Vous êtes sur la page 1sur 9

chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Modelling reactive absorption of CO2 in packed columns for


post-combustion carbon capture applications

F.M. Khan 1 , V. Krishnamoorthi 2 , T. Mahmud ∗


Institute of Particle Science and Engineering, School of Process, Environmental and Materials Engineering,
The University of Leeds, Leeds LS2 9JT, UK

a b s t r a c t

A rate-based process model for the reactive absorption of carbon dioxide (CO2 ) from a gas mixture into an aqueous
monoethanolamine (MEA) solution in a packed column is developed. The model is based on the fast second-order
kinetics for the CO2 –MEA reactions and takes into account the mass transfer resistances. The heat effects associated
with the absorption and chemical reaction are included through energy balances in the gas and liquid phases.
Appropriate correlations for the key thermodynamic and transport properties and for the gas–liquid mass transfer
are incorporated into the model to ensure reliable predictions. The model predictions are validated by simulating
a series of experiments conducted in pilot and industrial scale absorption columns with random and structured
packings reported in the literature. Comparisons between the simulation results and the experimental data reveal
good quality predictions of the gas phase CO2 and MEA concentrations and the liquid temperature along the column
height. The sensitivity studies reveal that the correlations for the gas- and liquid-film mass transfer coefficients given
by Onda et al. (1968) provide better predictions than the penetration theory of Higbie (1935) and the correlation of
Bravo et al. (1985).
© 2010 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords: Carbon capture; CO2 –MEA absorption; Reactive absorption; Process modelling

1. Introduction from flue gases which would otherwise be vented to the atmo-
sphere (Wilson and Gerard, 2007). The latter technology is
In recent years, there has been increasing demand for a sig- being considered by the power generation industry as one of
nificant reduction of carbon dioxide (CO2 ) emissions from the potential options. In principle, CO2 can be removed from
industrial sources to alleviate the global warming problem. a gas stream using several separation processes including
Following the 1997 Kyoto Protocol, the European Union has physical/chemical absorption into a liquid solvent, adsorp-
set a target of 20% reduction of CO2 emission by the year tion on solids, permeating through membranes, and chemical
2020. In order to meet this target, a significant reduction conversion. Reactive absorption using aqueous alkanolamine
of CO2 release from fossil fuel fired thermal power plants, solutions, including monoethanolamine (MEA) solutions, in
which contribute approximately 25% to the total global emis- packed columns are widely used for the separation of CO2 from
sions (WRI, 2007), will be required. This can be achieved by process gas streams in the chemical and petroleum industries.
adopting an effective strategy for carbon capture and stor- This is a well-developed technology and is currently the most
age (CCS) such as the pre-combustion CO2 capture technology preferred process approach for the CO2 capture from power
as used in the integrated gasification and combined-cycle plant flue gases (Freund, 2003; Idem and Tontiwachwuthikul,
(IGCC) plants, oxyfuel combustion for the production of 2006). However, packed-bed absorption columns are known to
sequestration-ready CO2 or the post-combustion CO2 capture suffer from various operational problems including high gas


Corresponding author at: Institute of Particle Science and Engineering, School of Process, Environmental and Materials Engineering,
The University of Leeds, Engineering (Houldsworth) Building, Leeds LS2 9JT, UK.
E-mail address: T.Mahmud@leeds.ac.uk (T. Mahmud).
Received 15 May 2010; Received in revised form 14 September 2010; Accepted 30 September 2010
1
BP Exploration, 1–4 Wellheads Avenue, Dyce, Aberdeen, AB21 7PB, UK.
2
Hindustan Unilever Research Centre, 64 Main Road, Whitefield, Bangalore-560066, India.
0263-8762/$ – see front matter © 2010 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.cherd.2010.09.020
chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608 1601

Gas-liquid interface the simulation of reactive absorption of CO2 from a mixture of


air–CO2 into aqueous MEA solutions in packed columns under
Gas
pA film
Liquid adiabatic conditions. The model accounts for the mass trans-
film fer resistances in the gas and liquid films and is based on
the fast second-order kinetics for the CO2 –MEA reactions in
CB
pAi the liquid film. The heat effects associated with the absorp-
tion and chemical reaction are taken into account through
A Bulk liquid
energy balances in the gas and liquid phases. Appropriate
CAi correlations for the key thermodynamic and physical proper-
ties and for the mass transfer coefficients and the gas–liquid
interfacial area are incorporated into the model to ensure
Bulk gas CA reliable predictions. The model predictions are validated by
CBi
simulating a number of experiments conducted in packed-
Fig. 1 – Gas–liquid reaction model based on the two-film bed absorption columns of different scale sizes, including
theory of mass transfer (actual concentration profiles in the pilot-plant columns packed with structured packing of 2.21 m
liquid phase depend on the reaction regime). height (Aroonwilas et al., 2003) and ceramic Berl saddles of
Adapted from Levenspiel (1999). 6.6 m height (Tontiwachwuthikul et al., 1992) and an industrial
column of 14.1 m packing height containing stainless steel Pall
phase pressure drop, liquid channelling and flooding of the rings (Pintola et al., 1993).
packing materials resulting in a poor gas–liquid contact. These
problems are prone to occur in the flue gas scrubbing opera- 2. Reactive absorption model formulation
tions, which involve low solvent flow rates and high superficial
gas velocities because of the large gas volumes (e.g., a typi- In order to develop a comprehensive mathematical model for
cal 400 MWe coal-fired power plant produces approximately the absorption of CO2 in aqueous MEA solutions, it is essential
1.1 × 106 Nm3 /h of flue gas (Raynal and Royon-Lebeaud, 2007)) to understand the underlying mechanism of the absorption
with relatively low CO2 concentrations (typically 3–5% for nat- process coupled with chemical reactions. This encompasses
ural gas and 10–15% for coal combustion) and require low mass transfer of CO2 from the bulk gas to the MEA solution,
pressure drop as the flue gas fan typically operates slightly reactions between CO2 and MEA species and the associated
above the atmospheric pressure. In order to achieve high CO2 kinetics regimes.
capture efficiency, there is a need for improved design of The transport of CO2 from the gas to the liquid phase can be
packed-bed absorption columns and optimisation of operat- modelled using the two-film mass transfer theory of Whitman
ing conditions, which can be facilitated via the use of advanced (1923), as depicted in Fig. 1. CO2 is transported through the
process models for reactive absorption. gas-film by molecular diffusion, absorbed in the liquid and
Rigorous theories of absorption with chemical reactions then diffuses through the liquid-film. In principle, the reaction
and the rate expressions for different gas–liquid reac- between a dissolved gas and a solvent can take place at any
tion regimes are well documented by Astarita (1967) and location in the liquid phase, which depends on the kinetics
Danckwerts (1970), which can be used as a basis for devis- regime. However, the CO2 –MEA reaction is characterised as a
ing a process model for the design and simulation of packed fast reaction (Astarita, 1967) which occurs within the liquid-
columns for the CO2 absorption processes. In the past, dif- film only, consequently the concentration of CO2 reduces to
ferent modelling approaches (see review in Kenig et al., 2001) zero somewhere within this region.
ranging from a relatively simple equilibrium stage model The reactive absorption process model developed in this
to more rigorous rate-based model with electrolyte solution study following the approaches used in the earlier work of
chemistry have been used for various gas–liquid systems. A Treybal (1969) and Pandya (1983) consists of a reaction model
rate-based methodology accounting for the interfacial mass using the two-film model for fast reaction incorporated into
transfer and finite rate chemistry is considered as an appro- a gas–liquid contactor model based on the mass and energy
priate approach for the modelling of industrial absorption balances. The model is described below.
columns as phase equilibrium is difficult to achieved in
such systems (Lawal et al., 2009). This approach also allows 2.1. Chemistry of CO2 –MEA system
the effects of column hydrodynamics and internals on the
absorption process to be included via the correlations for The mechanism of reaction between CO2 and amines is highly
the mass transfer coefficients, gas–liquid interfacial area complex and in spite of being studied extensively both the-
and hold-up (Kenig et al., 2001). A general rate-based mod- oretically and experimentally (see for details Astarita, 1967;
elling methodology for the gas absorption with chemical Danckwerts, 1970), it is not well understood (Pintola et al.,
reactions based on the steady-state, adiabatic physical absorp- 1993). The reaction mechanism for the CO2 –MEA system may
tion model of Treybal (1969) was proposed by Pandya (1983). be represented by the following overall reactions (Astarita,
Over the last two decades, a number of mathematical mod- 1967):
els of reactive absorption of CO2 have been developed (e.g.,
Tontiwachwuthikul et al., 1992; Pintola et al., 1993; Aroonwilas Carbamate formation:
et al., 2003; Freguia and Rochelle, 2003; Tobiesen et al., 2007;
Lawal et al., 2009) based on Pandya’s (1983) approach for the CO2 + 2RNH2 → RNHCOO− + RNH3 + (R1)
simulation of CO2 capture processes using alkaline and MEA
solutions. Bicarbonate formation:
The work reported in this paper focuses on the develop-
ment and detail validation of a rate-based process model for CO2 + RNH2 + H2 O → HCO3 − + RNH3 + (R2)
1602 chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608

Reversion of carbamate to bicarbonate: values of E given by Eq. (4) deviate by less than 3% from the
most accurate solution of van Krevelen and Hoftijzer (1948)
RNHCOO− + CO2 + 2H2 O → HCO3 − + 2RNH3 + (R3) (Tontiwachwuthikul et al., 1992).
The gas- and liquid-film mass transfer coefficients and the
where RNH2 is MEA and R is OH·CH2 ·CH2 . Astarita (1967) has interfacial area for random packings are obtained from the
suggested that for CO2 /MEA mole ratio less than 0.5, the rate of widely used correlations provided by Onda et al. (1968):
bicarbonate formation is insignificant and the overall reaction
 0.7  1/3
may be expressed by R1, which is considered to be approxi- kg RT G g −2.0
mately irreversible. The overall reaction rate is second-order, = C1 (av dp ) (5a)
av DAg av g g DAg
first-order with respect to both CO2 and MEA, and is expressed
in terms of the molar concentrations of CO2 and MEA as:   1/3  L 2/3   −0.5
l l 0.4
kol = 0.0051 (av dp ) (5b)
l g aw l l DAl
r = k2 [CO2 ][RNH2 ] (1)

Danckwerts (1970) has suggested that the use of molar con-   0.75  0.1  −0.05  0.2 
aw c L L2 av L2
centrations in the rate equation instead of thermodynamic = 1 − exp −1.45
av l av l l 2 g l l av
activities is valid. The second-order rate constant, k2 , depends
(5c)
on the temperature and is given by Hikita et al. (1977):

2152
log(k2 ) = 10.99 − (2) where av is the total packing surface area per unit volume of
T
packing, aw is the surface area of wetted packing per unit vol-
2.2. Reaction model ume of packing, C1 is a dimensionless constant which depends
on the packing size, DAg is the gas phase diffusivity of CO2 , dp
The overall rate of absorption of CO2 in an aqueous solution of is the nominal packing size, G and L are the flow rates of gas
MEA taking into account the mass transfer and chemical reac- and liquid, R is the universal gas constant, T is the gas temper-
tion rates can be expressed based on the two-film model for a ature,  c is the critical surface tension of packing material and
fast second-order reaction as (Levenspiel, 1999; Doraiswamy,  l is the surface tension of liquid, g and l are the viscosity
2001): of gas and liquid, and g and l are the density of the gas and
liquid. The surface area of wetted packing per unit volume of
pCO2 packing was used in Eq. (3). These parameters for structured
rCO2 = (3)
(1/kg a) + (HCO2 /kol aE) packing can be estimated using the correlations given by Bravo
et al. (1985).
where a is the gas–liquid interfacial area per unit volume of
packing, E is the enhancement factor, HCO2 is the Henry’s law
2.3. Physical and thermodynamic properties
constant, kg and kol are the gas- and liquid-film physical mass
transfer coefficients, respectively, and pCO2 is the partial pres-
The calculation of absorption rates requires data for the phys-
sure of CO2 in bulk gas phase.
ical properties of the gas and liquid phases, as well as the
The enhancement factor is defined as the ratio of the
solubility of gas and diffusivity in the liquid phase. The dif-
liquid-film mass transfer coefficient for chemical absorption,
fusivity of a solute gas in a liquid is strongly dependent on
kl , to the liquid-film mass transfer coefficient for physical
both temperature and viscosity of the liquid. The following
absorption, kol . One of the most important yet difficult aspects
semi-empirical correlations that relate CO2 diffusivity to both
of modelling reactive absorption processes is the determi-
of these parameters given by Versteeg and van Swaaij (1988)
nation of the enhancement factor (Tontiwachwuthikul et al.,
was used:
1992), which requires solutions of coupled diffusion-reaction
differential equations in the liquid film. As it stands, exact   0.8
w
analytical solutions do not exist for second-order kinetics. DCO2 ,l = DCO2 ,w (6a)
l
Approximate solutions have been provided by a number
researchers (e.g., van Krevelen and Hoftijzer, 1948; Wellek
 −2119 
DCO2 ,w = 2.35 × 10−6 exp (6b)
et al., 1978; DeCoursey, 1982). In this work, following previous T
studies (Tontiwachwuthikul et al., 1992; Pintola et al., 1993),
the explicit relations developed by Wellek et al. (1978) in terms where DCO2 ,l and DCO2 ,w are the diffusivity of CO2 in the solu-
of the enhancement factor for an instantaneous reaction, E∞ , tion and in water, l and w are the viscosity of the solution
and Hatta modulus, M, has be adopted: and water, and T is the solution temperature. The viscosity
ratio in Eq. (6a) was obtained from the following correlation
1 (Weiland et al., 1998) in terms of the weight percent of MEA in
E=1+ (4)
1.35 1.35 1/1.35 the solution, w, CO2 loading, ˛, and the solution temperature:
[(1/(E∞ − 1)) + (1/(E1 − 1)) ]

where 
l (21.186w + 2373)[˛(0.01015w + 0.0093T − 2.2589) + 1]w
 C D  √
M D k2 C w
= exp
T2
Bl B
E∞ = 1 + , E1 = √ and M2 = Al 2 Bl (7)
bDAl CAi tanh M (kol )

where CAi is the CO2 concentration at the gas–liquid interface,


CBl is the MEA concentration in bulk liquid, and DAl and DBl The Henry’s law constant was determined using the cor-
are the diffusivity of CO2 and MEA in liquid, respectively. The relation proposed by Danckwerts (1970) to calculate the
chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608 1603

solubility of CO2 in MEA solutions as a function of MEA con-


centration, CMEA , and solution temperature:

HCO2 = 10(5.3−0.035CMEA −1140/T) (8)

This approach was adopted instead of a rigorous method


based on the electrolyte NRTL model (Austgen et al., 1989) for
the vapour-liquid equilibrium of CO2 –MEA system for the ease
of computation.

2.4. Gas–liquid contactor model

A gas–liquid contactor model is developed using appropriate


forms of the species and energy conservation equations for the
predictions of CO2 and MEA concentrations and temperatures
in the gas and liquid phases as a function of packing height.
The model is based on the following assumptions:

• Steady-state and adiabatic operations.


• Plug-flow of gas and liquid. Fig. 2 – Schematic of a packed-bed CO2 absorption column
• Diffusion in the axial direction is negligible compared with with an infinitesimal element for mass and energy
convective transport. balances.
• The gas and liquid flow rates are constant throughout the
column (i.e., dilute gas–liquid system). Gas phase energy equation:
• Evaporation of solvent is negligible.
• Pressure is constant. dTg
GYCO2 Cpg = hg a(Tg − Tl ) (11)
dz
2.4.1. Mass balance Liquid phase energy equation:
The variation of CO2 concentration in the gas phase along the
height of the column is given by the following steady-state, dTl dTg dYCO2
LCpl = GYCO2 Cpg + G(HR + HS ) (12)
one-dimensional species conservation equation: dz dz dz
where Tg and Tl are the gas and the liquid temperature,
d(uz CCO2 ,g ) dNCO2 ,g respectively, hg is the gas phase heat transfer coefficient,
= = SCO2 (9)
dz dz and HR and HS are the heat of reaction and heat of solu-
tion, respectively. The values of HR and HS given by Akanksha
where CCO2 ,g is the molar concentration of CO2 in the gas
et al. (2007) were used in the calculation. Eqs. (11) and (12)
phase, NCO2 ,g (= uz CCO2 ,g = GYCO2 ) is the molar flux of CO2 , G is
were discretised using the forward difference scheme over
the molar flow rate of gas per unit cross-sectional area of the
the incremental height as shown in Fig. 2.
column, SCO2 is the source term representing the overall rate
of absorption of CO2 , uz is the gas velocity in the axial direc-
tion (along the column height), and YCO2 is the mole ratio of 2.5. Numerical solution
CO2 in the gas phase. The source term for a fast second-order
reaction is given by Eq. (3). Eq. (9) is discretised over an incre- The absorption column was divided into a number of infinites-
mental height of the column (z), as shown in Fig. 2, using a imal elements of height z and the discretised species
forward difference scheme. conservation and thermal energy equations were solved for
The MEA concentration in the solution can be obtained by each element using Microsoft Excel spreadsheet. For the sim-
performing a mass balance over the incremental height: ulation of counter-current absorption columns, the mass and
energy balance equations for the gas phase were solved
(YCO2 i − YCO2 i+1 )bG from the bottom to the top of the column, whilst the liq-
XMEAi = XMEAi+1 + (10) uid phase equations were solved from the top to the bottom
L
of the column. Calculations were performed using the given
where b is the stoichiometric factor of reaction (R1), XMEA is inlet concentration, temperature and flow rate of the gaseous
the mole ratio of MEA in water, and L is the molar flow rate of stream at the bottom and the liquid stream at the top of
water. the column (see Fig. 2). Although the compositions and tem-
peratures of the streams at the outlets were known for the
2.4.2. Energy balance experimental cases simulated in this study, they were not
The absorption of CO2 in a MEA solution results in the release specified as the exit conditions needed for the calculation.
of heat of solution followed by heat of reaction due to the Instead, these data were specified as initial guess and itera-
exothermic chemical reaction. Both of these effects increase tive calculations were performed until a converged solution
the liquid temperature, and consequently the gas temperature was achieved. For the estimation of the enhancement factor
due to it being in direct contact with the liquid, from the inlet using Eq. (4), an iterative procedure was adapted to determine
to the outlet of the column. The variations of the gas and the the unknown CO2 concentration at the gas–liquid interface in
liquid temperature are determined from the thermal energy each element of the absorption column following an approach
equations given below. similar to that used by Aroonwilas et al. (2003).
1604 chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608

Table 1 – Experimental conditions for the CO2 absorption using MEA solutions in counter-current packed-bed absorption
columns.
Scale size/packing type

Pilot-plant column with Pilot-plant with structured Industrial column with


random packinga packingb random packingc

T22d T13d T15d T19d I06d

Gas flow rate [mol/m2 s] 14.8 14.8 14.8 14.8 10.7 89.32
Inlet CO2 conc. [%mol] 19.1 15.3 19.5 11.5 15.0 1.53
Liquid flow rate [m3 /m2 h] 9.5 13.5 13.5 13.5 15.9 13.27
Inlet MEA conc. [kmol/m3 ] 3.00 2.00 2.03 2.00 5.20 4.10
Liquid temperature [◦ C] 19.0 19.0 19.0 19.0 19.7 38.7
CO2 removal efficiency [%] 100.0 100.0 100.0 100.0 – 99.92
Column pressure [kPa] 103.15 103.15 103.15 103.15 101.32 724.47

a
Tontiwachwuthikul et al. (1992).
b
Aroonwilas et al. (2001).
c
Pintola et al. (1993).
d
Run number.

3. Experimental cases data for run #T22 (Tontiwachwuthikul et al., 1992). Fig. 3(a)
shows the comparison between the predicted and measured
The rate-based process model was validated by simulat- gas phase CO2 concentration profiles along the column height.
ing a number of experiments on the absorption of CO2 in As can be seen, overall agreement between the prediction
aqueous MEA solutions carried out in counter-current packed- and measurement is generally good; however, the CO2 con-
bed absorption columns ranging from a small pilot-plant centration is somewhat underpredicted in the initial 2.5 m of
to a large industrial scale size. The following experimen- the column. This may be due to the use of empirical corre-
tal studies (Tontiwachwuthikul et al., 1992; Pintola et al., lations for the estimation of thermodynamic and transport
1993; Aroonwilas et al., 2001) reported in the literature were
selected. Aroonwilas et al. (2001) carried out experiments in 25.00 (a)
a small scale pilot-plant absorption column of 2.21 m packing
height and 0.1 m in internal diameter containing cylindrical 20.00
CO2 Concentration (mol%)

elements of stainless steel Gempak 4A structured packings


(element height = 0.245 m, specific surface area = 466 m2 /m3
15.00
and void fraction = 0.92). The gas mixture, air–CO2 , was intro-
duced into the column at the bottom and the solvent was
introduced at the top, thus providing a counter-current flow. 10.00
Having reached a steady-state, the gas phase CO2 concentra-
tions and solution temperatures were measured at different
5.00
locations along the column height. Tontiwachwuthikul et al.
(1992) carried out a series of experiments in a larger counter-
current pilot-plant absorption column consisting of six packed 0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
beds with a total packing height of 6.6 m and 0.1 m in internal Distance from the bottom of the column (m)
diameter. The column was packed with 12.7 mm ceramic Berl
saddles. Steady-state gas phase CO2 and MEA concentrations 3500.00 (b)
and solution temperatures were measured along the column
MEA Concentration (mol/m3)

height for different gas/liquid ratio, inlet CO2 and MEA concen- 3000.00

trations. Pintola et al. (1993) reported data taken from an acid


2500.00
gas treatment unit of an olefin plant. The absorption column,
26.6 m high and 1.9 m internal diameter, was packed with 2000.00
50 mm stainless steel Pall rings with a total packing height
of 14.1 m. The packing was divided into three sections with 1500.00
heights of 4.6, 4.6, and 4.9 m. The CO2 and MEA concentra-
1000.00
tions, temperatures, and flow rates were only measured at the
top and the bottom of the column. The operating conditions
500.00
of all the experimental runs simulated in this study are given
in Table 1. 0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
Distance from the bottom of the column (m)
4. Results and discussion
Fig. 3 – Comparison between the predicted and measured
4.1. Base case simulation (a) CO2 and (b) MEA concentration profiles along the column
height. (–) Prediction and () Expt. data of run #T22
For detail validation of the model described in Section 2, the (accuracy of CO2 data: ±2%).
predictions are compared with the pilot-plant experimental Tontiwachwuthikul et al. (1992).
chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608 1605

160.00 25.00

140.00

CO2 Concentration (mol%)


20.00
120.00
Enhancement Factor

100.00
15.00
80.00
10.00
60.00

40.00
5.00
20.00

0.00 0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
Distance from the bottom of the column (m) Distance from the bottom of the column (m)

Fig. 4 – The predicted enhancement factor as a function of Fig. 6 – Comparison between the predicted CO2
column height for run #T22. concentration profiles obtained using klo from Onda et al.
(1968) correlation (–) and penetration theory (- - -) for run
properties and parameters associated with mass transfer in #T22. () Expt. data (accuracy ± 2%).
the absorption rate Eq. (3). The error associated with CO2 con- Tontiwachwuthikul et al. (1992).
centration measurement might have also contributed towards
the discrepancy between the prediction and measurement.
The predicted and measured MEA concentration distribu-
of solvent evaporation in the energy balance. As shown in
tions along the column are shown in Fig. 3(b). The level of
Fig. 5, the heat effects associated with the CO2 absorption and
agreement between the predicted and measured CO2 con-
reaction with MEA cause a rapid variation of the solvent tem-
centrations revealed in Fig. 3(a) is also reflected here. The
perature in the bottom half of the column with a maximum
predicted outlet concentrations of both CO2 and MEA exactly
temperature at the exit. This is due to a significant amount of
match the experimental data. It is evident from these figures
CO2 absorption in this region as revealed in Fig. 3. This empha-
that the concentrations decrease rapidly in the bottom part of
sises that the variation of temperature due to the exothermic
the column, indicating that most CO2 removal from the gas
nature of the absorption and reaction within the absorption
takes place in that particular region. The remaining height of
column should not be neglected as the reaction rate as well as
the column serves to further reduce the CO2 concentration to
the gas solubility and transport properties is a strong function
a very small value. Fig. 4 shows the variation of the enhance-
of temperature.
ment factor along the height of the column, which increases
rapidly in the bottom half of the column.
Fig. 5 shows the comparison between the predicted and 4.2. Sensitivity study
measured solution temperature along the column height.
Although the predicted temperature profile is in good qual- The sensitivity of the predictions to the gas- and liquid-film
itative agreement with measurement, there are discrepancies mass transfer coefficients was assessed by simulating the
in the bottom section of the column where most CO2 absorp- experimental run #T22 (Tontiwachwuthikul et al., 1992). In the
tion occurs. The solvent temperature near the bottom exit is base case predictions shown in Fig. 3, the gas (kg ) and liquid
overpredicted in spite of the accurate prediction of the MEA (kol ) film mass transfer coefficients were obtained from Onda
concentration (see Fig. 3(b)). This may be due to the neglect et al.’s correlation (Eqs. (5a) and (5b)). The influence of kol on
the model predictions was assessed by estimating this param-
60.00 eter using the penetration theory of Higbie (1935), according
to which kol is given by:
50.00

Liquid Temperature (ºC)

DAl
40.00 kol =2 (13)
t

30.00
where t is the contact time. The calculated CO2 concentra-
tion profiles using kol obtained from the penetration theory
20.00
and from the correlation of Onda et al. (1968) are compared
in Fig. 6. As can be seen, the predicted concentrations in the
10.00
bottom half of the absorption column are somewhat differ-
ent, with the prediction obtained from Onda et al.’s correlation
0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 being in close agreement with measurements, while it dimin-
Distance from the bottom of the column (m) ishes in the upper region of the column. It is interesting to
note that the sensitivity of the predictions to kol suggests that
Fig. 5 – Comparison between the predicted and measured the diffusive transport of CO2 from the gas–liquid interface to
solution temperature profiles along the column height. (–) the reaction zone influences the overall absorption rate.
Prediction and () Expt. data of run #T22. A number of empirical correlations are reported in the
Tontiwachwuthikul et al. (1992). literature for the estimation of the gas-film mass transfer coef-
1606 chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608

25.00 18.00

CO2 Concentration (mol%)


16.00 Run #T13
14.00
CO2 Concentration (mol%)

20.00
12.00
10.00
15.00 8.00
6.00
4.00
10.00
2.00
0.00
5.00 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
Distance from the bottom of the column (m)

0.00 25.00

CO2 Concentration (mol%)


0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 Run #T15
Distance from the bottom of the column (m) 20.00

Fig. 7 – Comparison between the predicted CO2


15.00
concentration profiles obtained using kg from Onda et al.
(1968) (–) and Bravo et al. (1985) (- - -) correlations for run 10.00
#T22. () Expt. data (accuracy ± 2%).
Tontiwachwuthikul et al. (1992). 5.00

0.00
ficient. One such correlation proposed by Bravo et al. (1985) is
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
given by: Distance from the bottom of the column (m)

Shg = 0.0338Re0.8 0.333 14.00


g Scg (14)
CO2 Concentration (mol%)

12.00 Run #T19

where Reg is the Reynolds number, Scg is the Schmidt number,


10.00
and Shg is the Sherwood number. The predicted CO2 concen-
trations obtained using kg given by Eq. (14) and the correlation 8.00

(Eq. (5a)) of Onda et al. (1968) are shown in Fig. 7, which reveal 6.00
that the mass transfer rate in the gas phase is considerably 4.00
overestimated by the former correlation.
2.00
Cleary, the liquid-film mass transfer coefficient obtained
from the penetration theory and the gas-film coefficient given 0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00
by Bravo et al.’s correlation result in the CO2 concentration pre-
Distance from the bottom of the column (m)
dictions which deviate from the base case simulation results
obtained using the correlations of Onda et al. (1968) for both Fig. 8 – Comparison between the predicted and measured
parameters as well as the experimental data. A possible cause CO2 concentration profiles. (–) Prediction and () Expt. data
for this may lie in the fact that the film mass transfer coeffi- of run #T13, T15 and T19 (accuracy ± 2%).
cients in the expressions developed by Onda et al. (1968) are Tontiwachwuthikul et al. (1992).
correlated as functions of packing type (e.g., Berl saddles or
Raschig rings) as well as their dimensions. The packing type of process conditions if these models are to be used for the
affects the hydrodynamic conditions in the column, in addi- process design. The performance of the developed model was
tion to the gas–liquid interfacial area, and hence the mass examined by simulating three experimental runs (T13, T15
transfer rate between the two phases. In the penetration and T19) carried out by Tontiwachwuthikul et al. (1992) by
theory and the correlation of Bravo et al. (1985), the film coef- varying the inlet CO2 and MEA concentrations and the solution
ficients are not expressed as a function of packing geometry flow rate (i.e., the gas/liquid ratio) from those used in the base
and therefore not able to properly account for its effect on the case run #T22 (see Table 1). The predicted CO2 concentration
absorption rate. The penetration theory also neglects convec- profiles for these runs are compared with the experimental
tive mass transfer due to the liquid motion, which can also data in Fig. 8. As can be seen, for all three runs good levels
influence the absorption rate. of agreement with experimental data are achieved, with the
predicted trends similar to that observed for run #T22 in Fig. 3.
4.3. Simulation of the effects of process parameters
4.4. Simulation of an absorber with structured
In post-combustion CO2 capture plants, the process condi- packing
tions such as the flue gas and solution flow rates (and hence
the gas/liquid ratio) and inlet CO2 and solvent concentrations Structured packings are used to provide lower gas phase pres-
can vary depending on the type of fuel burnt and the load on sure drops than random packings while maintaining high
the boilers. Thus, the absorption columns must be designed mass transfer rates, which can be advantageous for the post-
and operated to handle the variations in the inlet conditions combustion CO2 capture applications. Although these types of
without adversely affecting the CO2 capture efficiency. It is packings are used in a number of applications, their industrial
therefore important that the absorption process models must use for CO2 absorption has not yet been reported in the open
demonstrate their ability to simulate accurately a wide range literature (Aroonwilas et al., 2003). In contrast with randomly
chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608 1607

16.00 1.80 (a)

CO2 Concentration (mol%)


14.00 1.60
CO2 Concentration (mol%)

1.40
12.00
1.20
10.00
1.00
8.00
0.80
6.00
0.60
4.00 0.40

2.00 0.20

0.00 0.00
0.00 0.50 1.00 1.50 2.00 2.50 0.00 2.00 4.00 6.00 8.00 10.00 12.00 14.00 16.00
Distance from the bottom of the column (m) Distance from the bottom of the column (m)

Fig. 9 – Comparison between the predicted and measured 4200.00 (b)


CO2 concentration profiles in an absorption column with

MEA Concentration (mol/m3)


structured packing. (–) Prediction and () Expt. data 4000.00
(accuracy ± 2%).
Aroonwilas et al. (2001). 3800.00

3600.00
packed absorption columns, modelling of gas absorption in
structured packings has received less attention (Aroonwilas
3400.00
et al., 2003). The present absorption model was used to sim-
ulate an experiment carried out by Aroonwilas et al. (2001)
3200.00
in a column of 2.21 m packing height and 0.1 m internal
diameter packed with stainless-steel Gempak 4A structured
3000.00
packing. The experimental conditions used in the simulation 0.00 2.00 4.00 6.00 8.00 10.00 12.00 14.00 16.00
are listed in Table 1. Fig. 9 shows the comparison between Distance from the bottom of the column (m)
the predicted and measured CO2 concentrations along the
absorption column. The predictions are in very good agree- Fig. 10 – The predicted (a) CO2 and (b) MEA concentration
ment with experimental data. It is interesting to note that profiles. (–) Prediction and measured exit concentrations ()
the level of agreement between the predictions and measure- in an industrial absorption column.
ments shown in Fig. 9 is similar to that obtained using a more Pintola et al. (1993).
rigorous absorption model incorporating a mechanistic model
of liquid distribution over the packings by Aroonwilas et al. 5. Conclusions
(2003).
A rate-based absorption model has been developed for the
4.5. Simulation of an industrial absorber simulation of post-combustion amine based CO2 capture
processes in packed-bed absorption columns. The model pre-
Finally, the model was applied to simulate an industrial dictions are validated against wide ranging experimental data
absorption column of an acid gas treatment unit in an olefin of gas phase CO2 and liquid phase MEA concentrations and the
plant for which data was reported by Pintola et al. (1993). The solution temperature reported in the literature. Good agree-
column packing consisted of three sections with heights of ment is obtained between the predictions and measurements
4.6, 4.6 and 4.9 m. The absorber operating conditions for the in pilot and industrial scale CO2 absorption columns for a
trial run (I06) simulated in this study are given in Table 1. As range of operating conditions. The predicted concentration
can be seen from the table, the gas and liquid flow rates are profiles along the column height, in agreement with the mea-
much larger and MEA concentration is higher (∼4 kmol/m3 ) in surement, show that most CO2 removal typically occurs in the
the industrial absorber than those used in the pilot-plant by bottom region of the absorber. The top region merely reduces
Tontiwachwuthikul et al. (1992). Fig. 10 shows the predicted the CO2 concentration down to a ppm level. The sensitivity
CO2 and MEA concentration profiles together with the mea- studies have revealed that amongst the various correlations
sured respective concentrations at the top and the bottom for the gas- and liquid-film mass transfer coefficients, the ones
exit of the column. Unfortunately, concentration data along given by Onda et al. (1968) provides better predictions. Further
the height of the column is not available for detail validation validation of the rate-based model using data from industrial
of the predictions. However, it is evident from the figure that scale columns is need before they can be used reliably for the
there is good agreement between the predicted and measured design and simulation of absorption columns for CO2 capture
exit concentrations. It should also be noted that the simulated from power plant flue gases.
concentration profiles are very similar to those obtained for
the pilot-plant runs. The predicted CO2 concentration profile
References
reveals that over 98% of CO2 is absorbed in the bottom two
packed beds of the column, which represents about 67.4% of
Akanksha, Pant, K.K., Srivastava, V.K., 2007. Carbon dioxide
the total packing height, while the top bed reduces the CO2 absorption into monoethanolamine in a continuous film
concentration to a ppm level. contactor. Chemical Engineering Journal 133, 229–237.
1608 chemical engineering research and design 8 9 ( 2 0 1 1 ) 1600–1608

Aroonwilas, A., Tontiwachwuthikul, P., Chakma, A., 2001. Effects Levenspiel, O., 1999. Chemical Reaction Engineering, 3rd ed. John
of operating and design parameters on CO2 absorption in Wiley & Sons.
columns with structured packings. Separation and Onda, K., Takeuchi, H., Okumoto, Y., 1968. Mass transfer
Purification Technology 24, 403–411. coefficients between gas and liquid phases in packed
Aroonwilas, A., Chakma, A., Tontiwachwuthikul, P., Veawab, A., column. Journal of Chemical Engineering of Japan 1 (1),
2003. Mathematical modelling of mass transfer and 56–61.
hydrodynamics in CO2 absorbers packed with structure Pandya, J.D., 1983. Adiabatic gas absorption and stripping with
packings. Chemical Engineering Science 58, 4039–4053. chemical reaction in packed towers. Chemical Engineering
Astarita, G., 1967. Mass Transfer with Chemical Reactions. Communications 19 (4), 343–361.
Elsevier, Amsterdam. Pintola, T., Tontiwachwuthikul, P., Meisen, A., 1993. Simulation of
Austgen, D.M., Rochelle, G.T., Peng, X., Chen, C., 1989. Model of pilot plant and industrial CO2 –MEA absorbers. Gas Separation
vapor–liquid equlibria for aqueous acid gas–alkanolamine & Purification 7 (1), 47–52.
systems using the electrolyte-NRTL equation. Industrial & Raynal, L., Royon-Lebeaud, A., 2007. A multi-scale approach for
Engineering Chemistry Research 28, 1060–1073. CFD calculations of gas–liquid flow within large size column
Bravo, J.L., Rocha, J.A., Fair, J.R., 1985. Mass transfer in gauze equipped with structured packing. Chemical Engineering
packings. Hydrocarbon Processing 64, 91–95. Science 62, 7196–7204.
Danckwerts, P.V., 1970. Gas–Liquid Reactions. McGraw-Hill, New Tobiesen, F.A., Svendsen, H.F., Juliussen, O., 2007. Experimental
York. validation of a rigorous absorber model for CO2
DeCoursey, W.J., 1982. Enhancement factors for gas absorption postcombustion capture. Journal of American Institute of
with reversible reaction. Chemical Engineering Science 37, Chemical Engineers 53 (4), 846–865.
1483–1489. Tontiwachwuthikul, P., Meisen, A., Lim, J., 1992. CO2 absorption
Doraiswamy, L.K., 2001. Organic Synthesis Engineering. Oxford by NaOH, monoethanolamine and
Press. 2-amino-2-methyl-1-propanol solutions in a packed column.
Freguia, S., Rochelle, G.T., 2003. Modeling of CO2 capture by Chemical Engineering Science 47 (2), 381–390.
aqueous monoethanolamine. Journal of American Institute of Treybal, E., 1969. Adiabatic gas absorption and stripping in
Chemical Engineers 49 (7), 1676–1686. packed towers. Industrial & Engineering Chemistry 61 (7),
Freund, P., 2003. Making deep reductions in CO2 emissions from 36–41.
coal-fired power plant using capture and storage of CO2 . van Krevelen, D.W., Hoftijzer, P.J., 1948. Kinetics of gas–liquid
Proceedings of Institute of Mechanical Engineers, Part A: J. reactions: part 1. General theory. Recueil des Travaux
Power and Energy 217, 1–7. Chimiques des Pays-Bas 67, 563–586.
Higbie, R., 1935. The rate of absorption of a pure gas into a still Versteeg, G.F., van Swaaij, W.P., 1988. Solubility and diffusivity of
liquid during short periods of exposure. Transactions of acid gases (CO2 , N2 O) in aqueous alkanolamine solutions.
American Institute of Chemical Engineers 31, 365–389. Journal of Chemical & Engineering Data 33, 29–34.
Hikita, H., Asal, S., Ishikawa, H., Honda, M., 1977. The kinetics of Weiland, R.H., Dingman, J.C., Cronin, D.B., Browning, G.J., 1998.
reactions of CO2 with MEA, DEA and TEA by a rapid mixing Density and viscosity of some partially carbonated aqueous
method. Chemical Engineering Journal 13, 7–12. alkanolamine solutions and their blends. Journal of Chemical
Idem, R., Tontiwachwuthikul, P., 2006. Preface for the special & Engineering Data 43, 378–382.
issue on the capture of carbon dioxide from industrial Wellek, R.M., Brunson, R.J., Law, F.H., 1978. Enhancement factors
sources: technological developments and future for gas absorption with second-order irreversible chemical
opportunities. Industrial & Engineering Chemistry reaction. Canadian Journal of Chemical Engineering 56,
Fundamentals 45 (8), 2413. 181–186.
Kenig, E.Y., Schneider, R., Gorak, A., 2001. Reactive absorption: Whitman, W.G., 1923. The two-film theory of gas absorption.
optimal process design via optimal modelling. Chemical Chemical and Metallurgical Engineering 29, 146–148.
Engineering Science 56 (2), 343–350. Wilson, E.J., Gerard, D., 2007. Carbon Capture and Sequestration:
Lawal, A., Wang, M., Stephenson, P., Yeung, H., 2009. Dynamic Integrating Technology, Monitoring and Regulation. Blackwell
modelling of CO2 absorption for post combustion capture in Publications, Oxford.
coal-fired power plants. Fuel 88, 2455–2462. WRI, 2007. Climate Analysis Indicators Tool., http://www.wri.org/.

Vous aimerez peut-être aussi