You are on page 1of 10

Ind. Eng. Chem. Res.

2006, 45, 6339-6348

6339

Neutralization of the Acidified Seawater Effluent from the Flue Gas


Desulfurization Process: Experimental Investigation, Dynamic Modeling, and
Simulation
Masahiro Tokumura, Mayumi Baba, Hussein Tawfeek Znad, and Yoshinori Kawase*
Research Center for Biochemical and EnVironmental Engineering, Department of Applied Chemistry,
Toyo UniVersity, Kawagoe, Saitama 350-8585, Japan

Chaturong Yongsiri and Kazuo Takeda


Fujikasui Engineering Co., Ltd., 1-4-3 Higashi-Gotanda, Shinagawa, Tokyo 141-0022, Japan

The neutralization of the seawater effluent in a wet flue gas desulfurization (FGD) process was investigated.
Before discharge, the acidified seawater effluent from a absorber is neutralized, using the natural alkalinity
in seawater, and aeration is performed to enhance the neutralization due to the stripping of carbon dioxide
dissolved in the effluent. Experimental investigation, dynamic modeling, and simulation have been performed
to elucidate the neutralization of the seawater effluent in a FGD process. We have measured changes in
solution pH and the dissolved carbon dioxide concentration and carbon dioxide concentration in the outlet
gas, relative to the time of aeration. Increases in the aeration rate, salinity, dissolved SO2 concentration, and
temperature increased the neutralization rate. We have developed a rate-based model to simulate the
neutralization process of the acidified seawater effluent. In the modeling, nonideal mixing in the gas phase
is described using a tanks-in-series model while the liquid is assumed to be completely mixed. The model,
which is based on reaction kinetics and equilibrium in the liquid phase and gas-liquid mass transfer, combined
with the nonideal mixing in the gas phase, consists of five ordinary differential equations. In simulating the
neutralization of the acidified seawater effluent, the effects of its temperature and salinity on the neutralization
rate were taken into account. The proposed model, which included no adjustable parameters, could describe
the present experimental results for the dynamic changes of solution pH, dissolved CO2 concentration, and
CO2 concentration in the gas phase successfully.
Introduction
Sulfur in fossil fuels used for power generation is converted
to sulfur dioxide (SO2) during combustion. SO2 is a precursor
to acid rain, which causes accelerated soil acidification and forest
degradation. To minimize the adverse impacts of SO2 emissions
on the environment, much effort has been put into developing
flue gas desulfurization (FGD) technologies for the control and
abatement of SO2 emissions.1,2 Various types of FGD processes
are currently available, and they are classified as wet, semidry, and dry processes. The installed worldwide FGD capacity
is more than 300 GW.3 Wet FGD processes have a tendency to
utilize sorbent more efficiently than dry processes and typically
can reduce SO2 emissions by more than 90%, up to 98%.4,5
The low cost of reactants (e.g., limestone) and, in most cases,
a commercial product (gypsum) are also listed among their
advantages. Therefore, they are the most widespread and have
been studied extensively.6,7 Attempts to establish a model for
the wet FGD processes also have been made.8-11
A new practical development in FGD processes for power
plants located near the coast is the utilization of seawater from
the cooling system of the plant to scrub SO2 from flue gas.12
The seawater FGD processes are categorized as pure seawater
or alkali seawater processes. Figure 1a depicts the schematic
diagram of a pure seawater FGD process. The chemical reactions
that occurred in the seawater FGD process are illustrated in
Figure 1b. Flue gas is introduced to an absorber and contacted
with seawater. The SO2 in flue gas reacts with seawater in the
* To whom correspondence should be addressed. Tel.: +81-49-2391377. Fax: +81-49-231-1031. E-mail: bckawase@mail.eng.toyo.ac.jp.

absorber and forms bisulfite (HSO3-) (via reaction 1). HSO3or SO32- (HSO3- T SO32- + H+) is oxidized to sulfate (SO42-),
because of oxygen in the flue gas and seawater (see reaction
2). Because the amount of oxygen in the flue gas and seawater
is not sufficient to oxidize completely to SO42-, air is blown
into the sump of the absorber. Via the chemical reactions
(reactions 1 and 2), H+ ions are produced and, as a result, the
seawater effluent becomes acidic. Therefore, the acidified
seawater effluent must undergo neutralization before discharge.
Seawater has a typical pH of 7.6-8.4, with an alkalinity of
100-110 mg/L (as CaCO3). The seawater from the cooling
system of the power plant is added to neutralize the acidified
seawater effluent from the scrubbing flue gas in a separate tank
(reactions 3 and 4). The air injection has the following important
assignments: oxidation of HSO3- and SO32- to SO42-, replenishment of dissolved oxygen in the seawater effluent, and
stripping of carbon dioxide. It can be understood from reactions
3 and 4 that the stripping of carbon dioxide highly improves
the neutralization of the acidified seawater effluent. The seawater
FGD process has already been performed, and its installed
worldwide capacity is 20 GW. However, there is only a small
amount of research in regard to neutralization via the stripping
of carbon dioxide.12 The low pH of the acidified seawater
effluent from the sump of the absorber (pH 3-4) necessitates
the neutralization process. By controlling the amounts of
seawater and aeration rate, the pH of the seawater effluent must
be adjusted to near 7 at the outlet of the neutralization tank. It
is necessary to have an understanding of the kinetic phenomena
in removal of carbon dioxide by air stripping process. The
removal of CO2 from natural, refinery, and synthesis gas streams

10.1021/ie0603619 CCC: $33.50 2006 American Chemical Society


Published on Web 07/28/2006

6340

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006

Figure 1. Schematic diagram of a pure seawater flue gas desulfurization (FGD) plant: (a) pure seawater FGD process, and (b) reactions in a pure seawater
FGD process.

is a significant operation, and there have been many studies of


reactive absorption of CO2 in chemical industries (e.g., Mandel
and Bandyopadhyay,13 Kongto et al.14). Furthermore, carbon
dioxide stripping has been also encountered in aquaculture
industries (e.g., Stumm and Morgan,15 Watten et al.16). However,
little information is available on the mechanism for the
neutralization of acidified seawater effluent in the FGD. In
particular, very few studies on the modeling of the neutralization
process by stripping of carbon dioxide have been reported.
In this paper, we have experimentally examined the effects
of operating parameters such as aeration rate, salinity, SO42concentration, and temperature on the neutralization of the

acidified seawater effluent. A simulation model for its process


dynamics has been developed, and its predictions have been
compared with the present experimental data.
Experimental Section
All experiments were performed in a cylindrical bubble
column reactor, the dimensions of which are 0.098 m in diameter
and 1.52 m in height. The working volume of the bubble column
reactor was 5 L. The operation was batch and continuous type,
with respect to the liquid phase and gas phase, respectively.
Air was injected through the porous glass filter sparger. The

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006 6341

volumetric inlet flow rate of air (Qg) was varied from 1 L/min
to 4 L/min and measured with a rotameter. Seawater, tap water,
and distilled water were used as a liquid phase. To examine
the effects of salinity, stripping of CO2 from tap water and
distilled water was performed. Sulfuric acid was added to change
the SO2 concentration in the model effluent. Before the start of
the measurement, 97% CO2 gas was injected to dissolve CO2
gas into the solution and then adjust the initial pH of the solution
to 4.5. Liquid samples were withdrawn intermittently from
the middle of the bubble column reactor for the determination
of the dissolved inorganic carbon (DIC) concentration. The
combustion-infrared total organic carbon (TOC) meter (TOCVE, Shimadzu Co.) was used to measure the DIC amount in
the solutions. The pH and temperature of the solution were
measured using a pH meter (HM-20P, Toa Denpa Co.), which
was immersed near the middle of the reactor. The outlet
concentration of carbon dioxide in the gas phase was measured
using a CO2 meter (Testo 535, Testo Co.). Because the CO2
concentration in the outlet gas phase for t < 5 min was rather
high, its CO2 concentration was measured after it was mixed
and diluted with air of measured flow rates.
Gas holdup (g) values were determined by the volume
expansion method. The volumetric mass-transfer coefficients
for CO2 (KLaL) were estimated from transient dissolved-oxygen
tension measurements, using a fast response dissolved oxygen
electrode (YSI model 57, Yellow Spring Instrument Co.).17
Of course, replicated data were taken to ensure the reliability
of the present experimental results. All experiments were
performed in triplicate.
Results and Discussion
Kinetics of Hydration of CO2 in Seawater. In seawater,
carbon dioxide exists in four different inorganic forms: as free
carbon dioxide (CO2), bicarbonate ion (HCO3-), carbonate ion
(CO32-), and true carbonic acid (H2CO3).
The concentration of CO2 dissolved in seawater is given by
Henrys law with Henrys constant (He) being the solubility
coefficient of CO2 in seawater:

[CO2] ) HeYCO2

(5)

where YCO2 is the partial pressure of CO2 in the gas phase.


The hydration of CO2 leads to the formation of H2CO3, but
it may also yield hydrogen ions (H+) and bicarbonate ions
(HCO3-). Therefore, the carbonate species are related by the
following equilibria:18
kf1

} HCO3- + H+
CO2 + H2O {\
k

(6)

b1

kf2

HCO3- {\
} CO32- + H+
k

(7)

b2

At low pH, the hydration of CO2 by reaction 6 is predominant.


Reaction 7 is considered to be instantaneous.13 The corresponding equilibrium relationships are defined as

K1 )
K2 )

[HCO3-][H+]
[CO2]
[CO32-][H+]
[HCO3-]

(8)

(9)

where K1 and K2 are equilibrium constants (K1 ) kf1/kb1 and

K2 ) kf2/kb2). At high pH value, the increase in the number of OH- ions enhances the following hydroxylation reaction:12
kf4

CO2 + OH- {\
} HCO3k

(10)

b4

This reaction indicates that a reaction of CO2 with a hydroxyl


ion gives a bicarbonate ion. The stoichiometric equilibrium
constants may be dependent on temperature, pressure, and
salinity. The concentrations of CO2, bicarbonate ion, and
carbonate ion are dependent on pH, and, consequently, the
predominant carbonate species in CO2 dissolved in seawater
changes relative to the solution pH.
In addition to the equilibrium relationships, a rate model has
been developed to predict the flux of CO2 from the seawater to
air bubbles. The following equation obtained from reactions 6
and 7 can describe the change in the H+ ion concentration in
seawater:

d[H+]
) kf1[CO2] - kb1[H+][HCO3-] +
dt
kf2[HCO3-] - kb2[CO32-][H+] (11)
By considering reactions 6, 7, and 10, we have the following equations for the material balances of [HCO3-] and
[CO32-]:

d[HCO3-]
) kf1[CO2] + kb2[CO32-][H+] +
dt
kf4[CO2][OH-] - kb1[H+][HCO3-] kf2[HCO3-] - kb4[HCO3-] (12)
and

d[CO32-]
) kf2[HCO3-] - kb2[CO32-][H+]
dt

(13)

respectively. The sum of the dissolved forms for the carbonate


systemsCO2, HCO3-, and CO32-sis called the total dissolved
inorganic carbon (DIC), which was experimentally measured
using the TOC meter.

DIC ) [CO2] + [HCO3-] + [CO32-]

(14)

The neutralization process involves mass transfer with


chemical reactions and equilibriums. Therefore, the next step
accounts for the mass transfer of carbon dioxide from the
liquid phase to the gas phase. Using a mathematical model that
is based on a tanks-in-series model, we have simulated the
dynamic behaviors of a neutralization process in a semibatch
bubble column reactor. The tanks-in-series model, in which
the flow in the reactor is considered as flow through a series
of equally sized, well-mixed stirred tanks, has been applied
to represent the performance of reactors under a wide range
of operating conditions (Figure 2).19,20 The number of hypothetical tanks indicates the degree of mixing in the reactor.
The liquid-phase mixing was assumed to be perfect mixing
and treated as a single, well-mixed tank. The gas phase was
described as a tanks-in-series model. Incidentally, the plugflow can be approximated with 20 hypothetical well-mixed
tanks.20
By taking the mass balance of dissolved carbon dioxide for
a completely mixed liquid phase in a semibatch bubble column

6342

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006

Figure 3. Gas holdup (g) and volumetric mass-transfer coefficient (KLaL)


for oxygen at T ) 22 C.

Figure 2. Schematic diagram of the tanks-in-series model for a semibatch


neutralization tank reactor.

reactor, the rate of change in dissolved carbon dioxide may be


written as

d[CO2]
) KLaL(C*CO2 - [CO2]) +
dt
(kb1[H+] + kb4)[HCO3-] - (kf1 + kf4[OH-])[CO2] (15)
In the case of CO2 stripping, the first, second, and third terms
on the right-hand side of eq 15 represent the mass transfer at
the gas/liquid interface, the increment of dissolved carbon
dioxide due to reactions in the liquid phase, and the CO2
reduction due to reactions in the liquid phase, respectively. The
gas-liquid mass transfer of CO2 may be controlled by the
liquid-phase mass transfer due to the relatively low solubilities.9 In the proposed model, in which the gas phase is described as the tanks-in-series, the equilibrium dissolved carbon
dioxide concentration at the gas/liquid interface, C*CO2, is
represented as19

C*CO2 ) He

YCO ,j
N j)1
2

(16)

where N is the number of hypothetical well-mixed tanks for


the gas phase, representing nonideal gas-phase mixing.
The increase in the gas-phase carbon dioxide concentration
for the jth tank, Yj, under continuous gas-flow conditions, can
be represented by the following equation:19

dYCO2,j
dt

( )( )
( )

1 - g N
Q (Y
- YCO2,j) g
VL g CO2,j-1
1 - g
(KLaL)CO2(C*
CO2 - [CO2])RT (17)
g

where VL is the liquid volume in the bubble column reactor.


The first and second terms on the right-hand side of eq 17
account for the gas flows (Qg) from the (j-1)th tank and to the
(j+1)th tank and the mass transfer at the gas/liquid interface,
respectively.

The gas holdups, g, were measured in seawater and tap water.
The gas holdup increased as the superficial gas velocity (Ug)
increased. As shown in Figure 3, gas holdups in seawater were
slightly higher than those in tap water. We obtained the
following empirical correlation for g in seawater:

g ) 13.5Ug1.15

(18)

The liquid-side volumetric mass-transfer coefficients, KLaL,


were measured for oxygen transfer from air to seawater or tap
water. The value of KLaL for oxygen increased as Ug increased.
From the experimental data presented in Figure 3, we obtained
the following correlation for CO2 mass transfer, in which the
volumetric mass-transfer coefficient was corrected for diffusivities, according to Higbies penetration theory:

(KLaL)CO2 )

( )
DCO2
DO2

0.5

(KLaL)O2 )

( )
DCO2
DO2

0.5

(0.190Ug0.44)
(19)

The diffusivity values for CO2 and O2 (DCO2 and DO2, respectively) were taken from the literature.21,22 It was assumed that
their ratio (DCO2/DO2) in seawater is same as that in water, as
well as in the study of Yoon et al.23 Figure 3 shows that the
KLaL values in seawater are 1.5 times greater than those in
tap water. This may be attributed to the increase of interfacial
area in electrolyte solutions and is consistent with the results
for water and NaCl solutions in the literature (Deckwer24).
By numerically solving the aforementioned five simultaneous
ordinary differential equations (eqs 11, 12, 13, 15, and 17), we
can estimate the amount of DIC (eq 14), the solution pH (which
is equal to -log [H+]), and the concentration distribution of
carbon dioxide in the gas phase (YCO2,j), as a function of the
aeration time, and compare the model predictions with the
measured values, as shown below. A fourth-order Runge-Kutta
technique was applied to solve the differential equations.
Experimental Results. (a) Effect of Aeration Rate. The
stripping of CO2 in tap water was conducted at a gas flow rate
of 2.35-9.41 10-3 m/s. Figure 4 depicts the neutralization
process by the stripping of CO2 from tap water as a function of
the aeration time. The dissolved CO2 concentration decreased
rather sharply and, as a result, the solution pH increased to the
natural pH of tap water with increased aeration time. As
mentioned previously, the stripping of CO2 by aeration shifted
the equilibriums represented by reactions 3 and 4 to the right

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006 6343

Figure 4. Effect of aeration rate on the neutralization of acidified seawater


effluent (tap water, T ) 22 C), relative to (a) dissolved inorganic carbon
(DIC) and (b) pH.

[H+]

and consequently decreased the value of


and then increased the solution pH. The initial solution pH was adjusted
to pH 5 by sparging of the CO2 gas. The dissolved CO2, given
as the DIC value, was stripped from the acidified seawater
effluent with time. The pH of the acidified seawater effluent
for Ug g 0.00705 m/s attained the natural pH value of 7.8 after
20 min, which implies that the acidified seawater effluent was
neutralized completely by an aeration of 20 min. At Ug )
0.00235 m/s, the decrease in the DIC value and the increase in
pH were somewhat slow and the neutralization was completed
after 40 min.
The changes in the total dissolved inorganic carbon (DIC)
or dissolved CO2 and pH at the initial stage of the neutralization
process could be described by the pseudo-first-order kinetics,
with respect to DIC and pH, respectively:

d(DIC)
) kDIC(DIC)
dt

(20)

d(pH)
) kpH(pH)
dt

(21)

Figure 5. Effect of salinity on the neutralization of acidified seawater


effluent (Ug ) 0.00705 m/s, T ) 22 C, no addition of H2SO4), relative to
(a) DIC and (b) pH.

where kDIC and kpH are pseudo-first-order kinetic constants for


changes in the DIC and pH values, respectively, at the initial
stage of the neutralization process. The pseudo-first-order rate
constants were determined from the semilogarithmic plots of
the DIC or pH values, as a function of the aeration time. Plots
of ln((DIC)/(DIC)0) or ln((pH)/(pH)0) versus time resulted in
straight lines, the slopes of which are pseudo-first-order kinetic
constants.
The figures for kDIC and kpH given in Figure 4 clearly indicate
that the neutralization was enhanced by the increase in gas flow
rate. The increase in gas flow rate caused an increase in the
gas-liquid mass transfer, as described below, and then improved
the stripping of CO2.
(b) Effect of Salinity. Figure 5 presents the experimental
data on the neutralization of the acidified seawater effluent using
three different solventssseawater, tap water, and distilled
waterswhich have salinity values of 35, 5, and 0 g/(kg-solution),
respectively. The major difference between distilled water and
seawater is the total concentration and the relative proportion

6344

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006

Table 1. Correlations Used in the Simulation of the Neutralization Process of the Acidified Seawater Effluent in the Literature18
expression

source

(1) Rate Constants (kf1, kb1, kf2, kb2, kf4, and kb4) and Equilibrium Constants (K1, K2, KW)
kf1 (s-1) ) exp(1246.98 - 61900/T - 183 ln T)
kb1 (kg mol-1 s-1) ) kf1/K1
kb2 ) 5 1010
kf4 (kg mol-1 s-1) ) A exp[-E/(RT)], where A ) 4.7 107 kg/mol and E ) 23.2 kJ/mol
kb4 (s-1) ) kf4 Kw/K1
K1 (mol/kg) ) 10-(3670.7/T-62.008+9.7944 ln T-0.0118S+0.000116S2)
K2 (mol/kg) ) 10-(1394.7/T+4.777-0.0118S+0.000118S2)
K*H2CO3 (mol/kg) ) 10-3.8
KW (mol2/kg2) ) [H+][OH-] ) exp(148.9802 - 13847.26/T - 23.6521 ln T +
(118.67/T - 5.977 + 1.0495 ln T)S1/2 - 0.01615S)
S ) 35 + [SO42-]add, where S is defined as the weight in grams of the dissolved
inorganic matter in 1 kg of solution
(2) Henrys Constant, He
He (mol kg-1 atm-1) ) exp(9345.17/T - 60.2409 + 23.3585 ln(T/100) + S(0.023517 0.00023656T + 0.047036 (T/100)2

Johnson25
Eigen26
Johnson25
Mehrbach et al.27
Mehrbach et al.27
Stumm and Morgan15
Millero28

Weiss29

(3) Density of Seawater (Used To Convert Equilibrium Constants, K1 and K2 (Given in Units of mol/L))
Fsw (kg/m3) ) Fpw + AS + BS3/2 + CS2, where
Millero and Poisson30
Fpw (kg/m3) ) 999.842594+6.793952 10-2(T - 273.15) - 9.9095290 10-3(T - 273.15)2 +
1.001685 10-4(T - 273.15)3 - 1.120083 10-6(T - 273.15)4 + 6.536332 10-9(T - 273.15)5
A (kg/m3) ) 8.24493 10-1 - 4.0899 10-3(T - 273.15) + 7.6438 10-5(T - 273.15)2 8.2467 10-7(T - 273.15)3 + 5.3875 10-9(T - 273.15)4
B ) -5.72466 10-3 + 1.0227 10-4(T - 273.15) - 1.6546 10-6(T - 273.15)2
C ) 4.8314 10-4

of ions dissolved in the solution. As shown in Figure 5, the


natural pH values for seawater and tap water were achieved
after 10 and 15 min, respectively. The equilibrium or natural
pH in distilled water was 6.5 and rather low, compared to
other solvents. The seawater with the highest salinity showed a
faster neutralization rate than other solvents. The correlations
in the literature given in Table 1 predict that the higher
concentration of salinity increases the equilibrium constant K1
and decreases the concentration of CO2, because of depression of the backward reaction rate constant, kb1. On the other
hand, the higher concentration of salinity decreases Henrys
constant, He, because of the salting-out effect.31 This effect
might overcome the increase in K1 and as a result cause the
increase in the dissolved CO2 concentration. Note that the S
term in the correlation for He presented in Table 1 is negative in the temperature range used in this study. As a result,
from eq 5, we can expect that the partial pressure of CO2 in the
gas phase, YCO2, increases as the salinity increases. The decrease
in He causes enhancement of the stripping of CO2, via the
increase in the salinity concentration. In fact, as can be observed
from the figures for kDIC and kpH presented in Figure 5, the rates
of CO2 stripping and pH change increased as the salinity
increased.
(c) Effect of Changes in Dissolved SO2. The concentration
of SO2 in seawater supplied to the absorber is increased by
scrubbing the flue gas from fossil fuel-fired power plants. The
concentration of SO2 in the acidified seawater effluent is
dependent on the design and operating parameters of the fossil
fuel-fired power plants, such as the fuels combusted, the
reactants used, and the liquid-gas ratio. Therefore, we examined
the effect of dissolved SO2 concentration on the neutralization
of the acidified seawater effluent by adding H2SO4, in amounts
of 0, 0.3928, and 0.7855 mM, to seawater. These values were
determined based on the sulfuric acid concentrations in the
seawater effluent from the absorber in the actual seawater FGD
process.12 The increase in the SO42- concentration of the
seawater due to absorption of SO2 gas in the absorber is only
2%-4% of the SO42- concentration of 28 mM in standard
seawater.12,18 The effect of sulfuric acid concentration on the
neutralization of the acidified seawater effluent is shown in
Figure 6. The figures for kDIC and kpH presented in this figure

Figure 6. Effect of the addition of H2SO4 on the neutralization of acidified


seawater effluent (seawater, Ug ) 0.00705 m/s, T ) 22 C), relative to (a)
DIC and (b) pH.

indicate that the addition of sulfuric acid to seawater that has


high salinity had a slight influence on the neutralization rate.
As mentioned previously, seawater itself includes a high concen-

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006 6345

Figure 7. Effect of temperature on the neutralization of acidified seawater


effluent (seawater, Ug ) 0.00705 m/s, addition of H2SO4 ) 0.7855 mM),
relative to (a) DIC and (b) pH.

tration of sulfur dioxide. In comparison with this concentration,


the added concentration of sulfur dioxide is too low to have a
significant effect on the neutralization. Therefore, additional
SO42- had marginally enhanced the neutralization rate or
removal of CO2 and the increase in pH. As depicted in Figure
6b, the equilibrium pH values obtained after 15 min of aeration
decreased as the amount of H2SO4 added to the seawater
increased.
(d) Effect of Temperature. Figure 7 presents the experimental data on the neutralization of the acidified seawater
effluent using three different temperatures of 31, 22, and 12 C
(( 2 C), respectively. The highest temperature showed the
fastest neutralization in this experimental temperature range. By
considering the effect of temperature on the equilibrium constant
and Henrys constant He, using the correlations in the literature
summarized in Table 1, it can be expected that the increase in
temperature increases K1 and decreases He. The effect of the
increase in He may be superior to that of the decrease in K1
and, consequently, leads to a faster rate of dissolved CO2
removal or neutralization of the acidified seawater effluent with
increasing temperature. As shown in the figures for kDIC and

kpH presented in Figure 7, this coincides with the experimental


results.
Simulation Results. The model was validated by comparing
its predictions with the present experimental data for the
stripping of CO2 from seawater, tap water, and distilled water.
As described previously, the model equations developed on the
kinetics of CO2 stripping from seawater and material balances
for species in the carbonate system contain several equilibrium
and kinetic constants and physicochemical properties. The
correlations for the equilibrium constants and reaction rate
constants used in the simulation are given in Table 1. It should
be emphasized that the proposed model includes no parameters
that have been determined by adjusting the proposed model to
the present experimental data. In the modeling of a wet FGD
process, for example, incidentally, Kiil et al.9 determined the
value of the overall volumetric mass-transfer coefficient of SO2
by adjusting their model to the experimental outlet concentration
of SO2. Mandel et al.31 used the reaction rate constant as an
adjustable parameter in the model for the removal of CO2 by
absorption in mixed amines. We assumed that the mixing in
the gas phase is almost a plug flow and then N ) 10 in the
simulation, because, in small-scale reactors, the mixing in the
gas phase is usually assumed to be in plug flow.10
Figure 8 illustrates typical comparisons of the simulation
results and the experimental data for the DIC, pH, and YCO2
values in the seawater effluent, tap water, and distilled water.
The lines in the figures illustrate the simulation results. Figure
8 indicates that the model, in which no adjustable parameters
are included, predicts the experimental data, including the results
in tap water (Ug ) 0.0047 m/s, no addition of H2SO4) and
distilled water (Ug ) 0.00705 m/s, addition of H2SO4 ) 0.7855
mM), very successfully. As can be seen in Figure 8a, the model
could reasonably predict the sharp decrease in the dissolved
CO2 concentration caused by the stripping of CO2. Figure 8b
also indicates that the experimental results for the neutralization
process or the dynamic increase in pH to the natural solution
pH in seawater, tap water, and distilled water were satisfactorily
simulated by the proposed model. For distilled water that
contained 0.7885 mM of H2SO4, aeration resulted in a decrease
in the dissolved CO2 concentration but it could not increase
the pH. The pH remained almost constant, because the influence
of sulfuric acid was more significant, in comparison to the
stripping of CO2. This result could also be predicted by the
model successfully. As can be observed from the result for YCO2
in the seawater system (Figure 8c), the CO2 concentration in
the gas phase drastically increased just after the aeration and
then gradually decreased. The flux of CO2 from the liquid phase
decreased as the aeration time increased, because of the decrease in the driving force for mass transfer at the interface. The outlet YCO2 value was expected to be larger than the
YCO2 value in the middle (j ) 5) and at the bottom (j ) 1),
because of the accumulation of CO2 transferred into the air
injected from the reactor bottom. It can be observed that the
outlet CO2 concentration in the gas phase (j ) 10) achieved
steady-state conditions after 8 min and the model could
reasonably predict the experimental results for the change in
the outlet YCO2 value.
Computed concentration distributions of CO2(aq), true carbonic acid (H2CO3), bicarbonate (HCO3-) and carbonate
(CO32-) ions, and the DIC value in seawater during the
neutralization process at Ug ) 0.00705 m/s and T ) 22 C are
illustrated in Figure 9.
The concentration of true carbonic acid (H2CO3) was
calculated using the following relationship:18

6346

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006

Figure 9. Computed concentration distributions of CO2(aq), true carbonic


acid (H2CO3), bicarbonate (HCO3-), and carbonate ion (CO32-), DIC, and
pH in seawater during the neutralization process. Calculations were based
on pH0 ) 4.25, [CO2]0 ) 0.026 M, Ug ) 0.00705 m/s, T ) 22 C, addition
of H2SO4 ) 0.7885 mM.

K*H2CO3 )

[H+][HCO3-]
[H2CO3]

(22)

As the solution pH increased, the amount of CO2(aq) sharply


decreased, whereas that of the CO32- ion significantly increased
and that of the HCO3- ion decreased very slightly. Although
the amount of H2CO3 decreased as the pH increased, its amount
is negligible in comparison with the dominating species, as well
as that of the CO32- ion in the neutralization process. When
the pH is reduced to <5, CO2(aq) is the dominating dissolved
carbon dioxide species. For 5 < pH < 6, CO2(aq) and HCO3are comparable among the aqueous carbonate species. In the
region of pH >6, HCO3- becomes the dominating dissolved
carbon dioxide species.
Conclusions

Figure 8. Comparison of experimental results with model predictions for


the neutralization of acidified seawater effluent (seawater and distilled water;
Ug ) 0.00705 m/s, addition of H2SO4 ) 0.7885 mM: tap water; Ug )
0.0047 m/s, no addition of H2SO4): (a) total dissolved inorganic carbon
(DIC) or dissolved CO2, (b) solution pH, and (c) distributions of carbon
dioxide concentration in the gas phase (YCO2) in seawater at T ) 22 C (N
) 10: top (j ) 10), middle (j ) 5), and bottom (j ) 1) of reactor).

We measured changes in the concentration of dissolved


inorganic carbon (DIC) or dissolved CO2 concentration, solution
pH, and the concentration of outlet carbon dioxide in the gas
phase, relative to the time of aeration in a bubble column reactor,
during the neutralization of an acidified seawater effluent. It
was experimentally determined that the aeration rate, salinity,
addition of dissolved SO2, and temperature enhanced the
neutralization rate. We developed the rate-based model, taking
into account the temperature and salinity of the effluent, to
simulate the neutralization of the acidified seawater effluent.
The model could satisfactorily describe the dynamic changes
in pH, the concentration of DIC, and the outlet concentration
of carbon dioxide in the gas phase. It should be emphasized
that the dynamic model includes no adjustable parameters that
are determined so that the simulations match the experimental
data. It may be concluded, therefore, that the proposed model,
which is based on the chemical equilibrium of the chemical
reactions, the gas-liquid mass-transfer rate, and the nonideal
mixing, is very useful for the design and scaleup of the

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006 6347

neutralization process or removal of CO2 by air stripping for


the acidified seawater effluent in the flue gas desulfurization (FGD) process. The applicability of the proposed model
to full-scale conditions should be verified using the data in
full-scale plants, and this will be one of our future research
subjects.
Nomenclature
A, B, C ) coefficients in the correlation for density of seawater
C*
CO2 ) dissolved carbon dioxide concentration at the gas/
liquid interface (M)
DCO2 ) diffusion coefficient of CO2 (m2/s)
DO2 ) diffusion coefficient of O2 (m2/s)
DIC ) dissolved inorganic carbon (M)
He ) Henrys constant (M/Pa)
K1 ) first dissociation constant of carbonic acid, defined by eq
8 (M)
K2 ) second dissociation constant of carbonic acid, defined by
eq 9 (M)
K*H2CO3 ) dissociation constant of true carbonic acid, defined
by eq 22 (M)
KW ) dissociation constant, defined in Table 1 (M2)
KLaL ) liquid-side volumetric mass-transfer coefficient (s-1)
kb1 ) backward rate constant of carbonate system in seawater,
used in reaction 6 (M-1 s-1)
kb2 ) backward rate constant of carbonate system in seawater,
used in reaction 7 (s-1)
kb4 ) backward rate constant of carbonate system in seawater,
used in reaction 10 (s-1)
kDIC ) pseudo-first-order kinetic constants, defined by eq 20
(s-1)
kpH ) pseudo-first-order kinetic constants, defined by eq
21 (s-1)
kf1 ) forward rate constant of carbonate system in seawater,
used in reaction 6 (s-1)
kf2 ) forward rate constant of carbonate system in seawater,
used in reaction 7 (s-1)
kf4 ) forward rate constant of carbonate system in seawater,
used in reaction 10 (M-1 s-1)
N ) the number of tanks for the gas phase
P ) pressure (Pa)
Qg ) volumetric gas flow rate (m3/s)
R ) gas constant (J mol-1 K-1)
S ) salinity
T ) temperature (K)
t ) time (s)
Ug ) superficial gas velocity (m/s)
VL ) the bubble column reactor volume (m3)
YCO2 ) carbon dioxide concentration in the gas phase (Pa)
[CO2] ) carbon dioxide concentration in the liquid phase
(M)
[CO32-] ) carbonate ion concentration in the liquid phase
(M)
[H+] ) hydrogen ion concentration in the liquid phase (M)
[HCO3-] ) bicarbonate ion concentration in the liquid phase
(M)
[OH-] ) hydroxide ion concentration in the liquid phase
(M)
Greek Letters
g ) gas holdup
Fpw ) density of pure water (kg/m3)
Fsw ) density of seawater (kg/m3)

Subscripts
j ) jth tank
0 ) initial
Literature Cited
(1) Ortiz, F. J. G.; Vidal, F.; Ollero, P.; Salvador, L.; Cortes, V.;
Gimenez, A. Pilot-Plant Technical Assessment of Wet Flue Gas Desulfurization Using Limestone. Ind. Eng. Chem. Res. 2006, 45, 1466.
(2) Srivastava, R. K.; Jozewicz, W. Flue Gas Desulfurization: The State
of the Art. J. Air Waste Manage. Assoc. 2001, 51, 1676.
(3) Srivastava, R. K.; Jozewicz, W.; Singer, C. SO2 Scrubbing Technologies: A Review. EnViron. Prog. 2001, 20 (4), 219.
(4) Bigham, J. M.; Kost, D. A.; Stehouwer, R. C.; Beeghly, J. H.; Fowler,
R.; Traina, S. J.; Wolfe, W. E.; Dick, W. A. Mineralogical and Engineering
Characteristics of Dry Flue Gas Desulfurization Products. Fuel 2005, 84,
1839.
(5) U.S. Environmental Protection Agency Air Pollution Control
Technology Fact Sheet (EPA-452/F-03-034); EPA: Washington, DC
(available at http://www.epa.gov/ttn/catc/dir1/ffdg.pdf).
(6) Kikkawa, H.; Nakamoto, T.; Morishita, M.; Yamada, K. New Wet
FGD Process Using Granular Limestone. Ind. Eng. Chem. Res. 2002, 41,
3028.
(7) Nygaard, H. G.; Kiil, S.; Johnsson, J. E.; Jensen, J. N.; Hansen, J.;
Fogh, F.; Dam-Johansen, K. Full-Scale Measurements of SO2 Gas-Phase
Concentrations and Slurry Compositions in a Wet Flue Gas Desulphurization
Spray Absorber. Fuel 2004, 83, 1151.
(8) Brogren, C.; Karlsson, H. T. A Model for Prediction of Limestone
Dissolution in Wet Flue Gas Desulfurization Applications. Ind. Eng. Chem.
Res. 1997, 36, 3889.
(9) Kiil, S.; Michelsen, M. L.; Dam-Johansen, K. Experimental Investigation and Modeling of a Wet Flue Gas Desulfurization Pilot Plant. Ind.
Eng. Chem. Res. 1998, 37, 2792.
(10) Meikap, B. C.; Kundu, G.; Biswas, M. N. Modeling of a Novel
Multi-Stage Bubble Column Scrubber for Flue Gas Desulfurization. Chem.
Eng. J. 2002, 86, 331.
(11) Gerbec, M.; Stergarcek, A.; Kocjancic, R. Simulation Model of
Wet Flue Gas Desulphurization Plant. Comput. Chem. Eng. 1995, 19, S283.
(12) Oikawa, K.; Yongsiri, C.; Takeda. K.; Harimoto, T. Seawater Flue
Gas Desulfurization: Its Technical Implications and Performance Results.
EnViron. Prog. 2003, 22 (1), 67.
(13) Mandal, B. P.; Bandyopadhyay, S. S. Simultaneous Absorption of
Carbon Dioxide and Hydrogen Sulfide into Aqueous Blends of 2-Amino2-Methyl-1-Propanol and Diethanolamine. Chem. Eng. Sci. 2005, 60, 6438.
(14) Kongto, A.; Limtrakul, S.; Ngaowsuwan, K.; Ramachandran, P.
A.; Vatanatham, T. Mathematical Modeling and Simulation for Gas-Liquid
Reactors. Comput. Chem. Eng. 2005, 29, 2461.
(15) Stumm, W.; Morgan, J. J. Aquatic Chemistry; 3rd Edition; Wiley:
New York, 1996.
(16) Watten, B. J.; Sibrell, P. L.; Montgomery, G. A.; Tsukuda, S. M.
Modification of Pure Oxygen Absorption Equipment for Concurrent
Stripping of Carbon Dioxide. Aquacult. Eng. 2004, 32, 183.
(17) Deckwer, W.-D. Bubble Column Reactors; Wiley: New York,
1985.
(18) Zeebe, R. E.; Wolf-Gladrow, D. CO2 in Seawater: Equilibrium,
Kinetics, Isotopes; Elsevier: Amsterdam, 2001.
(19) Kanai, T.; Ichikawa, J.; Yoshikawa, H.; Kawase, Y. Dynamic
Modeling and Simulation of Continuous Airlift Bioreactor. Bioprocess Eng.
2000, 23, 213.
(20) Andre, G.; Robinson, C. W.; Moo-Young, M. New Criteria for
Application of the Well-Mixed Model to Gas-Liquid Mass Transfer Studies.
Chem. Eng. Sci. 1983, 38, 1845.
(21) Wise, D. L.; Houghton, G. The Diffusion Coefficients of Ten
Slightly Soluble Gases in Water at 10-60 C. Chem. Eng. Sci. 1996, 21,
999.
(22) Unver, A. A.; Himmelblau, D. M. Diffusion Coefficients of CO2,
C2H4, C3H6, and C4H8 in Water from 6 to 65 C. J. Chem. Eng. Data
1964, 9, 428.
(23) Yoon, J.-H.; Baek, J.-I.; Yamamoto, Y.; Komai, T.; Kawamura, T.
Kinetics of Removal of Carbon Dioxide by Aqueous 2-Amino-2-Methyl1,3-Propanediol. Chem. Eng. Sci. 2003, 58, 5229.
(24) Deckwer, W.-D. Bubble Column Reactors; Wiley: Chichester, U.K.,
1992.
(25) Johnson, K. S. Carbon Dioxide Hydration and Dehydration Kinetics
in Seawater. Limnol. Oceanogr. 1982, 27, 849.
(26) Eigen, M. Proton Transfer, Acid-Base Catalysis, and Enzymatic
Hydrolysis. Angew. Chem., Int. Ed. Engl. 1964, 3, 1.

6348

Ind. Eng. Chem. Res., Vol. 45, No. 18, 2006

(27) Mehrbach, C.; Culberson, C. H.; Hawley, J. E.; Pytkowicz, R.


M. Measurement of the Apparent Dissociation Constant of Carbonic Acid
in the Seawater at Atmospheric Pressure. Limnol. Oceanogr. 1973, 18,
897.
(28) Millero, F. J. Thermodynamics of the Carbon Dioxide System in
the Oceans. Geochim. Cosmochim. Acta 1995, 59, 661.
(29) Weiss, R. F. Carbon Dioxide in Water and Seawater: The Solubility
of a Non-Ideal Gas. Mar. Chem. 1974, 2, 203.
(30) Millero, F. J.; Poisson, A. International One-Atmosphere Equation
of State of Seawater. Deep-Sea Res. 1981, 28A, 625.

(31) Mandal, B. P.; Guha, M.; Biswas, A. K.; Bandyopadhyay, S. S.


Removal of Carbon Dioxide by Absorption in Mixed Amines: Modeling
of Absorption in Aqueous MDEA/MEA and AMP/MEA Solutions. Chem.
Eng. Sci. 2001, 56, 6217.

ReceiVed for reView March 24, 2006


ReVised manuscript receiVed June 29, 2006
Accepted June 29, 2006
IE0603619

You might also like