International Journal of Greenhouse Gas Control: Sukanta Kumar Dash, Syamalendu S. Bandyopadhyay

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

International Journal of Greenhouse Gas Control 44 (2016) 227–237

Contents lists available at ScienceDirect

International Journal of Greenhouse Gas Control


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

Studies on the effect of addition of piperazine and sulfolane into


aqueous solution of N-methyldiethanolamine for CO2 capture and VLE
modelling using eNRTL equation
Sukanta Kumar Dash a , Syamalendu S. Bandyopadhyay b,∗
a
Department of Chemical Engineering, Pandit Deendayal Petroleum University, Gandhinagar 382007, Gujarat, India
b
Cryogenic Engineering Centre, Indian Institute of Technology Kharagpur, Kharagpur 721302, WB, India

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

Article history: In this work, new experimental vapour liquid equilibrium (VLE) data and modelling work of VLE of CO2
Received 7 May 2015 in piperazine (PZ) activated aqueous solutions of n-methyldiethylamine (MDEA) are presented. Besides,
Received in revised form 11 August 2015 experimental VLE data of CO2 in a hybrid solvent, containing sulfolane as physical solvent along with
Accepted 8 November 2015
aqueous (MDEA + PZ), have also been reported over the temperature range of 308–328 K and CO2 partial
Available online 8 December 2015
pressure range of 1–1400 kPa. The concentrations of MDEA in the aqueous solutions are 22–28 mass%
and also 42–48 mass%. PZ is used as a rate activator within a concentration range of 2–8 mass%. The total
Keywords:
amine concentration in the aqueous solutions of (MDEA + PZ) has been kept within 30 mass% and also
CO2 capture
VLE modelling
50 mass%, while for the hybrid solvent the concentration of sulfolane has been maintained at 10 mass%
Piperazine along with total 50 mass% (MDEA + PZ). Electrolyte non random-two liquid (eNRTL) theory has been used
N-methyldiethanolamine to model the VLE. It has been found that there is a good agreement between the experimental and model
Sulfolane results of CO2 solubility in PZ activated MDEA solutions. The CO2 cycle capacity is calculated considering
a set of lower and higher CO2 partial pressures corresponding to lower (lean solvent) and higher (rich
solvent) equilibrium CO2 loadings. For the sulfolane based hybrid solvent, it has been observed that the
CO2 cyclic capacity of the solvent decreases with the lower lean-rich partial pressure range of 10–100 kPa,
but increases at higher partial pressure ranges of 100–1000 kPa at 323 K. PZ mass% in the hybrid solvent
has also been found to have a positive effect on the CO2 cyclic capacity.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction regenerative absorption of CO2 into aqueous solutions of alka-


nolamines. Commercially important alkanolamines used for gas
Removal of acid gas impurities, such as carbon dioxide (CO2 ) treating purpose are monoethanolamine (MEA), diethanolamine
and hydrogen sulfide (H2 S), from natural gas, refinery off-gases, (DEA), N-methyldiethanolamine (MDEA), 2-amino-2-methyl-1-
hydrogen gas stream and synthesis gas for ammonia produc- propanol (AMP), etc. (Kohl and Nielsen, 1997). Aqueous solutions
tion is an important operation in industrial gas processing. In of MDEA along with a rate activator piperazine (PZ) have been suc-
the recent times, growing concentration of CO2 in the earth’s cessfully used for years in the natural gas industry for removal of
atmosphere leading to global warming and climate change is CO2 and/or hydrogen sulfide (H2 S) (Bishnoi and Rochelle, 2002).
also of serious concern necessitating the need for large scale Due to certain advantages of (MDEA + PZ) solvent, it is the focus
CO2 separation from various flue gas streams. This gives rise to of many recent research activities on gas treating (Zoghi et al.,
extensive research activities for developing more efficient and 2012). As a tertiary amine, MDEA does not form a carbamate and
improved processes for economic CO2 capture from large point it does not easily undergo amine polymerization like the primary
sources of CO2 such as fossil fuel based power plants for safe and secondary amines (Closmann et al., 2009). Both MDEA and PZ
CO2 sequestration. The widely used processes for the removal are more resistant to thermal and oxidative degradation (Freeman
of CO2 from natural gas and industrial gas streams are the et al., 2010). Besides, MDEA is much less corrosive (Jenab et al.,
2005). The heat of absorption of this mixed solvent (MDEA + PZ)
at high CO2 loading is low (Svensson et al., 2013), hence solvent
∗ Corresponding author.
regeneration energy is reasonable for the CO2 capture operation.
E-mail addresses: sk.dash@sot.pdpu.ac.in, ssbandyo@hijli.iitkgp.ernet.in
There is also a growing interest in using hybrid solvents to take the
(S.S. Bandyopadhyay). advantage of both physical and chemical absorption of CO2 . The

http://dx.doi.org/10.1016/j.ijggc.2015.11.007
1750-5836/© 2015 Elsevier Ltd. All rights reserved.
228 S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237

use of a physical solvent, e.g., sulfolane in varying concentration, thermodynamics and VLE have been modelled by some workers
has been developed to take advantage of the VLE of each class of (Tong et al., 2012; Zong and Chen, 2011). The modelling approach
solvents (Wang et al., 2010). Sulfolane has high physical absorption of this work is similar to those reported by Zong and Chen (2011).
capacity for CO2 and H2 S, hence used in physical and hybrid solvent Our thermodynamic model is based on eNRTL theory which has
processes (e.g., sulfinol-M process) (Angaji et al., 2013). The physi- been developed by regressing experimental data. Aspen Plus®
cal characteristics of the aqueous sulfolane–amine solution make it (2011) platform has been chosen for data regression and model
possible to tune the hybrid solvent by varying the relative amounts validation. Further, use of this model will lead to a paramet-
of water and sulfolane in order to reduce the solvent regeneration ric study of CO2 capture process using absorption-regeneration
energy. In view of this, sulfolane has been chosen as one of the column model as described in our previous work (Dash et al.,
solvents of the hybrid solvent system in this study. The mixing of 2014).
the chemical solvent, MDEA, and the physical solvent, sulfolane,
is expected to increase the CO2 loading capacity, while PZ will act
as a rate activator. In this work, a higher total amine concentra- 2. Experimental
tion in the aqueous (MDEA + PZ) solvents has been taken in order
to increase the cyclic loading capacity of the solvent giving rise to 2.1. Materials
reduced solvent circulation rate. Higher amine concentration along
with sulfolane reduces the quantity of water in the solvent and thus Reagent grade MDEA (>95% pure), PZ (>99%) and sulfolane (CAS:
reduces the regeneration heat duty. 126-33-0) (>99%) were obtained from Sigma–Aldrich, USA, and was
In the literature, a good deal of solubility data of CO2 in used without further purification. Carbon dioxide (>99.9% pure)
(MDEA + PZ) is available (Ermatchkov and Maurer, 2011). Xu et al. was obtained from Chemtron Science Pvt. Ltd., India. Double dis-
(1998) presented the effect of the addition of PZ on equilibrium CO2 tilled water degassed by prolonged boiling and cooled to ambient
partial pressure and CO2 loading in aqueous (1.75–4.28 mol dm−3 temperature under air tight condition was used to prepare aqueous
MDEA + 0.041–0.021 mol dm−3 PZ). Liu et al. (1999) presented solutions of (MDEA + PZ) and (MDEA + PZ + sulfolane). A digital bal-
experimental solubility data of CO2 over aqueous (MDEA + PZ) by ance (Acculab ALC 210.4, accuracy of ±0.0001 g) was used to weigh
changing relative concentration of MDEA and PZ and presented a PZ for solution preparation.
simple VLE model. Their VLE model did not include any of the PZ
carbamate species such as PZ-carbamate, PZ-dicarbamate, proton-
ated PZ-carbamate. Bishnoi and Rochelle (2002) investigated CO2 2.2. Apparatus
solubility in 4.0 mol dm−3 MDEA activated with 0.6 mol dm−3 PZ at
temperatures of 313 and 343 K and CO2 partial pressure up to about The VLE measurements were done in a high pressure stainless
7.5 kPa. They developed an eNRTL model using experimental sol- steel stirred equilibrium cell, which is connected to a high pressure
ubility and NMR data. Kamps et al. (2003) reported total pressure stainless steel gas buffer vessel. A detail description of the experi-
data and VLE model of CO2 in aqueous solution of 2 mol (kg water)−1 mental set up, procedure and validation of experimental procedure
MDEA + 2 mol (kg water)−1 PZ at 353 K and 0.18–6.4 MPa total pres- are presented by Dash et al. (2011a). A magnetic stirrer is used
sure. However, it was found by Derks et al. (2010) that their model for liquid phase agitation in the equilibrium cell. Both the cell and
did not accurately describe the CO2 solubility data below 100 kPa. the gas buffer are equipped with platinum temperature sensors
Recently, Najibi and Maleki (2013) reported high temperature VLE (Pt 100, JULABO, FRG) and PDCR 4010 (Druck, UK) pressure trans-
data of CO2 in aqueous (MDEA + PZ). A summary of literature review ducers (pressure range 0–140, 0–700, 0–1400, 0–2000 kPa with
on CO2 − (MDEA + PZ + H2 O + sulfolane) is presented in Table 1. accuracy of ±0.75% of full-scale), with digital pressure indicators
Although there are reasonable VLE data of CO2 in (MDEA + PZ) (DPI 280, Druck, UK). The temperatures of the equilibrium cell and
available in the literature, data on CO2 solubility in aque- buffer vessel were controlled within ±0.1 K of the desired temper-
ous (MDEA + PZ + sulfolane) are scarce. Jenab et al. (2005) ature by a circulator temperature controller (F 32, JULABO, FRG).
presented solubility of CO2 in aqueous (1.68–3 mol dm−3 Pressure of the equilibrium cell is controlled by precision pressure
MDEA + 0.36–1.36 mol dm−3 PZ + 0.36–1.36 mol dm−3 sulfolane) controllers (5866 series, Brooks Instruments, USA, pressure range
at temperature and CO2 partial pressure range of 313–343 K 0–140, 100–700, 700–1400 kPa), within an accuracy of ±1% of full
and 30–3900 kPa, respectively. However, no thermodynamic scale. A vacuum pump (Indvac Model IV 50, India) was employed
model for this system is available in the open literature. Reliable to evacuate the equilibrium cell and buffer vessel, and degas the
experimental data and improved thermodynamic models are amine solutions.
still necessary to explain the equilibrium behaviour of CO2 in For each run, the equilibrium cell and buffer vessel were allowed
aqueous (MDEA + PZ + sulfolane) for a wide range of temperature to reach the desired temperature for VLE measurements. A known
and pressure. volume (25 × 10−6 m3 ) of the mixed solvent was introduced into
For consistent design of the absorption-regeneration process the cell and the solution was then degassed so that the liquid exists
for CO2 separation, understanding of CO2 -amine solvent VLE and under its own vapour pressure. Once the temperature was uniform
thermodynamic VLE model is needed (Puxty and Maeder, 2013). throughout, the pressure of the cell (Pv ) and buffer vessel (Pt ) were
Because of chemical reactions of CO2 with aqueous (MDEA + PZ) recorded. The pressure of the equilibrium cell was then set to the
and due to the presence of PZ-carbamate and carbonate ions desired value with the pressure controller taken in line. The attain-
in the liquid phase, the liquid phase is strongly non-ideal. ment of equilibrium was ensured when there was no change in the
Hence, the thermodynamic modelling of the systems containing pressure of the equilibrium cell and the buffer vessel for at least one
(MDEA + PZ + sulfolane + CO2 ) is a difficult task. Accurate experi- hour at constant temperature. The difference between the initial
mental data for the solubility of CO2 in these solvents are required and final pressures in the buffer, and the equilibrium CO2 partial
to develop and test thermodynamic models used to describe pressure pCO2 (= Pt − Pv ) was used to calculate the corresponding
VLE of these systems (Puxty and Maeder, 2013). But, reliable CO2 loading, ˛CO2 in the liquid phase. Peng–Robinson equation of
data of these systems are often scarce. In this work, the solu- state was used for calculating the compressibility factor of CO2 at
bilities of CO2 in aqueous solutions containing (MDEA + PZ) and the corresponding temperature and pressure. ˛CO2 for both aque-
(MDEA + PZ + sulfolane) have been measured for a wide range of ous (MDEA + PZ) and (MDEA + PZ + sulfolane) was calculated using
concentrations and temperatures. Recently CO2 -aqueous amine Eq. (1).
S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237 229

Table 1
Summary of literature on CO2 solubility in (MDEA + PZ + sulfolane + H2 O) system.

Reference T (K) Solvent Concentration and the pCO2 (kPa)


ratios (mol dm−3 , M) or
(mol kg water−1 , m)

Liu et al. (1999) 323.15–363.15 MDEA + PZ MDEA: (1.35–4.77) M 13.16–935.3


PZ: (0.17–1.55) M
Jenab et al. (2005) 313.15–343.15 MDEA + PZ MDEA: 2–3 M 27.79–3938.43
PZ:0.36–1.36 M
Vahidi et al. (2009) 313.15–343.15 MDEA + PZ MDEA: (2–3) M 27.8–3938.4
PZ: (0.36–1.36) M
Bottger et al. (2009) 313.3–393.2 MDEA + PZ MDEA: 218–11880
(2.227–7.831) m
PZ: (1.966–2.072) m
Derks et al. (2010) 298–323 MDEA + PZ MDEA: (0.61–4.49) M 0.25–110
PZ: (1.12–1.835) M
Speyer et al. (2010) 313.15–393.15 MDEA + PZ MDEA: 0.11–99.8
(2.013–7.936) m
PZ: (1.024–4.099) m
Najibi and Maleki (2013) 363.15–423.15 MDEA + PZ MDEA: (1.6–3) M 27.3–204
PZ: (0.3–0.7) M
Murrieta-Guevara et al. (1988) 303.15–373.15 Sulfolane 100 wt% 81.2–2263.4
Jou et al. (1990) 298.15–403.15 Sulfolane 100 wt% 15.7–7590
Wang et al. (2010) 313.2–373.2 AMP + sulfolane AMP: 8.2–30.6 wt% 2.23–193
Sulfolane:
19.4–41.2 wt%
Jenab et al. (2005) 313.15–343.15 MDEA + sulfolane MDEA: (2–3) M 38.2–3151.06
313.15–343.15 MDEA + PZ + sulfolane PZ: (0.36–1.36) M 30.20–3154.51
MDEA: (1.68–2.5) M
PZ: (0.43–0.84) M
Sulfolane:
(0.43–0.84) M

g
nCO2 − nCO K8
˛CO2 = 2
(1) MDEA + H2 OMDEAH+ + OH− (R8)
n(MDEA+PZ)
g Sulfolane does not undergo any reaction in carbonated aqueous
where (nCO2 − nCO ) is the no. of moles of CO2 present in amine solu-
2 medium. The only reacting species are carbonate and amine ions in
tion and n(MDEA+PZ) is the total moles of amine (MDEA + PZ) present aqueous phase and their concentrations are governed by thermo-
in the liquid phase. The uncertainty of loading, ˛CO2 , is deter- dynamic equilibrium constant, Kj , defined in terms of mole fraction,
mined from the uncertainties in temperature, pressure, volume and xi , and activity coefficient,  i , of the ith component as,
molarity which are ±0.1 K, ±0.75%, ±0.5% and ±0.2%, respectively.
After analyzing possible systematic and random errors, the esti- 
mated uncertainty in the measured solubility with respect to CO2 Kj = (xi i )ij (2)
loading and pCO2 are about ±3% and ±2%, respectively. i

where ij is the reaction stoichiometric coefficient of component i in


3. Modelling jth reaction. Dash et al. (2011b) developed thermodynamic model
for (AMP + PZ + H2 O + CO2 ) system. Same equilibrium constants
3.1. Chemical and phase equilibria from Dash et al. (2011b) were used here. The values of these equilib-
rium constants are listed in Table 2. Equilibrium constant for MDEA
Model development of the chemical and phase equilibria protonation is taken from Aspen data bank (Aspen, 2011). In this
of the system (CO2 + PZ + H2 O) has been discussed by Dash work, water, MDEA and sulfolane are considered as solvents lead-
et al. (2011a). The models for (CO2 + MDEA + PZ + H2 O) and ing to a mixed solvent system. The standard state associated with
(CO2 + MDEA + PZ + sulfolane + H2 O) are developed by considering each solvent is the pure liquid at the system temperature and pres-
the following reactions in aqueous phase. sure. The chosen reference state for ionic solutes and the molecular
K1 solute (CO2 ) is the ideal, infinitely dilute aqueous solution at the
2H2 OH3 O+ + OH− (R1) system temperature and pressure.
K2
CO2 + 2H2 OH3 O+ + HCO3 − (R2)
3.2. Vapour phase model
K3
− + 2−
HCO3 + H2 OH3 O + CO3 (R3)
When phase equilibrium is established, fugacity of each compo-
K4 nent in the liquid and vapour phases is same. At this equilibrium
+ +
PZH + H2 EPZ + H3 O (R4) condition, CO2 molecules distribute themselves between the liquid
K5 and vapour phase according to Eq. (3), whereas the solvent species
PZ + CO2 + H2 OPZCOO− + H3 O+ (R5) such as water, MDEA, sulfolane and PZ follow Eq. (4) according to
the activity coefficient approach (Poling et al., 2001).
K6
H+ PZCOO− + H2 OPZCOO− + H3 O+ (R6)  
v∞ 0 )
(P − Psolv
K7 iv yi P = xi i∗ Hi exp i
(3)
PZCOO− + CO2 + H2 OPZ(COO− )2 + H3 O+ (R7) RT
230 S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237

Table 2
Coefficients for the chemical equilibrium constants used in the eNRTL model.

a b c d T (K)

K1 132.899 −13,445.9 −22.477 0.0 273–498


K2 231.465 −12,092.1 −36.782 0.0 273–498
K3 216.049 −12,431.7 −35.482 0.0 273–498
K4 514.314 −34,910.9 −74.602 0.0 273–323
K5 466.497 1614.5 −97.540 0.2471 273–343
K6 6.822 −6066.9 −2.290 0.0036 273–343
K7 −11.563 1769.4 −1.467 0.0024 273–343
K8 −3.68672 −6754.686 0.0 0.0 273–328

Table 3
Henry’s law constant: ln HiA (T, PA∗,l ) = aiA + (biA /T ) + ciA ln T + diA T (kPa).

a b c d

HCO2 –H2 O 163.805 −8477.71 −21.9574 0.005781


HCO2 –MDEA 1048 −12,160.1 0 −0.0919986
HCO2 –sulfolane 85.1206 −4652.9 −9.4077 0

 
vi (P − Pi0 ) The Born correction term for the excess Gibbs energy is given by:
iv yi P = xi i Pi0 exp (4)  
RT
G∗E,Born Qe2 1 1
 x z2
i i
= − 10−2 (9)
RT 2kT εs εw ri
where yi and xi are concentration of species in vapour and liquid i
phases, respectively, v∞ i
, partial molar volume of CO2 at infi-
nite dilution, can be calculated from the Brelvi-O’Connell method. Here dielectric constant of PZ is assumed to be the same as that
Hi is the Henry’s law constant, P0 is the vapour pressure. The of piperdine (Dash et al., 2011c). The mixed solvent dielectric con-
coefficients of Henry’s law constant for CO2 in water are taken stant, εs , is calculated by a simple mass fraction average. Dielectric
from DIPPER data base. Sulfolane vapour pressure is taken from constant data of sulfolane and MDEA are taken from Vahidi and
Fulem et al. (2011). The partial molar volume of CO2 is obtained Moshtari (2013) and listed in Table 4. The NRTL expression for the
from Zhang and Chen (2011). Experimental results for solubility short range interactions is given as:

of N2 O in pure MDEA from Derks et al. (2005) have been used to


G∗E,lc  XG
j j jm jm
   Xa 
estimate the physical solubility of CO2 in pure MDEA and a corre- = Xm
+ Xc

RT X G Xa
lation has been developed for Henry’s constant of CO2 in MDEA. m k k km c a a
However, the Henry’s constants for CO2 in aqueous (MDEA + PZ)

and (MDEA + PZ + sulfolane) have been adjusted in the model to XG  


j j jc,a c jc,a c
 
×
+ Xa
match the experimental VLE data within a reasonable accuracy and X G 
k k kc,a c a c
presented in Table 3. In this work, Redlich–Kwong–Soave equa-
 

tion of state has been chosen to compute vapour phase fugacity Xc XG  
j j ja,c a ja,c a
coefficients, ϕiv , and eNRTL theory is adopted to calculate liquid ×

(10)
c 
Xc X G 
k k ka,c a
phase activity coefficients.
where j and k can be any species. The definitions of all the terms
were given by Dash et al. (2011a). Only a few terms are defined
3.3. Electrolyte NRTL activity coefficient model
here.

The eNRTL equation representing the thermodynamics of CO2 - Gjc,a c = exp(−˛jc,a c jc,a c ) (11)
alkanolamine has been previously developed by Dash et al. (2011c).
Only the major equations for model development are discussed Gja,c a = exp(−˛ja,c a ja,c a ) (12)
here. According to this model, the excess Gibbs energy is calculated
ln Gcm
by Eqs. (5) and (6): cm = − (13)
˛cm
G∗E G∗E,PDH G∗E,Born G∗E,lc ln Gam
= + + (5) am = − (14)
RT RT RT RT ˛cm
 
G∗E,PDH   1000 1/2  4A I  where Xj = xj ·Cj (Cj = zj for ions and 1 for molecules), ˛ij is the
 x 1/2
=− xk ln(1 + Ix ) (6) nonrandomness factor and ij is the binary energy interac-
RT Ms 
k tion parameter. The activity coefficient for any species (ionic or

where Debye–Hückel parameter, A , and ionic strength of solvent,


Table 4
Ix , are given by Eqs. (7) and (8), respectively. Dielectric constants of MDEA, PZ, sulfolane and water.

1
 2N d 1/2  Q 2 3/2 Dielectric constant ␧ = a1 + b1 /T [1/T − 1/298.15]
A s e
A = (7) Component a1 b1
3 1000 εs kT
H2 O 78.65 31,989
1 2 PZ 4.253 1532.2
Ix = xi zi (8) MDEA 21.9957 8992.68
2 Sulfolane 44.1283 12,703.9
i
S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237 231

molecular, solute or solvent) is calculated from the partial deriva-


tive of the excess Gibbs energy with respect to mole number as,

1 ∂(nt G∗E )
ln i = , i, j = m, c, a (15)
RT ∂ni
T,P,nj =
/ i

3.4. Model parameters

The model parameters of the VLE model for (CO2 + PZ + H2 O)


are presented by Dash et al. (2011a). Recently Zhang and
Chen (2011) presented eNRTL model parameters for the sys-
tem (CO2 + MDEA + H2 O). Zong and Chen (2011) presented eNRTL
model parameters of aqueous (CO2 + MDEA + sulfolane) sys-
tem. Regressed parameters from Zhang and Chen (2011) for
(CO2 + MDEA + H2 O) system and from Zong and Chen (2011) for
(CO2 + MDEA + sulfolane + H2 O) system have been used in this work.
Fig. 1. Comparison of model predicted and experimental equilibrium partial
Only the additional parameters for CO2 solubility in aqueous
pressures of CO2 over aqueous 30 mass% MDEA at various temperatures.
(MDEA + PZ) have been regressed. The pure component parameters
such as critical constants, acentric factor, compressibility factor,
Brelvi-O’Connell parameter, are listed in Table 5. Antoine equa-
tion constants for the vapour pressure of molecular species have
been used from Aspen data bank and are listed in Table 6. The
model predicted physical properties e.g., density, viscosity, specific
heat of pure and mixed solvents are compared with the experi-
mental data available in literature. The thermophysical property
data such as molar heat capacity and thermal conductivity from
Shokouhi et al. (2013) and Mundhwa et al. (2009); vapour pres-
sure of sulfolane and (sulfolane + water) from Fulem et al. (2011)
and Steele et al. (1997); excess molar volume data of sulfolane
and (sulfolane + MDEA) from Vahidi and Moshtari (2013); density,
viscosity and liquid diffusivities of (MDEA + PZ) from Derks et al.
(2008) and Fu et al. (2013) have been compared with the model
predictions. Other parameters such as non-randomness factors for
molecule–molecule and molecule–electrolyte, binary interaction
parameters for molecule–molecule and molecule–ion pair and ion
pair–ion pair, etc. are assigned their default values as in our previ-
Fig. 2. Comparison of model predicted and literature results of VLE of CO2 over
ous work (Dash et al., 2011a). A few molecule–ion pair parameters aqueous (MDEA + PZ).
have been chosen for regression considering the equilibrium com-
position in the liquid phase. Experimental data of this work and and aqueous (MDEA + PZ) from literature (Jou et al., 1982; Kundu
data from Speyer et al. (2010), Bottger et al. (2009) and Jenab et al. and Bandyopadhyay, 2005) respectively, with our eNRTL model.
(2005) have been entered in the data regression system (DRS® ) and Fig. 3 shows the equilibrium CO2 partial pressure against loading
regressed by following standard procedure. for different (MDEA + PZ) composition at 328 K. The figure indi-
The PZ–H2 O interaction parameters are taken from Dash et al. cates that increase of PZ mass% in (MDEA + PZ) solution, while
(2011a) and MDEA–H2 O interaction parameters are taken form keeping total amine mass% constant, increases the CO2 solubil-
Zhang and Chen (2011). The regressed parameters obtained in this ity up to a loading of 1. After CO2 loading of about 1, the trend
work and default values of other parameters used for molecule–ion reveres. This anomaly is also observed by Bishnoi and Rochelle
pair and ion pair–molecule interactions are listed in Table 7.

4. Results and discussion

4.1. Equilibrium CO2 partial pressure over aqueous (MDEA + PZ)

The new solubility data of CO2 in aqueous (MDEA + PZ) over


the temperature range of 308–328 K and CO2 partial pressure
range of 1–1400 kPa are presented in Tables 8 and 9. Aqueous
amine concentrations of (22 mass% MDEA + 8 mass% PZ), (25 mass%
MDEA + 5 mass% PZ), and (28 mass% MDEA + 2 mass% PZ) and
(42 mass % MDEA + 8 mass% PZ), (45 mass% MDEA + 5 mass% PZ) and
(48 mass% MDEA + 2 mass% PZ) have been taken for this work. Using
the regressed value of eNRTL interaction parameters, the VLE of CO2
in aqueous (MDEA + PZ) is predicted by Aspen Plus® flash model as
described in Dash et al. (2011a,b). The model predicted CO2 partial
pressures have been plotted against experimental results of this
work and from literature in Figs. 1–3. Figs. 1 and 2 show the com- Fig. 3. Effect of relative MDEA/PZ composition on CO2 solubility in aqueous
parisons of experimental CO2 solubility data over aqueous MDEA (MDEA + PZ).
232 S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237

Table 5
Pure component physical properties for VLE model.

Properties H2 O CO2 MDEA Sulfolane PZ

Tc (K) 647.3 304.2 675 853 638.0


Pc (kPa) 22,048 7376 3880 5030 6870
Vc (m3 /kmol) 0.0559 0.0939 0.368 0.3 0.3100
Acentric factor (ω) 0.344 0.225 1.1649 0.382349 0.4138
Racket ZRA 0.2432 0.2736 0.252575 0.23901 0.2000
Brelvi-O’Connell 0.0464 0.0939 0.3 0.3 0.3

Table 6
Antoine equation coefficients of molecular species and ions.

Components H2 O MDEA Sulfolane PZ CO2 Ions

A 65.6422 246.162 143.752 63.5952 133.632 −1.00E+20


B −7206.7 −18,378 −13,283 −7914.5 −4735 0
C 0 0 0 0 0 0
D 0 0 0 0 0 0
E −7.1385 −33.972 −19.429 −6.6461 −21.268 0
F 4.05E−06 2.3348E−05 0.013441 5.21E−18 0.040909 0
G 2 2 1 6 1 0
B G
ln(p0i /Pa) = A + (T /K)+C
+ D(T /K) + E ln(T /K) + F(T /K) .

(2002). It is also observed that as the temperature increases the choosing a better solvent, as suggested by Mathias et al. (2013)
CO2 solubility decreases for a particular total amine (MDEA + PZ) based on solution thermodynamics. CO2 capacity is defined as the
concentration. The CO2 loading increases with partial pressure and moles of CO2 absorbed per kg of solvent over a given lean-rich CO2
also with PZ concentration. It is evident that PZ activated MDEA partial pressure range. This can be calculated using Eq. (16) (Dash
shows improved VLE, e.g., at 328 K and CO2 partial pressure of about et al., 2012) assuming the rich and lean streams are in equilib-
1 kPa, the CO2 loading, ˛, is 0.1 for 30 mass% MDEA, 1.6 for 28 mass% rium with the inlet and outlet CO2 partial pressure in an absorption
MDEA + 2 mass% PZ, 0.23 for 25 mass% MDEA + 5 mass% PZ, and 0.3 column.
for 22 mass% MDEA + 8 mass% PZ. It is clear from Fig. 3 that the  mole CO
2
equilibrium CO2 loading capacity of aqueous (MDEA + PZ) can be capacity = (˛pCO − ˛pCO )[MDEA + PZ]
kg solvent 2 ,rich 2 ,lean
>1 especially at higher CO2 partial pressure. From the results pre-
 mole (MDEA + PZ)
sented in Figs. 1–3, it is clear that there is a good agreement between
× (16)
the model predicted results and experimental results. kg solvent

4.2. CO2 cyclic capacity Table 10 presents the estimated CO2 capacity of the solvent
calculated at different partial pressures of CO2 and an absorption
CO2 cyclic capacity is an important performance indicator of temperature of 313 and 323 K for total 50 mass% of (MDEA + PZ).
the solvent as it is required to calculate the solvent circulation rate It is observed that the solvent capacity increases with the PZ
in an absorption-regeneration process. This is also important for mass% at lower partial pressure range of 10–100 kPa but decreases
Table 7
Summary of interaction parameters obtained from data regression and default values.

Parameter Aij Aji Bij Bji

Binary molecule–molecule interaction parameters obtained using DRS


H2 O–MDEA 2.15908 −4.03445 −959.666 1661.4
H2 O–sulfolane 0.5356 −1.841 346.9 784.93

Interaction Am,ca Aca,m Interaction Am,ca Aca,m

Molecule–electrolyte pair parameters obtained using DRS


H2 O- 11.053 −4.4819 PZ- 9.9165 −2.016
[PZH+ , HCO3 − ] [PZH+ , PZ(COO− )2 ]
PZ- 9.9167 −2.016 H2 O- 7.9342 −4.032
[PZH+ HCO3 − ] [PZH+ , PZ(COO− )2 ]
H2 O- 7.9349 −4.032 PZ- 9.9165 −2.016
[PZH+ PZCOO− ] (H3O+ , HCO3 − )
PZ- 9.916 −2.016 PZ- 9.9165 −2.016
[PZH+ PZCOO− ] [H3 O+ , PZ(COO− )2 ]
H2 O- 8.13 −4.08 Sulfo- 4.4469 2.7734
[MDEAH+ PZ(COO− )2 ] [MDEA+ , HCO3 − ]
MDEA- 27.947 −6.79 H2 O- 26.1614 −14.3515
[PZH+ , HCO3 − ] [MDEAH+ , HCO3 − ]
PZ- 3.76 −2.016 MDEA- 17.1643 −6.9124
[MDEAH+ HCO3 − ] [H3 O+ , HCO3 − ]
H2 O- 8.72 −2.73 MDEA- 8.6666 −5.8048
[MDEAH+ PZCOO− ] [MDEAH+ HCO3 − ]
PZ- 0.406 −10.85
[MDEAH+ PZCOO− ]

˛ij , nonrandomness parameter.


, energy interaction parameter.
S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237 233

Table 8
Solubility of CO2 in aqueous 30 mass% (MDEA + PZ) in the temperature range of 308–328 K. ˛CO2 is defined as mole CO2 /mole amine (MDEA + PZ).

T = 303 K T = 313 K T = 323 K

˛CO2 pCO2 /kPa ˛CO2 pCO2 /kPa ˛CO2 pCO2 /kPa

30 mass% MDEA + 0 mass% PZ


0.177877 1.985098 0.11659 2.908719 0.060196 1.309613
0.258702 6.417104 0.27418 8.684802 0.193339 8.643446
0.326853 7.444116 0.29159 11.45567 0.262099 14.62631
0.357307 7.016769 0.39500 19.52702 0.348954 24.48976
0.463834 18.09334 0.47790 28.59092 0.430622 36.46928
0.5646 18.73436 0.59547 47.50449 0.508771 52.00542
0.678045 55.41731 0.67155 66.81094 0.563896 62.62707
0.700627 34.06372 0.67171 61.89645 0.608779 74.54455
0.783457 49.64123 0.71592 82.25059 0.73246 250.8254
0.785604 92.60342 0.76201 88.08871 0.83509 595.5982
0.827614 102.908 0.83912 109.3182 1.007072 719.3911
0.861579 76.04716 1.00888 232.9733 1.195895 1039.902
0.899962 95.00898 1.19977 305.6223 1.360726 1238.411
0.979441 623.5136 1.40361 426.1067 1.427663 1306.649

28 mass% MDEA + 2 mass% PZ


0.318193 3.294711 0.16133 1.854136 0.224815 5.059242
0.376374 5.617551 0.35675 10.55272 0.383465 19.5408
0.444572 7.795644 0.52971 28.85973 0.491688 36.66227
0.557528 14.68834 0.61282 44.97487 0.54308 45.3195
0.687868 28.91488 0.67728 63.37148 0.597295 60.55237
0.843066 69.08553 0.72087 80.31374 0.667036 105.5272
0.878696 88.92962 0.74426 91.70048 0.885668 472.7703
0.893747 101.7707 0.989794 685.5479
0.99764 681.4123
1.084272 904.7358
1.28443 1066.025

25 mass% MDEA + 5 mass% PZ


0.295811 1.178652 0.14755 0.420455 0.11404 1.013227
0.448556 4.487148 0.28981 2.805329 0.253571 4.473362
0.606706 13.4063 0.42538 9.684244 0.308128 7.485472
0.707522 26.3439 0.52876 20.23697 0.377294 12.93071
0.79949 44.45792 0.61690 35.44226 0.449729 21.49144
0.855701 63.36459 0.68472 52.76362 0.513509 32.49219
0.896419 82.45737 0.74003 74.99258 0.575622 47.24257
0.912088 90.49426 0.77122 90.2668 0.607838 56.72692
0.928961 99.28934 0.629849 62.66843
0.678308 78.25972
0.711554 87.95774

22 mass% MDEA + 8 mass% PZ


0.04422 0.089605 0.06333 0.17921 0.150062 5.369413
0.149691 0.372206 0.14808 0.454918 0.255398 7.774966
0.237104 1.082154 0.2258 0.999442 0.370384 13.98529
0.353214 2.508943 0.32556 2.515836 0.425529 18.82396
0.422145 4.49404 0.38486 4.232118 0.495948 28.78392
0.422392 5.679585 0.48306 9.146613 0.557388 42.12818
0.507767 12.15183 0.56820 17.22486 0.62089 60.09056
0.590798 21.02274 0.63581 27.67419 0.661406 77.28785
0.662262 34.367 0.68994 40.14998 0.673682 78.30107
0.745573 56.69246 0.80647 87.75786 0.713514 88.31617
0.790882 87.95085 0.843874 242.0923
0.963171 593.9646
1.055765 696.2523
1.187385 1118.775
1.247569 1246.29

with the increase in PZ mass% at higher partial pressure range of and CO2 partial pressure range of 1–1400 kPa are presented
100–1000 kPa. This indicates that the promoter action of PZ is real- in Table 11. The concentration of sulfolane is maintained at
ized more in the lower partial pressure range, which is similar to the 10 mass% while keeping total amine (MDEA + PZ) concentration
CO2 partial pressure condition of the flue gas from coal fired power as 50 mass% in the aqueous solution. Fig. 4 shows the effect of
plants. Various authors used different units for solvent capacity, e.g. CO2 solubility with addition of 10 mass% sulfolane in aqueous
moles of CO2 /(kg H2 O + kg amine), moles of CO2 /mole of amine/(kg (45 mass% MDEA + 5 mass% PZ) mixed solvent at 50 ◦ C. The results
H2 O), etc. The cycle capacity is calculated following Eq. (16). show that at low loading region the CO2 solubility decreases
with replacement of water by sulfolane. However, at high loading
region, CO2 solubility increases with sulfolane. The solubility of
4.3. Equilibrium CO2 partial pressure over aqueous mixed amine
CO2 increases with PZ concentration and this proves that PZ is
(MDEA + PZ + sulfolane)
an effective component in absorbing CO2 into (MDEA + sulfolane)
The new solubility data of CO2 in aqueous as well. Jenab et al. (2005) reported that the solubility of
(MDEA + PZ + sulfolane) over the temperature range of 313–333 K CO2 in (MDEA + sulfolane) solutions decreases with sulfolane
234 S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237

Table 9
Solubility of CO2 in aqueous 50 mass% (MDEA + PZ) in the temperature range of 308–328 K. ˛CO2 is defined as mole CO2 /mole amine (MDEA + PZ).

T = 313 K T = 323 K T = 333 K

˛CO2 pCO2 /kPa ˛CO2 pCO2 /kPa ˛CO2 pCO2 /kPa

50 mass% MDEA + 0 mass% PZ


0.39461 32.0512 0.43732 71.341 0.20439 18.62006
0.59658 103.81 0.54839 141.44 0.28753 44.6132
0.77578 279.71 0.70435 262.27 0.41068 86.18844
0.8346 344.78 0.76979 359.05 0.56135 157.89367
0.91323 547.3 0.83071 463.41 0.68922 293.58202
0.95708 670.26 0.86097 537.5 0.80372 457.88351
0.90777 653.51 0.84509 550.34189
0.91279 681.29
0.91571 725.27

48 mass% MDEA + 2 mass% PZ


0.2582 8.134 0.2624 13.096 0.32048 41.35714
0.38429 23.987 0.42528 56.246 0.40969 90.98571
0.46774 50.042 0.5005 84.644 0.51931 173.7
0.62904 115.25 0.66448 216.92 0.63155 244.00714
0.78533 257.79 0.83214 432.53 0.69173 353.60357
0.90777 489.87 0.89231 523.17 0.77238 473.53929
0.92073 523.17 0.90777 636.49 0.83214 676.18929
1.00389 878.89 0.91822 672.05 0.92616 868.5
1.02478 949.14 0.99386 1029.83 0.97296 1122.8464
1.03021 1052.54 1.03021 1244.21 0.98842 1364.7857
1.05111 1329.63

45 mass% MDEA + 5 mass% PZ


0.34563 10.339 0.3153 12.407 0.38416 49.45712
0.44296 28.192 0.44819 41.839 0.45478 76.51071
0.58468 71.272 0.60698 118.83 0.84258 591.40714
0.70468 125.31 0.64788 139.44 0.87142 678.25714
0.74647 147.73 0.73104 213.68 0.89733 779.58214
0.81417 224.71 0.80124 322.52 0.96753 1091.8285
0.86097 285.91 0.83258 382.69 0.98592 1220.0357
0.90777 407.37 0.85303 417.71 1.00138 1327.5642
0.95458 538.54 0.87643 458.37
0.97296 602.64 0.88186 481.95
0.98592 671.98 0.91028 557.36
1.00138 781.65 0.96252 713.96
1.04567 1058.74 0.99093 899.52
1.06155 1298.61 1.03774 1195.22
1.05111 1358.58

42 mass% MDEA + 8 mass% PZ


0.37639 7.927 0.354 12.338 0.31901 20.40286
0.39482 8.961 0.366 13.924 0.33807 23.98714
0.45349 14.751 0.41901 22.884 0.3608 36.32536
0.57083 42.322 0.53569 57.279 0.49346 89.53821
0.76076 128.89 0.7052 152.95 0.60886 210.6457
0.91759 347.95 0.824 309.97 0.6463 269.5107
0.96461 523.17 0.877 424.05 0.70648 346.0214
0.97852 589.13 0.91279 525.23 0.72553 406.0582
1.02553 1007.04 0.93987 664.19 0.83786 629.4557
1.03494 1048.4 0.95458 688.59 0.85458 743.05
1.05825 1300.68 0.99603 1015.32 0.8678 749.1846
1.0654 1368.92 1.01074 1112.5 0.98742 1133.461
1.05019 1426.82 1.00422 1329.908

concentration. We have also observed the same trend, but this is MDEA and PZ disappear consistently with CO2 loading and the
mainly predominant at lower loading and lower partial pressure. main reaction products are protonated MDEA, protonated PZ, pro-
It is probably due to the fact that at lower partial pressure the tonated PZ carbamate, and bicarbonate ion. Initially PZ carbamate
physical absorption of CO2 in sulfolane is not much realized. reaches a maximum concentration of around 12% of total initial
amine at 0.14 loading and later it is converted to PZ dicarbamate at
4.4. Speciation and pH a loading of about 0.8. The concentration of PZ dicarbamate reaches
about 22% of total amine and it is then converted to protonated PZ
On estimation of the activity coefficients by eNRTL model, it carbamate. By comparing the speciation of (CO2 + PZ + H2 O) pre-
is possible to calculate the liquid phase concentration of vari- sented by Dash et al. (2011a) and (CO2 + MDEA + H2 O) presented
ous ionic and molecular species of (MDEA + PZ) solution. Fig. 5 in this work, it can be noted that there is a significant difference
shows the equilibrium concentration profiles of various ionic and among the speciation of aqueous PZ, aqueous MDEA and aqueous
molecular species of CO2 loaded aqueous (30 mass% MDEA) at (MDEA + PZ) in the presence of CO2 . In aqueous PZ, the carbamate
318 K. Fig. 6 shows the equilibrium concentration profiles of vari- species are less than 0.2 mole%. On the other hand, this can account
ous ionic and molecular species of CO2 loaded aqueous (25 mass% for as much as 0.5 mole% in (25 mass% MDEA + 5 mass% PZ). The con-
MDEA + 5 mass% PZ) at 318 K. It is evident from Fig. 6 that both centration of protonated PZ carbamate increases up to 0.8 mole%
S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237 235

Table 10
Solubility of CO2 in aqueous (MDEA + PZ + sulfolane) in the temperature range of 313–333 K, ˛CO2 is defined as mole CO2 /mole amine (MDEA + PZ).

T = 313 K T = 323 K T = 333 K

˛CO2 pCO2 /kPa ˛CO2 pCO2 /kPa ˛CO2 pCO2 /kPa

42 mass% MDEA + 8 mass% PZ + 10 mass% sulfolane


0.32075 5.37654 0.26375 16.79996 0.26375 16.79996
0.56687 49.49174 0.45486 65.7227 0.45486 65.7227
0.6711 101.3271 0.61483 174.9331 0.61483 174.93313
0.73772 145.0977 0.80488 326.321 0.79365 342.73702
0.84823 274.2725 0.89837 497.3534 0.89837 497.35342
0.8601 332.9319 0.93632 670.4675 0.95382 642.18746
0.97341 545.9256 0.9971 833.2492 1.02811 834.08662
1.07497 779.5983 1.06907 1095.729 1.08265 976.36986
1.12966 984.3204 1.09082 1164.451 1.15301 1141.7251
1.21163 1236.604

45 mass% MDEA + 5 mass% PZ + 10 mass% sulfolane


0.38134 39.28929 0.11222 6.20357 0.2816 39.89724
0.63747 97.18929 0.48818 86.85 0.49178 125.49825
0.87643 328.78929 0.73503 225.44595 0.66844 308.65525
0.9837 554.91509 0.91479 492.84171 0.7945 458.63004
1.0682 812.58225 0.98892 622.91002 0.86031 630.55857
1.13158 1046.3357 1.04362 851.53108 0.95382 834.35968
1.07888 1022.5100 1.04754 1070.1643
1.0944 1104.7673 1.15693 1334.1607
1.13734 1292.3657

48 mass% MDEA + 2 mass% PZ + 10 mass% sulfolane


0.10788 2.06786 0.2633 26.88214 0.24947 41.28132
0.36246 28.95 0.4736 82.71429 0.46432 192.62779
0.4929 53.76429 0.74143 246.075 0.63613 343.83639
0.63348 93.05357 0.89129 496.28571 0.75374 616.88314
0.78973 254.34643 0.96949 620.35714 0.89428 855.12786
0.93423 496.28571 1.02027 856.09286 0.99593 1082.8678
0.99676 622.425 1.10904 951.21429 1.06187 1127.3267
1.03187 756.83571 1.1198 1355.4114
1.07105 940.875

Table 11
Solvent capacitya of (MDEA + PZ) solvent.

50 mass% (48 + 2) mass% (45 + 5) mass% (42 + 8) mass%


MDEA MDEA + PZ MDEA + PZ MDEA + PZ

T = 313 K
Solvent capacity (10–100) kPa CO2 partial pressure 1.3637 1.43562 1.491875 1.518473
Solvent capacity(100–1000) kPa CO2 partial pressure 1.7833 1.64436 1.54 1.456131

T = 323 K
Solvent capacity (10–100) kPa CO2 partial pressure 1.128724 1.22262 1.30375 1.367071
Solvent capacity(100–1000) kPa CO2 partial pressure 2.173528 1.96812 1.82875 1.701046
a
Various authors used different units for solvent capacity, e.g. moles of CO2 /(kg H2 O + kg amine), moles of CO2 /mole of amine/(kg H2 O), etc. But the above calculation is
based on Eq. (16).

in the same solution at a loading of about 1. It is observed from


Fig. 7 that protonation represents a significant amount of product
with the addition of CO2 into aqueous (MDEA + PZ). Similar obser-
vation is also reported by Bishnoi and Rochelle (2002). The overall
effect is shown to be an increase in reactive PZ concentration and
hence the rate of reaction. The model estimated speciation of CO2
loaded (MDEA + PZ + sulfolane) is represented by Fig. 8. The mole
fraction of species has been calculated excluding H2 O for clarity. It is
observed from Fig. 8 that the species protonated piperazine carba-
mate reduces after a loading of 0.2. The concentration of piperazine
carbamate is maximum at 0.1 CO2 loading, then it decreases and
finally disappears at a loading of 0.5. PZ-dicarbamate forms after a
loading of 0.1 and reaches a maximum value at a loading of 0.4, then
reduces and disappears at a loading of 1.0. The concentration of pro-
tonated PZ is maximum at 0.05 loading. It decreases and disappears
at a loading of 0.3. It shows that PZ-carbamate gets converted to PZ-
Fig. 4. Comparison of experimental and model predicted equilibrium CO2 solubility dicarbamate, and hence, CO2 uptake increases. Both protonated-PZ
data in aqueous (45 mass% MDEA + 5 mass% PZ + 10 mass% sulfolane) at 50 ◦ C.
and PZ-carbamate are converted to PZ-di-carbamate. After loading
of 0.4 PZ-di-carbamate and protonated PZ-carbamate are converted
to bicarbonate. Hence, the final product is bicarbonate and the
236 S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237

Fig. 5. Comparison of experimental VLE data of CO2 − (MDEA + PZ + H2 O) with that


of CO2 − (MDEA + PZ + sulfolane + H2 O).
Fig. 8. eNRTL model predicted equilibrium liquid phase speciation of a CO2 loaded
aqueous (MDEA + PZ + sulfolane).

Fig. 6. Liquid phase speciation of CO2 loaded 30 mass% MDEA solution at 313 K.
Fig. 9. Model predicted amine volatility and pH of CO2 loaded aqueous
(MDEA + PZ + sulfolane).

partial pressure of amine with CO2 concentration. It is found from


Fig. 9 that PZ is more volatile up to a loading of 0.2 and volatility
is nearly constant after loading of 0.3. MDEA volatility is negligible
while sulfolane volatility is realized at a loading greater than 0.6.
It is observed that lean solutions have higher amine volatility then
loaded solutions where the free amine has been consumed.

5. Conclusion

In this work, new experimental results of VLE of CO2 in


aqueous (MDEA + PZ) [(22 mass% MDEA + 8 mass% PZ), (25 mass%
MDEA + 5 mass% PZ) and (28 mass% MDEA + 2 mass% PZ) and
(42 mass% MDEA + 8 mass% PZ), (45 mass% MDEA + 5 mass% PZ)
and (48 mass% MDEA + 2 mass% PZ)] have been presented. Effect
of hybrid solvent (MDEA + PZ + sulfolane) on CO2 solubility has
Fig. 7. eNRTL model predicted equilibrium liquid phase speciation of a CO2 loaded
PZ activated aqueous MDEA.
also been reported. The hybrid solvent is prepared by replacing
10 mass% water with 10 mass% sulfolane in aqueous (MDEA + PZ).
The eNRTL model of this work has been used to describe the equi-
reversible reaction of PZ to PZ carbamate is maintained within cer- librium thermodynamics of CO2 in these solvents. The model will
tain loading. The pH of the carbonated (MDEA + PZ + sulfolane) at be important for simulation of absorption-regeneration process for
various CO2 loadings is given in Fig. 9. CO2 removal. The liquid phase speciations of CO2 loaded aqueous
MDEA, (MDEA + PZ) and (MDEA + PZ + sulfolane) are different. The
4.5. Amine volatility major product is bicarbonate, which indicates the carbamate rever-
sion of PZ-carbamate is favoured in this solvent. The results show
The model predicted volatility of (MDEA + PZ + sulfolane) at dif- that replacement of water with sulfolane has a positive effect on
ferent CO2 loading is given in Fig. 9. Amine volatility is expressed in the solubility of CO2 in MDEA based solvents. The magnitude of the
S.K. Dash, S.S. Bandyopadhyay / International Journal of Greenhouse Gas Control 44 (2016) 227–237 237

effect of addition of sulfolane on loading and capacity is different Kamps, A.P.-S., Xia, J.-Z., Maurer, G., 2003. Solubility of CO2 in (H2 O + piperazine)
for different compositions and temperatures. This indicates that the and in (H2 O + MDEA + piperazine). AIChE J. 49, 2662–2670.
Kohl, A.L., Nielsen, R.B., 1997. Gas Purification, fifth ed. Gulf Publishing Company,
compositions of the hybrid solvent should be suitably tuned to get Houston.
better CO2 solubility based on its application for CO2 capture from Kundu, M., Bandyopadhyay, S.S., 2005. Modelling vapour–liquid equilibrium of
low pressure flue gas streams of power plant or for CO2 removal CO2 in aqueous N-methyldiethanolamine through the simulated annealing
algorithm. Can. J. Chem. Eng. 83, 344–353.
from high pressure natural gas streams. Liu, H.-B., Zhang, C.-F., Xu, G.-W., 1999. A study on equilibrium solubility for carbon
dioxide in methyldiethanolamine–piperazine–water solution. Ind. Eng. Chem.
References Res. 38 (10), 4032–4036.
Mathias, P.M., Reddy, S., Smith, A., Afshar, K., 2013. Thermodynamic analysis of CO2
capture solvents. Int. J. Greenh. Gas Control 19, 262–270.
Angaji, M.T., Ghanbarabadi, H., Gohari, F.K.Z., 2013. Optimizations of sulfolane
Mundhwa, M., Elmahmudi, S., Maham, Y., Henni, A., 2009. Molar Heat capacity of
concentration in proposed sulfinol-M solvent instead of MDEA solvent in the
aqueous sulfolane, 4-formylmorpholine, 1-methyl-2-pyrrolidinone, and
refineries of Sarakhs. J. Nat. Gas Sci. Eng. 15, 22–26.
triethylene glycol dimethyl ether solutions from (303.15 to 353.15) K. J. Chem.
Aspen Technology Inc., 2011. Aspen Physical Property System. V-7.3, Cambridge.
Eng. Data 54, 2895–2901.
Bishnoi, S., Rochelle, G.T., 2002. Thermodynamics of
Murrieta-Guevara, F., Romero-Martinez, A., Trejo, A., 1988. Solubilities of carbon
piperazine/methyldiethanolamine/water/carbon dioxide. Ind. Eng. Chem. Res.
dioxide and hydrogen sulfide in propylene carbonate, N-methylpyrrolidone
41, 604–612.
and sulfolane. Fluid Phase Equilib. 44 (1), 105–115.
Bottger, A., Ermatchkov, V., Maurer, G., 2009. Solubility of carbon dioxide in
Najibi, H., Maleki, N., 2013. Equilibrium solubility of carbon dioxide in
aqueous solutions of N-methyldiethanolamine and piperazine in the high gas
N-methyldiethanolamine + piperazine aqueous solution: experimental
loading region. J. Chem. Eng. Data 54, 1905–1909.
measurement and prediction. Fluid Phase Equilib. 354, 298–303.
Closmann, F., Nguyen, T., Rochelle, G.T., 2009. MDEA/piperazine as a solvent for
Poling, B.E., Prausnitz, J.M., O’Connell, J.P., 2001. The Properties of Gases and
CO2 capture. Energy Procedia 1, 1351–1357.
Liquids, fifth ed. McGraw-Hill, New York.
Dash, S.K., Samanta, A., Samanta, A.N., Bandyopadhyay, S.S., 2011a. Vapour liquid
Puxty, G., Maeder, M., 2013. A simple chemical model to represent CO2 -amine-H2 O
equilibria of carbon dioxide in dilute and concentrated aqueous solutions of
vapour–liquid-equilibria. Int. J. Greenh. Gas Control 17, 215–224.
piperazine at low to high pressure. Fluid Phase Equilib. 300, 145–154.
Shokouhi, M., Jalili, A.H., Mohammadian, A.H., Hosseini-Jenab, M., Nouri, S.S., 2013.
Dash, S.K., Samanta, A.N., Bandyopadhyay, S.S., 2011b. (Vapour + liquid) equilibria
Heat capacity, thermal conductivity and thermal diffusivity of aqueous
(VLE) of CO2 in aqueous solutions of 2-amino-2-methyl-1-propanol: new data
sulfolane solutions. Thermochim. Acta 560, 63–70.
and modelling using eNRTL-equation. J. Chem. Thermodyn. 43, 1278–1285.
Steele, W.V., Chirico, R.D., Knipmeyer, S.E., Nguyen, A., 1997. Vapor pressure, heat
Dash, S.K., Samanta, A.N., Bandyopadhyay, S.S., 2011c. Solubility of carbon dioxide
capacity, and density along the saturation line, measurements for
in aqueous solution of 2-amino-2-methyl-1-propanol and piperazine. Fluid
cyclohexanol, 2-cyclohexen-1-one, 1,2-dichloropropane,
Phase Equilib. 307, 166–174.
1,4-di-tert-butylbenzene, (+/−)-2-ethylhexanoic acid, 2-(methylamino)
Dash, S.K., Samanta, A.N., Bandyopadhyay, S.S., 2012. Experimental and theoretical
ethanol, perfluoro-n-heptane, and sulfolane. J. Chem. Eng. Data 42, 1021–1036.
investigation of solubility of carbon dioxide in concentrated aqueous solution
Speyer, D., Ermatchkov, V., Maurer, G., 2010. Solubility of carbon dioxide in
of 2-amino-2-methyl-1-propanol and piperazine. J. Chem. Thermodyn. 51,
aqueous solutions of N-methyldiethanolamine and piperazine in the low gas
120–125.
loading region. J. Chem. Eng. Data 55, 283–290.
Dash, S.K., Samanta, A.N., Bandyopadhyay, S.S., 2014. Simulation and parametric
Svensson, H., Hulteberg, C., Karlsson, H.T., 2013. Heat of absorption of CO2 in
study of the post combustion CO2 capture using aqueous
aqueous solutions of N-methyldiethanolamine and piperazine. Int. J. Greenh.
2-amino-2-methyl-1-propanol and piperazine. Int. J. Greenh. Gas Control 21,
Gas Control 17, 89–98.
130–139.
Tong, D., Trusler, M., Maitland, G.C., Gibbins, J., Fennell, P.S., 2012. Solubility of
Derks, P.W.J., Hogendoorn, J.A., Versteeg, G.F., 2010. Experimental and theoretical
carbon dioxide in aqueous solution of monoethanolamine or
study of the solubility of carbon dioxide in aqueous blends of piperazine and
2-amino-2-methyl-1-propanol: experimental measurements and modelling.
N-methyldiethanolamine. J. Chem. Thermodyn. 42, 151–163.
Int. J. Greenh. Gas Control 6, 37–47.
Derks, P.W.J., Hamborg, E.S., Hogendoorn, J.A., Niederer, J.P.M., Versteeg, G.F., 2008.
Vahidi, M., Matin, M.S., Goharrokhi, M., Jenab, M.H., Abdi, M.A., Najibi, S.H., 2009.
Densities, viscosities, and liquid diffusivities in aqueous piperazine and
Correlation of CO2 solubility in N-methyldiethanolamine + piperazine aqueous
aqueous (piperazine + N-methyldiethanolamine) solutions. J. Chem. Eng. Data
solutions using extended Debye–Huckel model. J. Chem. Thermodyn. 41,
53, 1179–1185.
1272–1278.
Derks, P.W.J., Hogendoorn, K.J., Versteeg, G.F., 2005. Solubility of N2 O in and
Vahidi, M., Moshtari, B., 2013. Dielectric data, densities, refractive indices, and
density, viscosity, and surface tension of aqueous piperazine solutions. J.
their deviations of the binary mixtures of N-methyldiethanolamine with
Chem. Eng. Data 50, 1947–1950.
sulfolane at temperatures 293.15–328.15 K and atmospheric pressure.
Ermatchkov, V., Maurer, G., 2011. Solubility of carbon dioxide in aqueous solutions
Thermochim. Acta 551, 1–6.
of N-methyldiethanolamine and piperazine: prediction and correlation. Fluid
Wang, C.-W., Soriano, A.N., Yang, Z.-Y., Li, M.-H., 2010. Solubility of carbon dioxide
Phase Equilib. 302, 338–346.
in the solvent system (2-amino-2-methyl-1-propanol + sulfolane + water).
Freeman, S.A., Dugas, R., Wagener, D.H.V., Nguyen, T., Rochelle, G.T., 2010. Carbon
Fluid Phase Equilib. 291, 195–200.
dioxide capture with concentrated, aqueous piperazine. Int. J. Greenh. Gas
Xu, G.-W., Zhang, C.-F., Qin, S.-J., Gao, W.-H., Liu, H.-B., 1998. Gas–liquid
Control 4, 119–124.
equilibrium in a CO2 –MDEA–H2 O system and the effect of piperazine on it. Ind.
Fu, D., Qin, L., Hao, H., 2013. Experiment and model for the viscosity of carbonated
Eng. Chem. Res. 37, 1473–1477.
piperazine-N-methyldiethanolamine aqueous solutions. J. Mol. Liq. 186, 81–84.
Zhang, Y., Chen, C.-C., 2011. Thermodynamic modelling for CO2 absorption in
Fulem, M., Ruzicka, K., Ruzicka, M., 2011. Recommended vapor pressures for
aqueous MDEA solution with electrolyte NRTL model. Ind. Eng. Chem. Res. 50,
thiophene, sulfolane, and dimethyl sulfoxide. Fluid Phase Equilib. 303,
163–175.
205–216.
Zoghi, A.T., Feyzi, F., Zarrinpashneh, A., 2012. Experimental investigation on the
Jenab, M.H., Abdi, M.A., Najibi, S.H., Vahidi, M., Matin, N.S., 2005. Solubility of
effect of addition of amine activators to aqueous solutions of
carbon dioxide in aqueous mixtures of
N-methyldiethanolamine on the rate of carbon dioxide absorption. Int. J.
N-methyldiethanolamine + piperazine + sulfone. J. Chem. Eng. Data 50,
Greenh. Gas Control 7, 12–19.
583–586.
Zong, L., Chen, C.-C., 2011. Thermodynamic modelling of CO2 and H2 S solubilities
Jou, F.-Y., Otto, F.D., Mather, A.E., 1982. Solubility of H2 S and CO2 in aqueous
in aqueous DIPA solution, and aqueous sulfolane-MDEA solution with
methyldiethanolamine solutions. Ind. Eng. Chem. Process Des. Dev. 21,
electrolyte NRTL model. Fluid Phase Equilib. 306, 190–203.
539–544.
Jou, F.-Y., Deshmukh, R.D., Otto, F.D., Mather, A.E., 1990. Solubility of H2 S, CO2 , CH4
and C2 H6 in sulfolane at elevated pressures. Fluid Phase Equilib. 56, 313–324.

You might also like