You are on page 1of 6

J. Chem.

Thermodynamics 42 (2010) 802807

Contents lists available at ScienceDirect

J. Chem. Thermodynamics
journal homepage: www.elsevier.com/locate/jct

Equilibrium solubility of carbon dioxide in the amine solvent system


of (triethanolamine + piperazine + water)
Pei-Yuan Chung a, Allan N. Soriano a,b, Rhoda B. Leron a, Meng-Hui Li a,*
a
b

R&D Center for Membrane Technology and Department of Chemical Engineering, Chung Yuan Christian University, Chung Li 32023, Taiwan, ROC
School of Chemical Engineering and Chemistry, Mapa Institute of Technology, Manila 1002, Philippines

a r t i c l e

i n f o

Article history:
Received 31 December 2009
Received in revised form 3 February 2010
Accepted 4 February 2010
Available online 10 February 2010
Keywords:
Carbon dioxide
Solubility
Blended amines
Triethanolamine
Piperazine

a b s t r a c t
In this study, a new set of data for the equilibrium solubility of carbon dioxide in the amine solvent system that consists of triethanolamine (TEA), piperazine (PZ), and water is presented. Equilibrium solubility
values were obtained at T = (313.2, 333.2, and 353.2) K and pressures up to 153 kPa using the vapourrecirculation equilibrium cell. The TEA concentrations in the considered ternary (solvent) mixture were
(2 and 3) kmol  m3 and those of PZs were (0.5, 1.0, and 1.5) kmol  m3. The solubility data (CO2 loading
in the amine solution) obtained were correlated as a function of CO2 partial pressure, system temperature, and amine composition via the modied KentEisenberg model. Results showed that the model
applied is generally satisfactory in representing the CO2 absorption into mixed aqueous solutions of
TEA and PZ.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Carbon dioxide has long been identied as one of the major contributors to global warming. Thus, its removal and recovery from
industrial process streams and emissions have become vital and
the search for efcient and economical absorbents has gained signicant interests [13].
In industrial gas treating processes, aqueous alkanolamine solutions have been widely used to remove acid gases such as CO2 and
H2S, triethanolamine (TEA), a tertiary amine, being one of the rst
amines used for such application. Although it was superseded by
methyldiethanolamine (MDEA) and monoethanolamine (MEA)
[4], its application in the removal of acid gas is still recommended.
This has been proven throughout the decades as research interests
on TEA continuously increased resulting in the availability of
numerous experimental solubility data (H2S and CO2 in aqueous
ethanolamine solutions) in a wide range of temperatures, pressures, solvent compositions, and acid gas loadings [48]. The use,
however, of such solutions is subject to limitations in loading
capacity and absorption rate. Thus, the search for an effective promoter, such as piperazine (PZ), has gained recent interests in research [9].
Studies have been conducted on the effect of piperazine (PZ)
on the rate of absorption of CO2 when blended in aqueous

* Corresponding author. Tel.: +886 3 265 4109; fax: +886 3 265 4199.
E-mail address: mhli@cycu.edu.tw (M.-H. Li).
0021-9614/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jct.2010.02.005

alkanolamines solutions. It has been found out that PZ has higher


reaction rate with CO2 than monoethanolamine (MEA), 2-amino2-methyl-1-propanol (AMP), and dimethylethanolamine (DMEA)
and has been reported to be superior than MEA, AMP, and DMEA
for CO2 absorption [9]. The small addition of PZ in aqueous AMP
solutions can also promote the reaction rates with CO2 [1,9] and
has signicant effect on the enhancement of the CO2 absorption
rate [2]. The mixture of MEA and PZ also exhibited efcient CO2
removal with the CO2 removal efciency increasing at higher
amounts of PZ [9]. In general, PZ-alkanolamine blends yield
improvements in absorption, such as higher loading capacity, faster rate of reaction, and lower solvent regeneration energy
requirements [3,10]. Although a number of studies have already
investigated the effect of PZ, when blended with known alkanolamines, on the absorption of CO2, no data is yet available on CO2
solubility in aqueous mixtures of TEA and PZ. Thus, it is the main
aim of this work to present new set of solubility data for CO2 in
aqueous solvent system containing blended amine solutions of
TEA and PZ. The solubilities of CO2 in the TEAPZ [(2 and
3) kmol  m3 and (0.5, 1.0, and 1.5) kmol  m3, respectively] solvent systems were measured using a vapour-recirculation equilibrium cell at temperatures (313.2, 333.2, and 353.2) K and CO2
partial pressures of up to 153 kPa. Validation of the applied
experimental methods was done by measuring the CO2 solubility
in aqueous MDEA solution (2.0 kmol  m3) at 313.2 K and CO2
partial pressures of up to 90 kPa. The obtained equilibrium solubility data (CO2 loading in the total amine solution) in the aqueous blended amine systems were correlated as a function of CO2

803

P.-Y. Chung et al. / J. Chem. Thermodynamics 42 (2010) 802807

partial pressure, system temperature, and amine composition


using a modied KentEisenberg model.
2. Experimental section
2.1. Chemicals
The alkanolamine samples, TEA, and (PZ and MDEA), which
were all reagent grade with a minimum mass fraction purity
0.99, were acquired from TEDIA and ACROS Organics Co., respectively. Reagent grade carbon dioxide with a minimum mass fraction purity 0.999 was obtained from Liehwa Industrial Gases,
(KUN Technology, Co., Ltd.). The 1-M (kmol  m3) standard solutions of NaOH(aq) and HCl(aq) were supplied by Fluka and the BaCl2
(mass fraction purity >0.99) was purchased from SHOWA. The liquid water used to prepare the solutions was a Type I reagentgrade water (resistivity = 18.3 MX  cm and total organic carbon
content (TOC) <15  109) and produced via Barnstead Thermodyne
(model Easy Pure 1052) water purication system. A MettlerToledo (model AL204) digital balance with accuracy 1  104 g was
used in the preparation of the aqueous alkanolamine solutions.
2.2. Measurement of CO2 solubility
The solubility apparatus (vapour-recirculation equilibrium cell)
and methods of analyses employed in this work are similar as
those used in our previous works [1114]. Only the most pertinent
details are discussed in the following paragraph.
A 1.0-dm3 stainless steel vapour-recirculation equilibrium cell
was used in the measurement of the equilibrium solubility of
CO2 in the investigated aqueous blended amine system, i.e., TEA
PZH2O. Another 0.5-dm3 stainless steel sample cylinder was connected to the equilibrium cell to increase the volume of the vapour
phase. Two cylinders were mounted vertically in a thermostat-installed air bath. The temperature of the system was controlled by a
basic immersion circulator obtained from NESLAB Instruments Inc.
(model EX-810B) whose stability was within 0.1 C of the set-point
temperature. To ensure low partial pressures of CO2, nitrogen was
injected into and mixed with CO2 stream and the partial pressure
of CO2 was determined via on-line gas chromatography. The solubility of CO2 (also denoted here as CO2 loading capacity, a) in
amine solutions was determined by the standard acidbase titration method. When both the system pressure and the gas concentration did not change for 2 h, equilibrium was assumed to have
been reached. Usually it would take (6 to 8) h for the system to attain equilibrium. At equilibrium, a liquid sample was withdrawn
from the equilibrium cell and placed directly into a vessel containing an excess of 1.0 M NaOH solution that would convert free dissolved CO2 into non-volatile ionic species. An excess amount of
BaCl2 solution was then added to the solution. The solution was

shaken well to permit all (physically and chemically) absorbed


CO2 to precipitate the carbonate as BaCO3. The excess NaOH was
titrated with HCl solution using phenolphthalein as indicator.
Then, the solution was titrated with HCl with methyl orange as
indicator. From the amount of HCl added, the solubility of CO2 in
terms of the moles of CO2 per mole of amine was then calculated.
Considering all possible sources of uncertainty the average
uncertainty for the following experimental results are estimated
as
follows:
concentration
in
aqueous
solutions
is
0.02 kmol  m3, temperature is 0.01 K, partial pressure is
5 kPa (readings below 100 kPa) and 10 kPa (readings above
100 kPa), and CO2 loading (mol of CO2 per mol of total amine)
is estimated to be around 2% to 3% (standard deviation). The standard deviation was obtained from the carried out replicate measurement runs (three to ve times). All the experimental results
for all the systems studied are the average of such replicate
experiments. The thermometer used to register the temperature
in the system was a Hart Scientic (Fluke) digital thermometer,
model 1502A, with a precision of 0.006 K, adapted with a thermocouple (Pt), which was compared with the readings from a
CROPICO digital thermometer, model 3002, having an accuracy
of 0.005 K. The equilibrium pressure during the process was
indicated by an Ashcroft test gauge with the scale reading range
of 414 kPa and has an accuracy of 0.01 kPa (readings below
100 kPa) and 0.1 kPa (readings above 100 kPa). The pressure
indicator was calibrated by a digital pressure gauge from GE
Druck, model PTX-1400, with an accuracy of 0.15%. The thermometer and pressure gauge used were periodically calibrated
by the National Measurement Laboratory (ROC), a standard centre
for calibration.
3. Results and discussion
Prior to the measurement of CO2 solubility in the concerned solvent systems,
the CO2 solubility in aqueous (2.0 kmol  m3) MDEA solution at T = 313.2 K and
CO2 partial pressure of up to 90 kPa were measured. This was done to ensure that
the applied procedures and method for the solubility measurements would yield
valid and acceptable results. The aqueous (2.0 kmol  m3) MDEA solution at
T = 313.2 K condition were chosen for the validation since there were a number
of consistent data available in open literature for such condition. Table 1 contains
the results of our solubility measurements for aqueous (2.0 kmol  m3) MDEA solution at T = 313.2 K from this work together with the available data from the literature [1517]. The graphical comparison among the different measurements for this
system is shown in gure 1. As presented in table 1 and also seen in gure 1, the
solubility results obtained from this work appear to be consistent with the available
data from the literature [1517]. Such consistency was quantied by the smoothed
curve in gure 1, which was regressed from the combined data from this work and
the available literature values [1517] and is represented as in equation (1). The
overall average absolute deviation (AAD), of the measured values (this work) from
the corresponding data predicted by equation (1) was 9.2%; thus, the methods used
in the measurements could be expected to give results with acceptable accuracy

P 0:6632  24:31a 355:4a2  1322a3 2215a4  1093a5 :

TABLE 1
Solubility of CO2 in aqueous (2.0 kmol  m3) MDEA solution at T = 313.2 K.
Jou et al. [15]

MacGregor and Mather [16]

Austgen and Rochelle [17]

This work

P CO2 =kPa

aa

P CO2 =kPa

P CO2 =kPa

P CO2 =kPa

0.184
2.38
11.2
101

0.0680
0.224
0.441
0.866

1.17
4.15
48.9

0.124
0.267
0.686

0.177
0.419
0.887
6.95
92.8

0.0440
0.74sss0
0.113
0.362
0.842

0.28
0.60
3.48
13.26
18.67
46.44
89.90

0.0580
0.0960
0.240
0.442
0.522
0.697
0.803

AAD %b = 21.7

AAD % = 2.7

a is the CO2 loading in mol of CO2 per mol of amine.

Calculated using equation (1).

AAD % = 4.1

AAD % = 9.2

804

P.-Y. Chung et al. / J. Chem. Thermodynamics 42 (2010) 802807


TABLE 2
Solubility of CO2 in aqueous (2.0 kmol  m3) TEA + PZ solutions.
T = 313.2 K
P CO2 =kPa

FIGURE 1. Plot of pressure against the solubility of CO2 in aqueous MDEA


(2.0 kmol  m3) solution at T = 313.2 K: j, This work; h, Jou et al. [15]; s,
MacGregor and Mather [16]; 4, Austgen and Rochelle [17]; and line, smoothed
values, calculated using equation (1).

In equation (1), P is the CO2 partial pressure in kPa and a the CO2 loading capacity in
(mol CO2  mol amine1).
After validating the methodology, the solubility data at the considered CO2
partial pressures, temperatures, and amine solvent systems compositions were
obtained by measurement. Tables 2 and 3 show the results obtained. Figures 2
to 4 show a representative trend of the general behaviour of the observed results.
From these tables and gures, it had been deduce that (1) at constant temperature and solvent (TEAPZH2O) composition, the CO2 loading (solubility) increases with increasing CO2 partial pressure, (2) at constant CO2 partial
pressure and solvent composition, the CO2 loading decreases with increasing temperature, and (3) at constant temperature and TEA concentration in the ternary
mixture, the increase in PZ concentration signicantly increases the CO2 loading
capacity of the aqueous blended amine system. Figures 2 to 4 also show that
the applied model, described in the succeeding paragraphs, satisfactorily predicts
the obtained solubility data.
For the purpose of comparison and application, a correlation to describe the
behaviour of the data obtained from the experimental measurements is essentially
required. For acid gas solubility in various aqueous alkanolamine systems, the modied form of the original KentEisenberg model [18] seems to be the simplest and
yet has a theoretical basis that could be used [15,1922]. In this work, the model
developed by Li and Shen [12], which was also a modication of the original
KentEisenberg model [18] was applied to represent the CO2 solubility in the aqueous blended amine system of (TEA + PZ + H2O).
The chemical reactions that describe the equilibrium of the quarternary system (CO2 + TEA + PZ + H2O) with TEA denoted as R3N, are given as equations (1a)
to (5a)

T = 333.2 K

P CO2 =kPa

(2.0 kmol  m3) TEA


2.60
0.077
8.58
0.114
32.91
0.207
53.88
0.243
60.17
0.257
90.97
0.340
153.4
0.485

5.42
8.05
20.56
43.01
98.14
137.9

0.066
0.080
0.108
0.134
0.178
0.191

TEA + (0.5 kmol  m3) PZ


1.09
0.215
6.30
0.275
16.19
0.346
19.14
0.368
37.91
0.405
43.80
0.433
103.1
0.553

1.01
3.97
11.15
22.45
43.04
87.77

0.161
0.203
0.245
0.283
0.342
0.399

2.31
4.43
8.00
21.59
35.44
46.12
91.73

0.296
0.342
0.359
0.398
0.433
0.446
0.490

1.22
2.91
6.84
16.47
47.08
87.94

0.335
0.356
0.402
0.446
0.493
0.538

P CO2 =kPa

1.62
6.56
9.56
14.98
36.27
92.98

0.085
0.128
0.169
0.205
0.345
0.534

0.99
2.37
7.83
12.73
23.08
41.49
87.77

0.265
0.305
0.384
0.416
0.486
0.567
0.699

3.51
7.84
18.37
30.26
55.87
103.0

0.474
0.545
0.603
0.645
0.740
0.827

1.00
3.17
9.44
15.28
25.83
52.68
98.13

0.432
0.519
0.598
0.635
0.701
0.785
0.861

T = 353.2 K

TEA + (1.0 kmol  m3) PZ


2.28
0.374
7.75
0.421
18.15
0.470
30.53
0.533
45.44
0.569
67.97
0.633
117.9
0.682
TEA + (1.5 kmol  m3) PZ
1.75
0.389
6.24
0.460
18.00
0.544
32.67
0.587
54.18
0.651
80.22
0.691
120.1
0.746

a is the CO2 loading in mol of CO2 per mol of amine.

R3 NH 
;
R3 NH 

PZH 
;
K2
PZH 


H HCO3 
;
K3
CO2 


K 4 H OH ;

K1

K5

2
3
4
5

H CO2
3 
:
HCO3 

Henrys law relating the CO2 partial pressure to the concentration of the physically dissolved CO2 in the solvent is given as

PCO2 HCO2 CO2 ;

where P CO2 is the CO2 partial pressure in kPa and HCO2 is the Henrys law constant.
The mass balance equations for the reacting species are as follows:

mTEA R3 NH  R3 N;

mPZ PZH  PZ;


K1

R3 NH $ H R3 N;

1a

mTEA mPZ a CO2  HCO3  CO2


3 ;

10

R3 NH  PZH  H  OH  HCO3  2CO2


3 ;
K2

PZH $ H PZ;
K3

H2 O CO2 $ H HCO3 ;
K4

H2 O $ H OH ;
K5

HCO3 $ H CO2
3 :

2a
3a
4a

11
3

where mTEA and mPZ are the molar concentrations (kmol  m ) of TEA and PZ,
respectively, and a is the CO2 loading capacity of the solvent system (mol CO2  mol
amine1).
The apparent equilibrium constants, K3 to K5, mentioned in equations (4) to (6)
and the Henrys law constant for CO2, HCO2 , were modied forms (for units consistency) adapted from Kent and Eisenberg [18]. These relations are given as follows:

K 3 =kmol  m3 exp241:818 29:8253  104 =T=K  1:48528  108 =T=K2


5a

Equations (2) to (6) show the expressions for the apparent equilibrium constants, Ki,
as a function of molar concentrations of the species (the one in bracket) present in
their specic reactions

0:332647  1011 =T=K3  0:282393  1013 =T=K4 ;

12

K 4 =kmol  m3 2 exp39:5554  9:879  104 =T=K 0:568827  108 =T=K2 


0:146451  1011 =T=K3 0:136145  1013 =T=K4 ;

13

805

P.-Y. Chung et al. / J. Chem. Thermodynamics 42 (2010) 802807


TABLE 3
Solubility of CO2 in aqueous (3.0 kmol  m3) TEA + PZ solutions.
T = 313.2 K

T = 333.2 K

P CO2 =kPa

P CO2 =kPa

T = 353.2 K

P CO2 =kPa

1.43
3.05
11.31
23.09
36.01
55.43
89.44

0.034
0.045
0.064
0.080
0.107
0.125
0.147

2.66
4.64
13.36
22.18
37.85
61.66
106.0

0.074
0.086
0.157
0.182
0.259
0.313
0.425

(3.0 kmol  m3) TEA


2.37
0.055
6.26
0.073
16.11
0.128
29.34
0.137
32.14
0.145
54.74
0.175
99.41
0.241

1.18
6.42
14.78
28.40
40.52
64.91
109.0

0.200
0.239
0.285
0.353
0.386
0.443
0.519

TEA + (0.5 kmol  m3) PZ


1.24
0.150
6.42
0.196
15.00
0.226
30.01
0.265
48.21
0.303
69.16
0.329
99.55
0.362

1.85
6.95
14.84
24.44
36.68
55.31
84.91

0.134
0.165
0.184
0.209
0.226
0.237
0.265

1.26
6.25
14.73
29.34
34.16
58.84
95.39

0.258
0.315
0.368
0.444
0.472
0.546
0.601

TEA + (1.0 kmol  m3) PZ


2.17
0.250
8.75
0.300
15.96
0.316
28.53
0.353
45.18
0.393
63.41
0.408
103.7
0.434

0.73
2.18
4.76
15.86
31.43
49.00
80.29

0.180
0.212
0.245
0.269
0.309
0.323
0.353

2.63
3.59
6.40
15.50
29.05
46.83
82.22

0.249
0.255
0.279
0.308
0.329
0.345
0.373

3

1.17
2.68
8.32
31.33
49.77
88.87
121.1
a

0.311
0.342
0.392
0.507
0.562
0.613
0.642

TEA + (1.5 kmol  m ) PZ


1.34
0.258
4.77
0.308
12.67
0.338
27.18
0.362
42.77
0.411
68.81
0.437
101.0
0.462

FIGURE 3. Plot of pressure against the solubility of CO2 in {aqueous


(2.0 kmol  m3) TEA + (0.5 kmol  m3) PZ} solution at the considered temperatures: (j, 313.2 K; d, 333.2 K; N, 353.2 K), this work; and lines, calculated using the
modied KentEisenberg model.

a is the CO2 loading in mol of CO2 per mol of amine.

FIGURE 4. Plot of pressure against the solubility of CO2 in {aqueous (2.0 kmol  m3
and 3.0 kmol  m3) TEA + PZ} solution at T = 353.2 K: [j, (2.0 kmol  m3) TEA; d,
TEA + (0.5 kmol  m3) PZ; N, TEA + (1.0 kmol  m3) PZ; ., TEA + (1.5 kmol  m3)
TEA;
s,
TEA + (0.5 kmol  m3)
PZ;
4,
PZ];
[h,
(3.0 kmol  m3)
TEA + (1.0 kmol  m3) PZ; O, TEA + (1.5 kmol  m3) PZ]; and [solid lines,
3
3
(2.0 kmol  m ) TEA solutions; dashed lines, (3.0 kmol  m ) TEA solutions],
calculated using the modied KentEisenberg model.
FIGURE 2. Plot of pressure against the solubility of CO2 in aqueous (2.0 kmol  m3)
TEA solution at the considered temperatures: (j, 313.2 K; d, 333.2 K; N, 353.2 K),
this work; and lines, calculated using the modied KentEisenberg model.

HCO2 =kPa  kmol  m3 1  exp2:97068  0:184393  104 =T=K


0:009217  108 =T=K2  0:002078  1011 =T=K3
0:0016  1013 =T=K4 :

3

K 5 =kmol  m exp294:74 36:4385  10 =T=K  1:84157  10 =T=K


0:415792  1011 =T=K3  0:354291  1013 =T=K4 ;

14

15

In this work, the apparent equilibrium constants, K1 and K2, for the main amine
reactions, equations (1a) and (2a), were assumed to be dependent on temperature
(T), CO2 loading (a), and amine compositions (mTEA and mPZ). Applying the least-

806

P.-Y. Chung et al. / J. Chem. Thermodynamics 42 (2010) 802807

squares t to the equilibrium solubility values obtained for CO2 in aqueous blended
amine system of (TEA + PZ), K1 and K2 were determined to give the following
expressions as in equations (16) and (17), respectively:

K 2 =kmol  m3 expf55:928 10263=T=K  2627900=T=K2 65:922a


42:321a2  4:5132mTEA =kmol  m3
6:9725mPZ =kmol  m3  3:8987mTEA mPZ 3:6595m2PZ g:

K 1 =kmol  m3 expf58:088  45718=T=K 6820300=T=K 3:7833a

17

7:6473a2 0:050349mTEA =kmol  m3 


6:4299mPZ =kmol  m3 1:0139mTEA mPZ 1:939m2PZ g;
16

FIGURE 5. Plot of the predicted (the predicated values are calculated by the
modied KentEisenberg model) partial pressure against the measured partial
pressure for CO2 in {aqueous (2.0 kmol  m3) TEA + PZ} solutions: d,
(2.0 kmol  m3) TEA; s, TEA + (0.5 kmol  m3) PZ; j, TEA + (1.0 kmol  m3) PZ;
h, TEA + (1.5 kmol  m3) PZ.

The constants presented in Eqs. (16) and (17) were regressed from the experimental
solubility results of CO2 in aqueous system of (TEA + PZ) obtained in this work, which
has a total of 163 data points. The constants satisfactorily represent the experimental
solubility results as shown by an acceptable overall AAD of 18.9%. Aside from gures
2 to 4, such a satisfactory result is also shown in gure 5, which contains the plot of
the comparison between the predicted and measured partial pressure of CO2 in the
considered aqueous blended amine system of (TEA + PZ).

FIGURE 7. Plot of pressure against the solubility of CO2 in {aqueous


(2.0 kmol  m3) TEA + (1.5 kmol  m-3) PZ} solution at different temperatures: (j,
313.2 K; d, 333.2 K; N, 353.2 K), this work; and (solid lines, interpolated; dashed
lines, extrapolated), calculated using the modied KentEisenberg model.

PCO2 / kPa

4000

3000

2000

1000

0.0

0.2

0.4

0.6

0.8

-1

/ (mol CO2 mol solvent )

FIGURE 6. Plot of pressure against the solubility of CO2 in aqueous (2.0 kmol  m3)
TEA solution to compare data: T = (j, 313.2 K; d, 333.2 K; N, 353.2 K), this work;
(h, 273.2 K; s, 298.2 K; 4, 323.2 K; O, 348.2 K), Mason and Dodge [5]; and lines,
calculated using the modied KentEisenberg model.

FIGURE 8. Plot of pressure against the solubility of CO2 to compare solubility for
various solvent systems at T = 313.2 K: j, aqueous (2.0 kmol  m3)
TEA + (1.5 kmol  m3) PZ solution, this work; h, [Emim][BF4], reference [23]; (O,
[Emim][MDEGSO4]; ., [Bmim][PF6]), reference [24]; (4, [Emim][triate]; },
[Bmim][BF4]), reference [25]; (s, [Emim][EtSO4]; /, [Bmim][triate]), reference
[26].

P.-Y. Chung et al. / J. Chem. Thermodynamics 42 (2010) 802807


The applicability of the determined constants, K1 and K2, was validated using
the available data of Mason and Dodge [5] for CO2 solubility in aqueous
(2.0 kmol  m3) TEA solution at T = (273.2, 298.2, 323.2, and 348.2) K, as shown
in gure 6. As this gure shows, with the exception of the 273.2 K isotherm, the
determined constants satisfactorily represent the solubility data of Mason and
Dodge [5]; thus, the applied model (together with the obtained parameters) can
also be recommended to be used for a wider range of conditions of temperature
and CO2 partial pressures. gure 6 also shows that the solubility results obtained
are consistent with the reported data of Mason and Dodge [5], which further validates the applied methodology for this work. The model was further validated by
investigating its capability to predict values for conditions where there are no data
available yet. Figure 7 representatively shows such validation. As seen in this gure
(dashed lines), the model applied systematically extrapolates the behaviour of the
solubility data at conditions beyond the experimental setting at T = (293.2 and
373.2) K.
The CO2 absorption performance of the considered system was also compared
to the performance of another class of solvent called ionic liquids (ILs). Here, the
CO2 absorption performance was based on the CO2 solubility in terms of mol of
CO2 per mol of solvent. For the present system, the condition in which the maximum solubility was observed, i.e., in aqueous {(2.0 kmol  m3) TEA +
(1.5 kmol  m3) PZ} solutions at T = 313.2 K, was considered for such comparison.
For the representative IL system, results from our previous works on ILs [2326]
were used at similar temperature condition (T = 313.2 K). Figure 8 contains the plot
of the comparison of the CO2 solubility of the present system against the IL systems
considered. As shown in this gure, the studied system in this work has the highest
CO2 loading per mole of the solvent used. This observation arises from the fact that
the CO2 absorption of ILs was merely physical and that of the considered system is
chemical in nature.

4. Conclusions
A new set of results for the equilibrium solubility of CO2 in the
aqueous blended amine system of (TEA + PZ) at T = (313.2, 333.2,
and 353.2) K, and pressures up to 153 kPa have been determined.
The equilibrium solubility values of CO2 were obtained via measurements in a vapour-recirculation equilibrium cell for TEA and
PZ concentrations of (2 and 3) kmol  m3 and (0.5, 1.0, and
1.5) kmol  m3, respectively. A modied KentEisenberg model
satisfactory correlated the measured equilibrium loading (solubility)/partial pressure pairs at different temperatures and amine
concentration levels, as shown by an acceptable AAD of 18.9% for
a total of 163 data points. The model applied was also proven to
be applicable over a wider range of conditions as it represented
well the available data from the literature and systematically al-

807

lowed extrapolation of the data at conditions beyond the experimental setting.


Acknowledgement
This research was supported by Grants, NSC 98-2221-E-033029 and NSC 98-3114-E-007-013, of the National Science Council
of the Republic of China.
References
[1] D.J. Seo, W.H. Hong, Ind. Eng. Chem. Res. 39 (2000) 20622067.
[2] W.-C. Sun, C.-B. Yong, M.-H. Li, Chem. Eng. Sci. 60 (2005) 503516.
[3] S.-H. Lin, P.-C. Chiang, C.-F. Hsieh, M.-H. Li, K.-L. Tung, J. Chin. Inst. Chem. Eng.
39 (2008) 1321.
[4] Y.G. Li, A.E. Mather, Ind. Eng. Chem. Res. 35 (1996) 48044809.
[5] J.W. Mason, B.F. Dodge, Trans. AIChE 32 (1936) 2748.
[6] M.A. Lyudkovskaya, A.G. Leibush, Zh. Prok. Khim. 22 (1949) 558567.
[7] F.-Y. Jou, F.D. Otto, A.E. Mather, Can. J. Chem. Eng. 63 (1985) 122125.
[8] F.-Y. Jou, F.D. Otto, A.E. Mather, J. Chem. Eng. Data 41 (1996) 1181.
[9] C.-S. Tan, J.-E. Chen, Sep. Purif. Technol. 49 (2006) 174180.
[10] T. Chakravarty, U.K. Phukan, R.H. Weiland, Chem. Eng. Prog. 81 (1985) 3236.
[11] K.-P. Shen, M.-H. Li, J. Chem. Eng. Data 37 (1992) 96100.
[12] M.-H. Li, K.P. Shen, J. Chem. Eng. Data 38 (1993) 105108.
[13] M.-H. Li, B.-C. Chang, J. Chem. Eng. Data 39 (1994) 448452.
[14] M.-H. Li, B.-C. Chang, J. Chem. Eng. Data 40 (1995) 328331.
[15] F.-Y. Jou, A.E. Mather, F.D. Otto, Ind. Eng. Chem. Process Des. Dev. 21 (1982)
539544.
[16] R.J. MacGregor, A.E. Mather, Can. J. Chem. Eng. 69 (1991) 13571366.
[17] D.M. Austgen, G.T. Rochelle, Ind. Eng. Chem. Res. 30 (1991) 543555.
[18] R.L. Kent, B. Eisenberg, Hydrocarbon Process. 55 (1976) 8790.
[19] J.I. Lee, F.D. Otto, A.E. Mather, Can. J. Chem. Eng. 54 (1976) 214219.
[20] W. Hu, A. Chakma, Chem. Eng. Commun. 94 (1990) 5361.
[21] P. Tontiwachwuthikul, A. Meisen, C.J. Lim, J. Chem. Eng. Data 36 (1991) 130
133.
[22] D. Le Tourneux, I. Iliuta, M.C. Iliuta, S. Fradette, F. Larachi, Fluid Phase Equilib.
268 (2008) 121129.
[23] A.N. Soriano, B.T. Doma Jr., M.-H. Li, J. Chem. Eng. Data 53 (2008) 25502555.
[24] A.N. Soriano, B.T. Doma Jr., M.-H. Li, J. Chem. Thermodyn. 40 (2008) 1654
1660.
[25] A.N. Soriano, B.T. Doma Jr., M.-H. Li, J. Chem. Thermodyn. 41 (2009) 525
529.
[26] A.N. Soriano, B.T. Doma Jr., M.-H. Li, J. Taiwan Inst. Chem. Eng. 40 (2009) 387
393.

JCT 10-2

You might also like