You are on page 1of 11

Available online at www.sciencedirect.

com

ScienceDirect
Energy Procedia 63 (2014) 665 675

GHGT-12

Development of a Potassium Carbonate-Based Absorption Process


with Crystallization-Enabled High-Pressure Stripping for CO2
Capture: VaporLiquid Equilibrium Behavior and CO2 Stripping
Performance of Carbonate/Bicarbonate Aqueous Systems
Shihan Zhang, Xinhuai Ye, Yongqi Lu*
Advanced Energy Technology Initiative, Illinois State Geological Survey, Prairie Research Institute, University of Illinois at Urbana-Champaign,
615 E Peabody Drive, Champaign, Illinois 61820, U.S.A.

Abstract

A novel absorption-based post-combustion CO2 capture process is being developed to reduce the energy use associated with CO2
stripping and compression. The process features the use of a concentrated potassium carbonate/bicarbonate (PCB) aqueous
solution as a solvent and high-pressure CO2 stripping with a bicarbonate-dominant slurry enabled by crystallization of the CO2-
rich solution. Vaporliquid equilibrium data under the typical stripping conditions of the process were measured and CO2
stripping tests were performed under various pressures, temperatures, PCB concentrations, and CO2 loading levels using a bench
packed-bed stripping column. The results showed that CO2 stripping with a highly concentrated, highly CO2-loaded PCB
solution/slurry at a high temperature was favored to attain a high stripping pressure and a low water vapor-to-CO2 pressure ratio
in the product gas stream. The total heat use measured for CO2 stripping with the PCB system was up to three times lower than
that of the benchmark 5 M monoethanolamine system under favorable conditions.
2014
2013TheTheAuthors.
Authors. Published
Published by Elsevier
by Elsevier Ltd. is an open access article under the CC BY-NC-ND license
Ltd. This
Selection and peer-review under responsibility of GHGT.
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Organizing Committee of GHGT-12
Keywords: CO2 stripping, vaporliquid equilibrium, potassium carbonate, potassium bicarbonate, absorption, crystallization

1. Introduction

Absorption-based processes are believed to be suitable for large-scale post-combustion CO2 capture from coal-

* Corresponding author. Tel.: +1-217-244-4985; fax: +1-217-333-8566.


E-mail address: yongqilu@illinois.edu

1876-6102 2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Organizing Committee of GHGT-12
doi:10.1016/j.egypro.2014.11.074
666 Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675

fired power plants. However, the state-of-the-art monoethanolamine (MEA) absorption processes are highly energy-
intensive and costly [1,2]. One important cause is associated with the low pressure of CO2 stripping (1 to 2 atm) [3],
resulting in great amounts of heat loss due to water evaporation and the compression work required to compress CO2
to a sequestration-ready pressure. Thus, increasing the stripping pressure is a viable option for reducing the energy
use in CO2 absorption processes.
A novel hot carbonate absorption process (Hot-CAP) with crystallization-enabled high-pressure stripping is
currently being developed to reduce the energy use associated with CO2 stripping and compression. Hot-CAP uses a
concentrated potassium carbonate/bicarbonate (K2CO3/KHCO3, or PCB) aqueous solution as a solvent. The process
involves three major unit operations, i.e., CO2 absorption, bicarbonate crystallization, and CO2 stripping. A
schematic diagram of the process is displayed in Fig. 1 [4]. This process features high-pressure CO2 stripping with a
bicarbonate slurry containing a much higher CO2 loading than a homogeneous solution, enabled by crystallization of
the CO2-rich solution. As a result, it reduces 1) the stripping heat associated with water vaporization, 2) the sensible
heat because of the low specific heat capacity of the slurry, and 3) CO2 compression work.

Cleaned
K2CO3/KHCO3 lean SO42- High-pressure
flue gas
solution (40 wt% removal
CO2 (t6-8 atm)
K2CO3-equiv.,
15-20% CTB)
KHCO3 slurry
(t50 wt%)
Absorption
column Cross-heat High-pressure
(60-80 qC) exchanger stripper
Cross-heat (t140 qC)
Crystallization
tank exchanger
Flue gas (30-35 qC) Steam from
IP turbine

Hydro
cyclone Reboiler
K2CO3/KHCO3 K2CO3/KHCO3
rich solution semi-lean solution
(40-50% CTB) (d70% CTB)
Slurry pump

Fig. 1. Schematic diagram of the Hot-CAP process (CTB: carbonate-to-bicarbonate conversion).

Several other CO2 absorption processes involving crystallization or precipitation have also been reported
recently. The CO2CRC UNO MK3 process is a precipitating PCB-based system claimed to achieve significant
reductions in cost owing to the reduced solvent flow and equipment size, elimination of the SOx and NOx removal
equipment, lower solvent replacement, and production of valuable desulfurization by-products [5]. The process was
tested with both promoted and un-promoted 20 to 40 wt% potassium carbonate solutions using a slipstream of flue
gas from a power plant [6]. Shell has tested a precipitating PCB-based technology at a pilot scale (25 kg CO2/day)
[7]. Researchers in the Netherlands have investigated two process concepts based on the use of precipitating amino
acid salts [8,9]. These studies have concluded that the precipitating processes require less energy use for CO2
stripping to varying extents than do the MEA-based processes. Compared with these processes, Hot-CAP allows for
the use of a more concentrated solution for CO2 absorption and attains a higher pressure for CO2 stripping, which
thus has greater potential to improve the overall energy use for CO2 capture.
Knowledge of the phase equilibria of CO2-H2O-PCB systems at elevated pressures and temperatures is critical
for the design and evaluation of Hot-CAP. However, vaporliquid equilibrium (VLE) data under typical Hot-CAP
stripping conditions are not available in the literature, except for limited data for PCB concentrations up to 40 wt%
at temperatures up to 140 C [10,11]. It is also necessary to investigate the high-pressure CO2 stripping performance
of PCB solutions/slurries in a column setup to gain a better understanding of the kinetic behavior, energy use, and
operating issues of the stripping process. In this study, the VLE data were measured using a stirred cell reactor, and
CO2 stripping tests were conducted using a bench-scale packed-bed column for 40 to 60 wt% PCB solutions/slurries
with carbonate-to-bicarbonate (CTB) conversion levels up to 90% at temperatures between 120 and 200 qC.
Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675 667

2. Experimental methods

2.1. VLE measurements

A schematic of the experimental setup used to measure the VLE data of the CO2-H2O-K2CO3/KHCO3 systems
is shown in Fig. 2. The system consisted of a high-pressure stirred cell reactor, a temperature controller, an inert gas
(N2) supply, and a gas chromatograph (GC). The reactor (Parr Instrument Company, model 4531) had an inner
diameter (ID) of 10 cm and a depth of 14 cm, and was equipped with a variable-speed magnetic stirrer to provide
stirring speeds ranging from 0 to 600 rpm for both the gas and liquid phases. The reactor was also equipped with
multiple valves and gauges, including a gas inlet valve, gas release valve, liquid sampling valve, pressure gauge, and
safety rupture disc. An external furnace was attached to heat the reactor to the desired temperatures through a
separate Parr Series 4840 Temperature Controller. A vacuum pump (Dekker, RVL002H-01) was used to remove
residual gases, if any, from the system before each measurement. The pressure of the reactor was measured by a
pressure transducer (Omega, PX409-1.0KAUSB) and recorded by a computer. The GC (Shimadzu GC-2014),
equipped with an RT-Q-Bond Plot column, was used to measure gas compositions. Gas samples were obtained
using a 100-mL syringe (Restek Corporation) through a 10-mm septum fit into a tee tubing gas sampling port.

To DAQ

PrC

GC

TC PrT
Regulator
Parr reactor
DAQ
Temperature
N2 controller

Fig. 2. Experimental setup for the VLE measurement (PrT: pressure transducer; PrC: pressure controller; TC: thermocouple; GC: gas
chromatograph; DAQ: data acquisition).

The measurement was based on a desorption method. In a typical operation, 1,000 g of PCB slurry with the
desired composition was fed to the Parr reactor. After careful assembly and leakage checking, the system was
vacuumed to the desired level at room temperature. The stirrers for both the gas and liquid phases were turned on to
250 rpm and the furnace heater was turned on to heat the reactor to the desired temperature. When the system
reached an equilibrium state (in 3 to 4 hours), as evidenced by little pressure change over time, a pure N2 gas stream
was introduced into the reactor as an inert tracer to a predetermined pressure. After stabilization, the partial pressure
of N2 (PN2) in the reactor was determined by the difference between the total pressures before and after N2 gas
injection. The partial pressure of CO2 (PCO2) was determined based on the concentration ratio of CO2 to N2 in the gas
samples measured by the GC and the N2 gas partial pressure in the reactor. The partial pressure of water vapor
(PH2O) was calculated from the measured total pressure and the partial pressures of CO2 and N2. The liquid
composition at equilibrium was assumed to be equal to the initial composition because the change in PCB slurry
composition was negligible with CO2 desorption and water vaporization during the measurement.

2.2. CO2 stripping tests

A bench-scale stripping column system rated at 200 C and 34.0 atm was used for the stripping tests. The
system was composed of a stripping column, a liquid supply tank, and instrumentation for measuring temperature,
pressure, and gas and liquid flow rates (Fig. 3). The stripping column was a 2.54-cm-ID, 2.1-m-high stainless steel
tube packed with a 1.8-m height of pall rings (Hai-Yan New Century Petrochemical Device Co., Ltd.). A stainless
steel reboiler was attached to the bottom of the column, with heat supplied via three 1.0-kW mica band heaters
668 Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675

(Tutco, Model BH91911) with a temperature control and power input control. The liquid level in the reboiler was
monitored by a Klinger magnetic level gauge (Davis Control Ltd.). The temperature profile along the column was
measured by five evenly mounted thermocouples, and the pressure drop was measured by two pressure gauges
installed at the top and bottom of the column. The CO2 product gas stream exiting the top of the column was cooled
by passing it through a stainless steel cooling coil to condense the water vapor. The flow rate of the dried CO2 gas
was measured by a flow meter after the pressure was relieved by a back-pressure controller. The regenerated lean
solution was discharged through a release valve at the bottom to a liquid container.

5 PRD
Gas
Vent cooler P
T
FM PCr Cooling FM TCr
P
6 4
water 3
Water
7 knockout Hot water
T
2
Heater with
temperature
Gas
control
analyzer T
Stripping
column
1
Data T, P monitoring
logger & control T
High-pressure
stirred tank

Liquid T
level P
Discharge indicator PCr
container TCr
~
Reboiler heater
CO2
Cooling 8 with amp and
volt recorder cylinder
water

T: Thermal couple; P: Pressure transducer; FM: Flow meter


TCr: Temperature controller; PRD: Pressure rupture disk; PCr: Pressure regulator
Fig. 3. Schematic diagram of a bench-scale, high-pressure stripping column experimental system.

The PCB solution/slurry was continuously fed to the stripping column from a liquid supply tank. The supply
tank was a stainless steel cylinder with an ID of 0.3 m and a height of 0.9 m. The flow rate of the feed solution was
controlled by adjusting the pressure difference between the tank (with the headspace balanced with the N2 gas from
a cylinder under an adjustable pressure) and the stripping column and was monitored by a high-pressure liquid flow
meter (Micro Motion, Inc., Model F025). Two 2.5-kW electrical mica band heaters with temperature control (Tutco,
Model BH91910) were attached to heat the liquid in the tank to the desired temperatures. The tank was equipped
with a variable-speed magnetic drive agitator controlled through an inverter motor speed regulator (PDC Machines,
Inc., Model 1699-0113) to provide stirring speeds up to 500 rpm. A built-in Klinger magnetic level gauge (Davis
Control Ltd.) was used to monitor the liquid level in the tank.
In a typical experiment, approximately 20 L of PCB slurry with the desired composition was charged to the
liquid supply tank. The slurry was preheated to the desired temperature, about 5 to 10 C lower than the reboiler
temperature but sufficiently high to dissolve the solids in the tank. Approximately 1.5 L of a randomly diluted PCB
was initially charged to the reboiler and preheated to the desired temperature. Once the temperatures and pressures
in both the supply tank and stripping column were stable, the hot PCB solution in the supply tank was continually
fed to the stripping column at a controlled rate, ranging between 0.05 and 0.12 L/min. The liquid level of the
reboiler was held constant during the experiment. Each experiment was operated continuously for more than 40 to
60 min (>3 times greater than the residence time of the liquid in the reboiler) after it reached steady state, as
indicated by negligible changes in the readings of temperature, pressure, and flow rate of the dried product CO2. The
lean PCB solution exiting the reboiler was sampled for analysis of the CO2 loading by an acid titration method using
a Chittick apparatus.
Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675 669

3. Results and discussion

3.1. Phase equilibrium behavior

The VLE data for the 40 wt% PCB solution with various CTB conversion levels at 140 C was measured and
compared with those reported by Tosh et al. [11]. At a CTB conversion level below 50%, the data measured in this
study were comparable; however, at a CTB conversion level above 50%, both the total pressures (sum of PCO2 and
PH2O) and the PCO2 values measured in this study were lower. To confirm the measurement, the vapor pressures over
pure water at 140 to 200 C were measured using the same VLE cell, and the measured values closely matched
those in the literature. Thus, the measurements in this work were deemed trustworthy.
Fig. 4 shows the VLE data for the 40, 50, and 60 wt% PCB solutions with different CTB conversion levels at
140 to 200 qC. The curves of PCO2 as a function of CTB conversion are reverse S-shaped. Note that the PCB
solutions containing high concentrations of KHCO3 are slurries at room temperature but become soluble at high
temperatures (140 to 200 qC).

a a' 1200
1000 40 wt% PCB
1000 140qC
160qC
CO2 partial pressure, kPa

Water vapor pressure, kPa


800 180qC
100
600

40 wt% PCB 400


10
140qC
160qC 200
180qC
1 0
10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
K2CO3 converted, % K2CO3 converted, %
b b' 1200
50 wt% PCB
1000 1000 140qC
160qC
CO2 partial pressure, kPa

Water vapor pressure, kPa

180qC
800 200qC
100
600

50 wt% PCB 400


10 140qC
160qC
180qC 200
200qC
1 0
10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
K2CO3 converted, % K2CO3 converted, %
c c' 1200
60 wt% PCB
1000 1000 160qC
CO2 partial pressure, kPa

Water vapor pressure, kPa

180qC
800
200qC
100
600

50 wt% PCB
60 400
10 160qC
180qC 200
200qC
1 0
10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
K2CO3 converted, % K2CO3 converted, %
Fig. 4. Equilibrium partial pressures of CO2 and water vapor for 40 wt% (a and ac), 50 wt% (b and bc), and 60 wt% (c and cc) PCB solutions at
140, 160, 180, and 200 qC.

As shown in Fig. 4, a PCB solution with a higher CO2 loading was favored to obtain a higher total pressure and
670 Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675

a lower water vapor-to-CO2 pressure ratio (PH2O/PCO2). For example, at 160 qC, the total pressure increased from
538 to 954 kPa and the PH2O/PCO2 decreased from 1.90 to 0.35 as the CTB conversion level of the 50 wt% PCB
solution was increased from 50 to 73%. Increasing the temperature increased the total pressure but did not
substantially change the PH2O/PCO2. When the temperature was increased from 160 to 200 qC, the total pressure over
the 50 wt% PCB solution with 73% CTB conversion increased from 954 to 2,020 kPa, whereas the PH2O/PCO2
remained almost unchanged.
Note that at the same temperature and CTB conversion level, the PCO2 increased with increasing PCB
concentration. However, the PCO2 increase was not so substantial as that with an increasing CTB conversion,
especially when the PCB concentration reached 50 wt% or above. At the same time, the value of PH2O decreased
remarkably as the PCB concentration was increased from 40 to 60 wt%, suggesting that the stripping heat associated
with water vaporization can be lowered by using a more concentrated PCB feed. For example, at 160 qC, the
PH2O/PCO2 for the 60 wt% PCB with 83% CTB conversion was 0.05 compared with 0.11 for the 50 wt% PCB with
the same CTB conversion.
A higher stripping pressure and a lower PH2O/PCO2 will reduce both the use of CO2 stripping heat and the
compression work requirement. The measured VLE data suggest that CO2 stripping from a concentrated PCB
solution with a higher CTB conversion level at a high temperature thermodynamically favored achieving these
desirable conditions.

3.2. Performance of CO2 stripping

CO2 stripping was tested in the stripping column under different temperatures, PCB concentrations, and inlet
CTB conversion levels. The PH2O values at the top of the column were estimated from the measured VLE data for 40
to 60 wt% PCB solutions or from the data in the literature for 30 wt% PCB solutions [11] under the assumption that
water evaporation reached an equilibrium state. The PCO2 and PH2O/PCO2 at the top of the column were consequently
calculated from the total stripping pressure measured.
The heat used for CO2 stripping in Hot-CAP is the sum of the following heat elements, assuming the solution
does not flash upon entrance to the stripper:
Qtotal Qsensible  Qstripping  Qreaction  Qdissolutio n , (1)

where Qsensible is the sensible heat required to heat the inlet CO2-rich solution to the reboiler temperature (kJ/kg),
Qstripping is the stripping heat associated with water evaporation (kJ/kg), Qreaction is the heat of reaction required to
desorb the CO2 from the rich solution (kJ/kg), and Qdissolution is the heat taken to dissolve KHCO3 solids in the slurry
(kJ/kg). Phase equilibrium is usually not reached in the stripping column. The four heat elements can be calculated
based on the measured process parameters according to Eqs. (25):
(Treboiler  Tinlet ) M solution , (2)
Qsensible Cp
'CTB x solv M CO2

M H 2O PH 2O , (3)
Qstripping 'H vap , H 2O
M CO2 PCO2

Q reaction 'H abs ,CO 2 , (4)

2 M KHCO3
Qdissolution 'H disl , KHCO3 , (5)
M CO2

where Cp is the specific heat capacity of the PCB solution (kJ/kgK); Treboiler and Tinlet are the temperatures of the
reboiler and the inlet feed solution (K); CTB is the change in CTB conversion of the PCB solution during stripping
(%); xsolv is the molar fraction of PCB (K2CO3-equivalent) in the solution; 'Hvap,H2O is the heat of water evaporation
(kJ/kg); PH2O and PCO2 are the partial pressures of water vapor and CO2 at the top of the stripping column (kPa);
'Habs,CO2 is the heat of absorption (kJ/kg); 'Hdisl,KHCO3 is the heat of KHCO3 dissolution (kJ/kg); and Msolution, MH2O,
MCO2, and MKHCO3 are the molecular weights of the solution, water, CO2, and KHCO3, respectively.
Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675 671

Note that in the current study, the reboiler heat duty in the stripping column was attributed to only Qsensible,
Qstripping, and Qreaction because the Qdissolution associated with the dissolution of KHCO3 solids occurred in the liquid
supply tank. For the MEA-based processes, the Qdissolution term is not applied because no dissolution process is
involved.

3.2.1 Effect of temperature

As expected, higher temperatures released more CO2 from the hot PCB solution because of the higher driving
force for CO2 stripping, resulting in a leaner PCB solution exiting the stripping column. For example, with the
PCB30-80 feed, the CTB conversion of the regenerated solution decreased from 76 to 65% as the stripping
temperature was increased from 120 to 160 C [Fig. 5(a)].

a 80 b 700 c 6000
PCB30-80 PCB30-80 PCB30-80
CTB conversion of regenerated

70

Reboiler heat duty, kJ/kg CO2


600 CO2 partial pressure 5000 Reaction heat
60 Water vapor partial pressure Stripping heat
500
4000 Sensible heat
Pressure, kPa

50
PCB, %

400
40 3000
300
30
2000
200
20
100 1000
10
0 0 0
T=120 T=140 T=160 T=120 T=140 T=160 T=120 T=140 T=160
Stripping temperature, qC Stripping temperature, qC Stripping temperature, qC
Fig. 5. (a) Compositions of regenerated lean PCB solutions, (b) partial pressures of CO2 and water vapor at the top of the column, and (c) reboiler
heat duty for CO2 stripping with the PCB30-80 feed solution at different stripping temperatures.

As shown in Fig. 5(b), the PCO2 at the top of the column with the PCB30-80 feed increased from 200 to 441 kPa
as the stripping temperature was increased from 120 to 160 C. Note that the PCO2 was affected by both the
composition of the lean solution and the stripping temperature. A higher temperature corresponded to a higher CO2
equilibrium pressure; however, it also produced a leaner regenerated solution, which corresponded to a lower
equilibrium CO2 pressure. The results indicate that the net effect of increased temperature favored an increase in
CO2 partial pressure.
The PCO2 and PH2O data displayed in Fig. 5(b) also suggest that the PH2O/PCO2 in the gas stream exiting the
stripping column increased slightly from 0.59 to 0.66 as the temperature was increased from 120 to 160 C. On the
one hand, because the heat of CO2 absorption in the PCB solution (28 kJ/mol) is lower than the heat of water
evaporation (41 kJ/mol), the PH2O should increase more significantly than the PCO2 with increasing temperature,
resulting in an increase in PH2O/PCO2. On the other hand, for concentrated PCB solutions with high CTB conversion
levels, the PH2O at a higher temperature decreased from that of pure water more significantly than did the PH2O at a
lower temperature (Fig. 4). As a result, the combined effect of temperature on the PH2O/PCO2 was not pronounced for
the concentrated PCB solutions with high CTB conversion levels.
The results of reboiler heat duty for CO2 stripping with the PCB30-80 feed are displayed in Fig. 5(c). The heat
duty decreased with increasing stripping temperature from 120 to 160 qC. Compared with the stripping heat or
reaction heat use, the sensible heat use was more dependent on the temperature. A greater amount of CO2 was
stripped off at a higher stripping temperature, resulting in a proportional decrease in the sensible heat use on a per-
unit mass of CO2 release basis. Meanwhile, increasing the temperature slightly increased the stripping heat use as
the PH2O/PCO2 slightly decreased. Note that the reaction heat did not depend on the temperature. Overall, the reboiler
heat duty was reduced by approximately one-half when the temperature was increased from 120 to 160 qC.

3.2.2 Effect of PCB concentration

Fig. 6(a) shows the lean CO2 loading obtained from the stripping operation at 160 C for the 30 to 50 wt% PCB
solutions with 80% CTB conversion in the feed. The higher the PCB concentration, the greater the amount of CO2
672 Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675

desorbed, indicating that a more concentrated PCB feed favored a deeper level of solvent regeneration. The change
in CTB conversion throughout the column for the 30 to 50 wt% PCB varied from 14 to 24 percentage points, which
is equivalent to an 18 to 30% regeneration efficiency of KHCO3 contained in the feed solution.

a 80 b 700 c 2500
160 qC stripping 160 qC stripping 160 qC stripping
CTB conversion of regenerated

70 CO2 partial pressure

Reboiler heat duty, kJ/kg CO2


600 Reaction heat
Water vapor partial pressure 2000
60 Stripping heat
500
Sensible heat

Pressure, kPa
50
PCB, %

400 1500
40
300 1000
30
200
20
500
10 100

0 0 0
PCB30-80 PCB40-80 PCB50-80 PCB30-80 PCB40-80 PCB50-80 PCB30-80 PCB40-80 PCB50-80
PCB feed solution PCB feed solution PCB feed solution
Fig. 6. (a) Compositions of regenerated lean PCB solutions, (b) partial pressures of CO2 and water vapor at the top of the column, and (c) reboiler
heat duty for CO2 stripping at 160 C with different PCB feed solutions.

As illustrated in Fig. 6(b), both a higher total pressure (PH2O + PCO2) and a higher CO2 partial pressure were
attained for the more concentrated PCB solution. At 160 qC, when the total PCB concentration was increased from
30 to 50 wt% with the same 80% CTB conversion in the feed, the total stripping pressure increased from 731 to 834
kPa, whereas the PH2O/PCO2 ratio decreased from 0.66 to 0.26. This result verified the findings from the VLE
measurements.
Using a more concentrated PCB feed solution also led to a lower reboiler heat duty for CO2 stripping, as shown
in Fig. 6(c). The reboiler heat duty for CO2 stripping at 160 qC decreased from 2,323 to 1,218 kJ/kg CO2 when the
PCB concentration was increased from 30 to 50 wt% with the same initial 80% CTB conversion. Such a trend was
attributed to both the reduced sensible heat and stripping heat use. The sensible heat use was decreased because a
greater amount of CO2 was stripped off when a more concentrated solution was used. The stripping heat use for the
PCB50-80 feed was lower than that for the PCB30-80 or PCB40-80 feed because the PH2O/PCO2 obtained was the
lowest, as mentioned.

3.2.3 Effect of CO2 loading of the feed solution

The effect of the initial CO2 loading on the level of solvent regeneration for the 40 wt% PCB feed solution at
160 C is illustrated in Fig. 7(a). The initial CTB conversion of the solution was varied from 40 to 80%. For a richer
feed solution (higher CTB conversion), a greater amount of KHCO3 was regenerated during stripping. The change in
CTB conversion along the stripping column increased from 9.5 to 18 percentage points when the initial CTB
conversion was increased from 40 to 80%.

a 80 b 700 c 6000
160 qC stripping 160 qC stripping 160 qC stripping
CTB conversion of regenerated

70
Reboiler heat duty, kJ/kg CO2

600 CO2 partial pressure 5000 Reaction heat


60 Water vapor partial pressure Stripping heat
500
4000 Sensible heat
Pressure, kPa

50
PCB, %

400
40 3000
300
30
2000
200
20
100 1000
10
0 0 0
PCB40-40 PCB40-60 PCB40-80 PCB40-40 PCB40-60 PCB40-80 PCB40-40 PCB40-60 PCB40-80
PCB feed solution PCB feed solution PCB feed solution
Fig. 7. (a) Compositions of regenerated lean PCB solutions, (b) partial pressures of CO2 and water vapor at the top of the column, and (c) reboiler
heat duty for CO2 stripping at 160 C with the 40 wt% PCB feed at three different levels of CTB conversion.
Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675 673

As displayed in Fig. 7(b), the use of a feed solution with a higher CTB conversion level achieved a higher PCO2
and a lower PH2O in the CO2 product gas stream. As the CTB conversion of the 40 wt% PCB feed was increased
from 40 to 80%, the total stripping pressure increased from 545 to 793 kPa and the PCO2 increased from 108 to 538
kPa at 160 C (i.e., the PH2O/PCO2 decreased from 4.03 to 0.47). Such a tendency is consistent with that observed in
the VLE measurements.
As expected, using a richer PCB feed solution led to a lower reboiler heat duty for CO2 stripping. As shown in
Fig. 7(c), when the CO2 loading of the 40 wt% PCB feed was enriched from 40 to 80% CTB conversion, the heat
duty for stripping at 160 qC decreased from 5,359 to 1,677 kJ/kg CO2. The results in Fig. 7(c) also show that the
sensible heat use was comparable for the 40 wt% feed solution with either 40, 60, or 80% CTB conversion.
Therefore, the lower reboiler heat duty attained for the richer feed solution was attributed mainly to the lower
stripping heat use because the PH2O/PCO2 for such a system was significantly lower than that for the leaner feed
solution.

3.2.4 Comparison with MEA stripping

For comparison purposes, a 5 M MEA solution loaded with 0.475 mol CO2/mol MEA (i.e., 95% MEA
conversion, MEA5-95) was tested in the same stripping column. The tests with MEA were operated at or near
atmospheric pressure, a slightly lower pressure than but still close to typical MEA stripping pressures. The CO2
loading levels of the lean MEA solution were 0.37 and 0.20 mol CO2/mol MEA at stripping temperatures of 100 and
108 qC, respectively. Similar to the PCB system, the higher stripping temperature facilitated a deeper level of MEA
regeneration and a lower PH2O/PCO2 in the product gas stream, as demonstrated in Fig. 8(a).

a 10 b 7000
Stripping at 160 qC for PCB Total pressure, atm Stripping at 160 qC for PCB
9 and 100 or 108 qC for MEA
Total pressure, atm; PH2O/PCO2

and 100 or 108 qC for MEA PH2O/PCO2 6000


Reboiler heat duty, kJ/kg CO2

8 Dissolution heat
Reaction heat
7 5000 Stripping heat
6 Sensible heat
4000
5
4 3000

3 2000
2
1000
1
0 0

PCB or MEA feed solution PCB or MEA feed solution

Fig. 8. Comparison of CO2 stripping at 160 qC with the 40 and 50 wt% PCB feeds and at 100 qC with the MEA5-95 feed: (a) stripping pressure
and PH2O/PCO2 and (b) total heat use [MEA5-90(1): 100 qC stripping and lean CO2 loading of 0.37 mol CO2/mol MEA; MEA5-90(2): 108 qC
stripping and lean CO2 loading of 0.20 mol CO2/mol MEA].

The reboiler heat duty for the MEA stripping was determined based on the same approach used for the PCB
stripping. The heat duty for the MEA stripping amounted to 6,200 kJ/kg CO2 at 100 qC and 4,300 kJ/kg CO2 at 108
qC. The leaner the CO2 loading of the regenerated solution, the lower the PH2O/PCO2 in the product gas stream,
resulting in a greater amount of stripping heat use associated with water vaporization [Fig. 8(b)]. The heat duty
obtained for the MEA stripping in this study is comparable with those reported in the literature [12,13].
A comparison of the total heat use for CO2 stripping with the PCB and the MEA systems is displayed in Fig.
8(b). As mentioned, the reboiler heat duty for the PCB system [as displayed in Figs. 5(c), 6(c), and 7(c)] did not
include the term for dissolution heat Qdissolution. Thus, the total heat use for CO2 stripping with PCB should be the
sum of the reboiler heat duty and the heat of dissolution, as indicated by Eq. (1). Based on the enthalpy change of
KHCO3 dissolution, the heat of dissolution on a CO2 basis was estimated at 42 kJ/mol CO2 stripped (955 kJ/kg) for
the PCB system.
As shown in Fig. 8(b), the total heat use for 160 qC stripping from the 40 to 50 wt% PCB feed solutions with 40
674 Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675

to 80% initial CTB conversion ranged from 2,172 to 6,313 kJ/kg. The total heat use for CO2 stripping was lower for
a more concentrated PCB solution or a feed solution with richer CO2 loading. These results clearly demonstrate that
the total heat use for the PCB system was lower than that for the MEA system. For example, the total heat use for
CO2 stripping from the 40 or 50 wt% PCB feed solution with 80% CTB conversion was about 1.5 to 3 times lower
than that for the MEA. In addition, the total pressures attained for PCB stripping were higher than those for MEA
stripping [(Fig. 8(a)], suggesting less CO2 compression work is required for the PCB system.

4. Conclusions

Vaporliquid equilibrium (VLE) data were measured for the 40 to 60 wt% PCB solutions with CTB conversion
levels varying from approximately 20 to 90% at 140 to 200 C, which covered the full range of CO2 stripping
conditions in Hot-CAP. A high concentration or a higher CTB conversion level of the PCB solution was favored to
obtain a higher total pressure and a lower PH2O/PCO2. An increase in temperature facilitated obtaining a higher total
pressure but did not significantly change the PH2O/PCO2. These results suggest that using a concentrated, CO2-rich
PCB solution/slurry for CO2 stripping at high temperatures thermodynamically favors a reduction in energy use for
CO2 stripping.
A bench-scale packed-bed stripping column was used to investigate the performance of CO2 stripping for the
PCB system. The effects of stripping temperature, PCB concentration, and initial CO2 loading on the stripping
pressure and PH2O/PCO2 observed in the column tests followed trends similar to those in the VLE measurements.
Increasing the stripping temperature, PCB concentration, or initial CTB conversion significantly decreased the total
heat use for CO2 stripping. The PCB system could reduce the total heat use by up to three times under favorable
conditions compared with the 5 M MEA system. The total heat use for 160 qC stripping from the 40 to 50 wt% PCB
feed solutions with 40 to 80% CTB conversion ranged from 2,172 to 6,313 kJ/kg CO2, compared with 4,300 to
6,200 kJ/kg CO2 for 100 to 108 qC stripping from a rich 5 M MEA solution. In addition, the PCB system allowed
higher total stripping pressures than did the MEA system, thus requiring less compression work after stripping to
deliver supercritical CO2 to a storage facility.

Acknowledgements

The authors gratefully acknowledge funding support for this work by the Department of Energys National
Energy Technology Laboratory (DOE/NETL) under Award Number DE-FE0004360 and by the Illinois Department
of Commerce and Economic Opportunity (IDCEO) through the Office of Coal Development (OCD) and the Illinois
Clean Coal Institute (ICCI) under Contract Number 11/US-6.

References

[1] Rochelle GT. Amine scrubbing for CO2 capture. Science 2009; 325:16521654.
[2] Singh D, Croiset E, Douglas PL, Douglas MA. Techno-economic study of CO2 capture from an existing coal-fired power plant: MEA
scrubbing vs. O2/CO2 recycle combustion. Energy Convers Manage 2003; 44(19):30733091.
[3] Abu-Zahra MR, Schneiders LH, Niederer JP, Feron PH, Versteeg GF. CO2 capture from power plants: Part I. A parametric study of the
technical performance based on monoethanolamine. Int J Greenh Gas Control 2007; 1:3746.
[4] Lu Y, Zhang S, Devries N, Sahu M, Ye X, Hirschi J, Jones A. Experimental studies of a carbonate-based absorption process with high
pressure stripping for post-combustion CO2 capture. In: Sakkestad BA, editor. Proceedings of the 38th International Technical Conference
on Clean Coal & Fuel Systems, North Potomac, MD: Coal Technologies Associates; 2013.
[5] Anderson C, Ho M, Harkin T, Wiley D, Hooper B. Large scale economics of a precipitating potassium carbonate CO2 capture process for
black coal power generation. Greenh Gases: Science & Technol 2014; 4(1):819.
[6] Anderson C, Harkin T, Ho M, Mumford K, Qader A, Stevens G, Hooper B. Developments in the CO2CRC UNO MK3 process: A multi-
component solvent process for large scale CO2 capture. Energy Procedia 2013; 37:225232.
[7] Moene R, Schoon L, van Straelen J, Geuzebroek F. Precipitating carbonate process for energy efficient post-combustion CO2 capture. Energy
Procedia 2013; 37:18811887.
[8] Sanchez-Fernandez E, Heffernan K, van der Ham LV, Linders MJ, Eggink E, Schrama FN, Vlugt TJ. Conceptual design of a novel CO2
capture process based on precipitating amino acid solvents. Ind Eng Chem Res 2013; 52(34):1222312235.
Shihan Zhang et al. / Energy Procedia 63 (2014) 665 675 675

[9] Sanchez-Fernandez E, Heffernan K, van der Ham L, Linders MJ, Brilman DWF, Goetheer EL, Vlugt TJ. Analysis of process configurations
for CO2 capture by precipitating amino acid solvents. Ind Eng Chem Res 2014; 53(6):23482361.
[10] Benson HE, Field JH, Haynes WP. Improved process for carbon dioxide absorption uses hot carbonate solutions. Chem Eng Progr 1956;
52:433438.
[11] Tosh JS, Field JH, Benson HE, Haynes WP. Equilibrium study of the system potassium carbonate, potassium bicarbonate, carbon dioxide,
and water. Washington, DC: US Bureau of Mines, Report of Investigations 5484; 1959.
[12] Sakwattanapong R, Aroonwilas A, Veawab A. Behavior of reboiler heat duty for CO2 capture plants using regenerable single and blended
alkanolamines, Ind Eng Chem Res 2005; 44:44654473.
[13] Ciferno J, Dipietro P, Tarka T. An economic scoping study for CO2 capture using aqueous ammonia. Pittsburgh, PA: Department of
Energy/National Energy Technology Laboratory; 2005.

You might also like