyldiethanolamine. AIChE J. 48 (12), 2788–2799, http://dx.doi.org/ Chem. Eng. Data 59, 3610–3623, http://dx.doi.org/10.1021/je500526m.
10.1002/aic.690481208. Razavi, E., Khoshsima, A., Shahriari, R., 2019. Phase behavior modeling of mixtures
Bravo, J.L., Fair, J.R., 1982. Generalized correlation for mass transfer in packed dis- containing N-, S-, and O-Heterocyclic compounds using PC-SAFT equation of
tillation columns. Ind. Eng. Chem. Process Des. Dev. 21, 162–170, http://dx.doi. state. Ind. Eng. Chem. Res. 58, 11038–11059, http://dx.doi.org/10.1021/acs.iecr.
org/10.1021/i200016a028. 9b01429.
Chen, C., Britt, H.I., Boston, J.F., Evans, L.B., 1982. Local composition model for excess Rodriguez, G., Beckman, E.J., 2020. Modelling phase behavior of triglycerides, diglyc-
Gibbs energy of electrolyte systems, Part I: single solvent, single completely erides and monoglycerides related to biodiesel transesterification in mixtures
dissociated electrolyte systems. AIChE J. 28 (4), 588–596, http://dx.doi.org/10. of alcohols and CO2 using a polar version of PC-SAFT. Fluid Phase Equilib. 503,
1002/aic.690280410. 112303, http://dx.doi.org/10.1016/j.fluid.2019.112303.
Chen, Y., Mutelet, F., Jaubert, J.N., 2012. Modeling the solubility of carbon dioxide Samanta, A., Roy, S., Bandyopadhyay, S.S., 2007. Physical solubility and diffusivity of
in imidazolium-based ionic liquids with the PC-SAFT equation of state. J. Phys. N2 O in aqueous solutions of piperazine and (n-methyldiethanolamine + piper-
Chem. B 116, 14375–14388, http://dx.doi.org/10.1021/jp309944t. azine). J. Chem. Eng. Data 52, 1381–1385, http://dx.doi.org/10.1021/je700083b.
Cullinane, J.T., Rochelle, G.T., 2005. Thermodynamics of aqueous potassium carbon- Shahriari, R., Dehghani, M.R., Behzadi, B., 2012. A modified polar PHSC model for
ate, piperazine, and carbon dioxide. Fluid Phase Equilib. 227, 197–213, http:// thermodynamic modeling of gas solubility in ionic liquids. Fluid Phase Equilib.
dx.doi.org/10.1016/j.fluid.2004.11.011. 313, 60–72, http://dx.doi.org/10.1016/j.fluid.2011.09.029.
Dash, S.K., Bandyopadhyay, S.S., 2016. Studies on the effect of addition of piperazine Sodeifian, G., Razmimanesh, F., Sajadian, S.A., Soltanipanah, H., 2018. Solubility mea-
and sulfolane into aqueous solution of N-methyldiethanolamine for CO2 capture surement of an antihistamine drug (loratadine) in supercritical carbon dioxide:
and VLE modelling using eNRTL equation. Int. J. Greenhouse Gas Control 44, assessment of qCPA and PCP-SAFT equations of state. Fluid Phase Equilib. 472,
227–237, http://dx.doi.org/10.1016/j.ijggc.2015.11.007. 147–159, http://dx.doi.org/10.1016/j.fluid.2018.05.018.
Dash, S.K., Samanta, A., Samanta, A.N., Bandyopadhyay, S.S., 2011. Vapor liquid Versteeg, G.F., van Swaaij, P.M., 1988. Solubility and diffusivity of acid gases (CO2 ,
equilibria of carbon dioxide in dilute and concentrated aqueous solutions of N2 O) in aqueous alkanolamine solutions. J. Chem. Eng. Data 33, 29–34, http://
piperazine at low to high pressure. Fluid Phase Equilib. 300, 145–154, http://dx. dx.doi.org/10.1021/je00051a011.
doi.org/10.1016/j.fluid.2010.11.004. Xia, J., Kamps, A.P., Maurer, G., 2003. Solubility of H2 S in (H2 O + Piperazine) and in
Derks, P.W.J., Dijkstra, H.B.S., Hogendoorn, J.A., Versteeg, G.F., 2005. Solubility of (H2 O + MDEA + Piperazine). Fluid Phase Equilib. 207, 23–34, http://dx.doi.org/
carbon dioxide in aqueous piperazine solutions. AIChE J. 51 (8), 2311–2327, 10.1016/S0378-3812(02)00331-X.
http://dx.doi.org/10.1002/aic.10442. Xu, Q., 2011. Thermodynamics of CO2 loaded aqueous amines. In: Ph.D. Dissertation.
Derks, P.W.J., Hamborg, E.S., Hogendoorn, J.A., Niederer, J.P.M., Versteeg, G.F., University of Texas.
2008. Densities, viscosities, liquid diffusivities in aqueous piperazine and aque- Yan, Y., Chen, C., 2010. Thermodynamic modeling of CO2 solubility in aqueous solu-
ous (piperazine + N-methyldiethanolamine) solutions. J. Chem. Eng. Data 53, tions of NaCl and Na2 SO4 . J. Supercrit. Fluids 55, 623–634, http://dx.doi.org/10.
1179–1185, http://dx.doi.org/10.1021/je800031p. 1016/j.supflu.2010.09.039.
Ermatchkov, V., Maurer, G., 2011. Solubility of carbon dioxide in aqueous solutons Zhang, Y., Chen, C., 2011a. Thermodynamic modeling for CO2 absorption in aqueous
of N-methyldiethanolamine and piperazine: prediction and correlation. Fluid MDEA solution with electrolyte NRTL model. Ind. Eng. Chem. Res. 50, 163–175,
Phase Equilib. 302, 338–346, http://dx.doi.org/10.1016/j.fluid.2010.06.001. http://dx.doi.org/10.1021/ie1006855.
Gross, J., Sadowski, G., 2002. Application of the Perturbed-Chain SAFT equation of Zhang, Y., Chen, C., 2011b. Modeling gas solubilities in the aqueous solution of
state to associating systems. Ind. Eng. Chem. Res. 41, 5510–5515, http://dx.doi. methyldiethanolamine. Ind. Eng. Chem. Res. 50, 6436–6446, http://dx.doi.org/
org/10.1021/ie010954d. 10.1021/ie102150h.
Hemmati, A., Farahzad, R., Surendar, A., Aminahmadi, B., 2019. Validation of Zhang, X., Wang, J., Zhang, C., Yang, Y., Xu, J., 2003. Absorption rate into a MDEA
mass transfer and liquid holdup correlations for CO2 absorption process with aqueous solution blended with piperazine under a high CO2 partial pressure.
methyldiethanolamine solvent and piperazine as an activator. Process Saf. Env- Ind. Eng. Chem. Res. 42, 118–122, http://dx.doi.org/10.1021/ie020223t.
iron. Prot. 126, 214–222, http://dx.doi.org/10.1016/j.psep.2019.02.030.
Hilliard, M.D., 2008. A predictive thermodynamic model for an aqueous blend of
potassium carbonate, piperazine, and monoethanolamine for carbon dioxide
capture from flue gas. In: Ph.D. Dissertation. University of Texas.

You might also like