Вы находитесь на странице: 1из 12

Chemical Engineering Science 58 (2003) 3589 – 3600

www.elsevier.com/locate/ces

Rate-based modelling of SO2 absorption into aqueous NaHCO3=Na2CO3


solutions accompanied by the desorption of CO2
S. Ebrahimia; b;∗ , C. Picioreanua , R. Kleerebezema , J. J. Heijnena , M. C. M. van Loosdrechta
a Kluyver Laboratory for Biotechnology, Delft University of Technology, Julianalaan 67, 2628 BC Delft, The Netherlands
b Chemical Engineering Department, Sahand University of Technology, Tabriz, Iran

Received 21 February 2003; received in revised form 9 May 2003; accepted 16 May 2003

Abstract

A rate-based model of a counter-current reactive absorption/desorption process has been developed for the absorption of SO2 into
NaHCO3 =Na2 CO3 in a packed column. The model adopts the 9lm theory, includes di:usion and reaction processes, and assumes that
thermodynamic equilibrium among the reacting species exists in the bulk liquid. Model predictions were compared to experimental data
from literature. For the calculation of the absorption rate of SO2 into NaHCO3 =Na2 CO3 solutions and concomitant CO2 -desorption, it is
important to take into account all reversible reactions simultaneously. It is clear that the approximate analytical based model cannot be
expected to predict the absorption rates under practical conditions because of the complicated nature of the reactive absorption processes.
The rigorous numerical approach described here only requires de9nition of the individual reactions in the system, and subsequent solution
is independent of speci9c assumptions made, or operational variables like pH or compound concentrations. As an example of the =exibility
of this approach, additional calculations were conducted for SO2 absorption in a phosphate-based bu:er system.
? 2003 Elsevier Ltd. All rights reserved.

Keywords: Absorption; Flue gas; Sulfur dioxide; Numerical analysis; Multiphase reactions; Modelling

1. Introduction into the atmosphere, it is apparent that studies on =ue gas


desulfurization methods and development of =ue gas desul-
Sulfur dioxide in =ue gas generated as a result of com- furization plants have become numerous.
bustion of fossil fuel in, e.g., thermal power plants, etc., Although various processes have been proposed for =ue
is the main cause of global environmental problems such gas desulfurization, the wet-type scrubbing is still the dom-
as air pollution and acid rain. Sulfur dioxide has also been inant process. The wet-type processes include methods us-
reported to support the reactions that create ozone deple- ing alkaline solutions of sodium, calcium and magnesium
tion in the stratosphere (Karlsson, 1997). Many countries compounds as absorbent. The sodium method above all
have therefore adopted strict regulations regarding SO2 is excellent in reactivity between the absorbent and SO2 ,
emissions from coal- and oil-9red boilers in power plants, but the sodium compounds used are relatively expensive
which are one of the primary sources of SO2 emissions. The for this purpose. For this reason, the calcium method us-
sulfur dioxide content of the =ue gas generated is usually ing relatively cheap calcium compounds such as calcium
quite small and below about 0.1– 0.4% by volume (Astarita, carbonate is most widely employed as a =ue gas desulfur-
Savage, & Bisio, 1983). However, the volume of the gas ization system for large boilers in power plants. However,
produced globally is so large that considerable amount of when sodium compounds are used in a closed-loop system
sulfur dioxide is introduced into the atmosphere. In view of (with regeneration of solutions), it is estimated that the costs
the large number of processes which introduce sulfur dioxide will be comparable or lower than those of calcium-based
processes.
In this study, we have developed a combined chem-

ical/biological process for SO2 removal, NaHCO3 re-
Corresponding author. Kluyver Laboratory for Biotechnology, Delft
University of Technology, Julianalaan 67, 2628 BC Delft, The Nether-
covery and elemental sulfur production. An aqueous
lands. Tel.: +31-15-278-1551; fax: +31-15-278-2355. NaHCO3 =Na2 CO3 solution is used as absorbent in a
E-mail address: s.ebrahimi@tnw.tudelft.nl (S. Ebrahimi). closed-loop process, schematically depicted in Fig. 1.

0009-2509/03/$ - see front matter ? 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0009-2509(03)00231-8
3590 S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600

Cleaned H2S/ H2
Gas NaHCO3
Fe2(SO4)3

1 2 3 4 1 SO2 Absorber
2 Sulfite Reduction Bioreactor
3 H2S Absorber
4 Ferric Regeneration Bioreactor

Flue Gas HSO3- H2S/ H2 Air

H2/ CO2 Elemental Sulfur

Fig. 1. Process scheme of chemo-biological SO2 removal.

The process consists of two major liquid circulation loops. studies, approximate analytical solutions were used, which
The 9rst loop contains a sulfur dioxide absorber, which con- limits the applicability to the speci9c conditions investigated
verts SO2 into HSO− 3 , and a sul9te reduction bioreactor based on simplifying assumptions. Therefore, most of the
where HSO− 3 is converted into H2 S, using H2 . In the sec- above-mentioned aspects cannot be studied with the analyti-
ond loop, an aqueous Fe2 (SO4 )3 solution is used as H2 S ab- cal model. As it will be shown later, the use of simpli9ed ex-
sorbent. H2 S is absorbed and oxidized to elemental sulfur, pressions may easily lead to erroneous results. In this work
while Fe3+ is reduced to Fe2+ . Elemental sulfur is removed the expressions for the rate of absorption describing the in-
from the solution by a separator, and the reactant Fe3+ is terfacial =uxes were implemented in design calculations for
regenerated from Fe2+ by biological oxidation in an aer- absorber columns.
ated bioreactor (Ebrahimi, Kleerebezem, van Loosdrecht, &
Heijnen, 2003).
The present study aims at developing a rigorous 2. System description
rate-based steady-state model for the design and simulation
of packed columns used for absorption of SO2 into aqueous A gas mixture containing SO2 is fed at the bottom of
NaHCO3 /Na2 CO3 . The goal is to estimate the absorption the column, schematically shown in Fig. 2, at a volumet-
rates, enhancement factors and concentration pro9les of ric =ow rate G. The gas comes in contact with a liquid
all chemical species involved. The model is based on the containing the absorbing reactant =owing from the top at
9lm theory of gas absorption. A numerical solution of the a =ow rate L. The following assumptions were made to
model was used, allowing for integration of the following obtain a mathematical model describing the mass trans-
aspects: fer with simultaneous chemical reaction in a di:erential
absorber:
• reversibility e:ect of all reactions involved;
• the formation, absorption and desorption of CO2 ; • mass transfer can be described by the 9lm theory;
• the e:ect of 9nite CO2 reaction rate, and the second dis- • isothermal, steady state operation;
sociation of CO2 ; • the gas and liquid phases are in plug =ow; and
• the contribution of the gas phase resistance to mass trans- • solute concentrations are low so that the amount of ab-
fer; and sorption and reactions do not cause a signi9cant change
• unequal di:usivities of the species. in the =ow rates of gas and liquid.

There have been a few studies on the mechanism of chem- 2.1. Chemistry
ical absorption of SO2 into aqueous solutions (e.g., Chang
& Rochelle, 1981, 1985; Hikita, Asai, & Tsuji, 1977; Hikita When dilute sulfur dioxide is absorbed into aqueous
& Konishi, 1983; Sada, Kumazawa, & Butt, 1980). In those NaHCO3 =Na2 CO3 solutions, the following reactions should
S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600 3591

Cleaned gas Liquid in (1997) and Astarita et al. (1983), respectively:


17265:4
log(k3 ) = 329:85 − 110:541 log(T ) −
T
L (m3/s) K3 (1=s) and T (K) (9)
Gas Liquid
CO2 log(k4 ) = 13:635 − 2895=T + 0:08I

SO2 K4 (m3 =Kmol ·s) and T (K): (10)


dz
In any solution of pH ¿ 10, the CO2 reaction rate accord-
G (m3/s) ing to Eq. (4) will be more than 30 times higher than that
consumed in reaction (3). Thus, reaction (3) is normally of
negligible importance in determining the rate of absorption
of CO2 into alkaline solutions with pH ¿ 10. However, at
z
pH values below 8, as in our process, reaction (3) is faster
Liquid out
Flue gas than reaction (4) (Astarita et al., 1983; Danckwerts, 1970).
In this model the sul9te oxidation reaction has not taken
Fig. 2. Countercurrent gas–liquid contacting in packed column. into account because of the relatively low reaction rate (Eden
& Luckas, 1998).

be considered: 2.2. Reactor model


SO2 + H2 O  H + +
HSO−
3 ; (1)
The steady-state mass balance for SO2 and CO2 in the
gas phase plug =ow through the column can be written as
HSO− + 2−
3  H + SO3 ; (2)  
G dpSO2 ;b g
= −NSO 2 ;i
a; (11)
RTS dz
CO2 + H2 O  H+ + HCO−
3 ; (3)
 
G dpCO2 ;b g
= −NCO 2 ;i
a: (12)
CO2 + OH−  HCO−
3 ; (4) RTS dz
The steady-state mass balance for each species (A) in the
HCO− + 2−
3  H + CO3 ; (5) bulk liquid phase, also assumed in plug =ow, is
L dCA; b
H2 O  H+ + OH− : (6) = NA;l b a + RlA; b l : (13)
S dz
Reaction (1) is very fast, with an estimated forward rate The mass balances for individual species can be combined
constant of 3:4 × 106 (1/s) (Chang & Rochelle, 1981). Re- to obtain mass balances for total S and C species. Further-
actions (2), (5) and (6) are even faster than reaction (1) more, according to the law of elemental conservation, the
since they are based on simple proton transfer and there- steady-state =uxes of total S and C species through the liq-
fore regarded as instantaneous. Consequently, instantaneous uid 9lm are constant and are equal to SO2 and CO2 =uxes at
g g
equilibrium is assumed for the reactions (1), (2), (5) and the gas side (NSl total; b =NSO 2 ;i
and NCl total; b =NCO2 ;i
). There-
(6) throughout the liquid 9lm. fore, a continuity equation for total S and C species in the
The hydrolysis of CO2 is slow. Two reactions (3) and liquid phase can be written as
(4) may occur when CO2 is absorbed by aqueous alka-
L dCS total; b g
line solutions (Danckwerts, 1970). Forward reaction (3) is = NSO2 ;i
a; (14)
S dz
pseudo-9rst order, whereas forward reaction (4) is second
order. The rate of reversible reactions (3) and (4) can be ex- L dCC total; b g
pressed as the di:erence between the forward and the back- = NCO2 ;i
a: (15)
S dz
ward reaction rates, respectively: g
Eqs. (11), (12), (14) and (15) contain the =uxes NSO 2 ;i
and
g
RCO2 ;1 = k3 (CCO2 − CHCO− CH+ =K3 ); (7) NCO2 ;i . These =uxes are related to the concentration gradients
3
of SO2 and CO2 at the gas/liquid interface (see Eqs. (27)
RCO2 ;2 = k4 (CCO2 COH− − CHCO− =K4 ): (8) and (29)), so that the concentration pro9les of SO2 and CO2
3
in the liquid 9lm are required. For this purpose, the mass
The rate constant of the forward hydrolysis reactions (3) balance equations in liquid 9lm for di:erent species have to
and (4) can be calculated according to Brogren and Karlsson be integrated.
3592 S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600

Interface total sodium:


Gas bulk phase Liquid bulk phase d 2 CSO2 d 2 CHSO− d 2 CSO2−
NA,i DSO2 + D −
HSO3
3
+ D 2−
SO3
3
= 0;
d x2 d x2 d x2
PA,b (18)

PA,i• d 2 CCO2 d 2 CHCO− d 2 CCO2−


• CA,i DCO2 + DHCO− 3
+ DCO2− 3
= 0;
CA,b
d x2 3 d x2 3 d x2
(19)
SO2 (aq), HSO3 -, SO3 2-,
CO2, SO2 CO2 (aq), HCO3 -, CO32-,
x H+,OH-, Na+ d 2 CNa+
DNa+ = 0: (20)
x=0 x=δ d x2
Because the hydrolysis reaction of CO2 is slow, a separate
Fig. 3. Schematic diagram of the two 9lm model.
material balance for CO2 has to be included:
d 2 CCO2
DCO2 − RCO2 ;1 − RCO2 ;2 = 0: (21)
2.3. Film region d x2
It has been shown that the impact of any electric potential
The mass transfer between the gas and liquid phase is de- gradient on the =ux of ions may be disregarded under =ue
scribed based on the 9lm model. Even though 9lm renewal gas desulfurization conditions as long as the mass =ux equa-
theory for gas liquid mass transfer represents the most re- tions are combined with a =ux charge equation (Brogren
alistic mass transfer model conditions, it is rarely used in & Karlsson, 1997). Therefore, the mass balances must be
practice since this leads to a complex mathematical descrip- combined with a =ux of charge balance when the potential
tion and evaluation of reacting systems. In the 9lm model, gradient is disregarded. The =ux of charge in the liquid 9lm:
all the resistance to mass transfer is concentrated in a thin
9lm adjacent to the phase interface and the mass transfer dCH+ dCNa+ dCOH−
DH+ + DNa+ − DOH−
occurs within this 9lm by steady-state molecular di:usion dx dx dx
(Kenig, Schneider, & GLorak, 1999) (Fig. 3). Reactions be- dCHSO− dCSO2−
tween the absorbed gas and the liquid reactants are assumed − DHSO− 3
− 2DSO2− 3
3 dx 3 dx
to be complete within the liquid 9lm. This implies that the
bulk liquid is in a state of chemical equilibrium. dCHCO− dCCO2−
Since the chemical reactions occur only in the liquid − DHCO− 3
− 2DCO2− 3
= 0: (22)
3 dx 3 dx
phase, the molar =ux of each component in the gas-phase
9lm is constant along the direction x (i.e., normal to the Since instantaneous equilibrium is assumed for the reactions
gas/liquid interface). The di:erential equations describing (1), (2), (5) and (6) throughout the liquid 9lm, the chemi-
the di:usion of each component A in the gas phase (without cal equilibrium relations of these reactions also apply at all
reaction) have the following form: points in liquid phase:
CH+ CHSO−
dNAg 3
= K1 ; (23)
= 0: CSO2
dx
The di:erential equations describing the process of di:usion CH+ CSO2−
3
= K2 ; (24)
with simultaneously reaction of each species A in the liquid CHSO−
3
phase are
CH+ CCO2−
dNAl 3
= K5 ; (25)
= RA : (16) CHCO−
dx 3

The simple Fick’s law is used to express the di:usive =ux CH+ COH− = K6 : (26)
of each species:
dCA 2.4. Boundary conditions
NA = −DA : (17)
dx
The model system consisting of mass and charge balances
After replacing the =uxes (17) in Eqs. (16), the mass bal- Eqs. (18)–(22) and the chemical equilibrium relations (23)–
ances for individual species can be combined to obtain (26) must be completed by the boundary conditions relevant
the following balances for total sul9te, total carbonate and to the 9lm model at x = 0 and .
S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600 3593

2.4.1. Boundary conditions at the interface (x = 0) 2.4.2. Boundary conditions in bulk liquid (x = )
The absorption rate of SO2 is equal to the sum of the Equilibrium is assumed for the all reactions in the bulk
=uxes of the sulfur species at the gas–liquid interface, which liquid, therefore, besides mass action laws (23)–(26) also
also must be equal to the =ux of SO2 in the gas 9lm: equilibrium equation:
g kg; SO2 CH+ CHCO−
NSO2 ;i
= (pSO2 ;b − pSO2 ;i ) 3
= K3 ; (34)
RT CCO2
  
dCSO2  dCHSO−  is considered, together with another four equations obtained
= − DSO2 + DHSO− 3

d x x=0 3 dx  from mass balances for total sulfur species, total carbon
x=0
  species, sodium and a charge balance:
dCSO2− 
+ DSO2− 3
 ; (27) Ctot; S = CSO2 ;b + CHSO− ;b + CSO2− ;b ; (35)
3 dx  3 3
x=0

where Ctot; C = CCO2 ;b + CHCO− ;b + CCO2− ;b ; (36)


3 3

pSO2 ;i = HSO2 CSO2 ;i : (28) Ctot; Na = CNa+ ; b ; (37)


Since the rate of the hydrolysis reaction is slow, the =ux of
CO2 in the gas 9lm is equal to the =ux of dissolved CO2 at CH+ + CNa+ − COH− − CHSO− − 2CSO2− − CHCO−
3 3 3
the interface:
− 2CCO2− = 0: (38)
g 3
NCO2 ;i
= kg; CO2 =RT (pCO2 ;b − pCO2 ;i )
 The system of equations (18)–(22) together with the
dCCO2  four equilibrium equations (23)–(26) and the boundary
= −DCO2 ; (29)
d x x=0 conditions (27)–(38) are used to 9nd the concentration
pro9les of the nine unknown species in the liquid 9lm
where (SO2 ; HSO− 2− − 2− + + −
3 ; SO3 ; CO2 ; HCO3 ; CO3 ; Na ; H ; OH ) at
pCO2 ;i = HCO2 CCO2 ;i : (30) each position z in the column height. These concentration
pro9les allow the calculation of the =uxes of SO2 and
The =ux of HCO− 2−
3 and CO3 at the interface must be zero CO2 . The =uxes are needed for integration of di:erential
because the transport of carbon species is represented by mass balance equations in the bulk gas and liquid along the
CO2 : column (11), (12), (14) and (15).
  The liquid 9lm thickness and reactor height were both dis-
dCHCO−  dCCO2− 
DHCO− 3
 + DCO2− 3
 = 0: (31) cretized in a spatially uniform grid and second-order 9nite
3 dx  3 dx  di:erencing was applied. The resulting system of non-linear
x=0 x=0
algebraic equations was solved numerically with a tradi-
The =ux of Na is zero at the interface: tional Newton-based method.

dCNa+ 
DNa+ =0 (32)
dx  x=0
3. Estimation of physical properties and model parameters
and also the net =ux of charge must be zero at x = 0:
 
dCH+  dCNa+  The di:usion coeNcients of gases were calculated from
DH + DNa
dx  dx  the equation given by Reid, Prausnitz, and Poling (1988).
+ +

x=0 x=0
 The di:usion coeNcients in liquid, used in model calcula-

dCOH−  dCHSO−  tions, are listed in Table 1. The di:usion constants were
− DOH− − DHSO− 3
 extrapolated from 25◦ C to 55◦ C using the Stokes–Einstein
dx  x=0
3 dx 
x=0 equation:
 
dCSO2−  dCHCO−  DA 
− 2DSO2− 3
 − DHCO− 3
 = constant: (39)
3 dx  3 dx  T
x=0 x=0
 Correlations for the determination of the dissociation equi-
dCCO2−  librium constants and solubility values for SO2 and CO2
− 2DCO2− 3
 = 0: (33)
3 dx  as a function of temperature are given in Table 2. The ac-
x=0
tivity coeNcients take into account the deviations of the
Instantaneous equilibrium is assumed for the reactions (1), thermodynamic equilibrium of real mixture from those of
(2), (5) and (6) at the interface, thus Eqs. (23)–(26) apply an ideally diluted solution. Depending on whether species
also at x = 0. are charged or not, two di:erent types of activity coeNcient
3594 S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600

Table 1 2
E:ective di:usitivities in water at 25◦ C and in9nite dilution, DA (Vanysek, a = ap (1 − exp[ − 1:45(L =al ):0:1 (ap L =g"2l )−0:05
2001)
2
Species DA (m2 =s)
×(L =ap %"l )0:2 (%=%c )−0:75 ]): (44)

H+ 9:311 × 10−9 The mass transfer coeNcient kl was used for the estimation
Na+ 1:334 × 10−9 of the thickness of mass transfer boundary layer in the liquid
OH− 5:273 × 10−9 phase, , using the di:usivity for SO2 as reference:  =
SO2 (aq) 1:83 × 10−9 DSO2 =kl; SO2 .
HSO− 3 1:545 × 10−9
SO2−
3 0:959 × 10−9
CO2 (aq) 1:91 × 10−9
HCO− 3 1:185 × 10−9 4. Results and discussion
CO2−
3 0:923 × 10−9
4.1. Model validation

expressions are: Validation of the developed 9lm model was established


using SO2 absorption data into aqueous Na2 CO3 solutions
• Simple salting relation for uncharged species A (Gerard, as presented by Hikita and Konishi (1983).
Segantini, & Vanderschuren, 1996; Hikita et al., 1977) These authors have carried out absorption experiments
using a baUed agitated vessel operated batchwise with
log A = 0:076I: (40) respect to the liquid, and compared the experimental re-
sults with an approximate analytical solution based on
• Extended Debye–HPuckel model (B dot equation) for in-
the Leveque model. They proposed a two reaction plane
dividual ions (Parkhurst & Appelo, 1999)
√ model and showed that the measured absorption and des-
AzA2 I orption rates were in good agreement with the theoretical
log A = − √ + ḂI: (41) predictions.
1 + BaA I
Using the same parameters and conditions as used by
The following correlations, which cover a wide range of Hikita and Konishi (1983), the pH and concentration pro-
packing types, sizes and test systems, were used to calcu- 9les for all species were calculated and the results are
late individual mass transfer coeNcients (kl and kg ) and shown in Fig. 4. These concentration pro9les clearly show
the interfacial area (a) for packed column (Onda, Sada, & the rapid depletion of SO2 near the gas–liquid interface
Okumoto, 1968a; Onda, Takahashi, & Okumoto, 1968b): and agree with the existence of the two reaction planes
kg (RT=ap Dg ) = 5:23(G  =ap g )0:7 (g ="g Dg )1=3 in this system. These two reaction planes divide the liq-
uid phase into three regions. Therefore, the intuitive pro-
×(ap dp )−2 ; (42) 9les suggested by Hikita and Konishi (1983) for similar
cases agree well with the more exact prediction of our
kl ("l =gl ) = 0:0051(L =al )2=3 (l ="l Dl )−0:5 model.
As suggested by Hikita and Konishi (1983) the following
×(ap dp ):0:4 ; (43) four reactions may occur irreversibly and instantaneously at

Table 2
E:ect of temperature on dissociation constants of weak electrolytes in water, K, and solubility coeNcients of gases in pure water, H

Reaction A B C D Range of Ref.


validity, ◦ C

ln(K) = A=T + B ln(T ) + C(T ) + D K(mol=kg) and T (K)


CO2 + H2 O  HCO− 3 +H
+ −12092:1 −36:7816 0 235.482 0 –225 a
HCO− 3  H + + CO2−
3 −12431:7 −35:4819 0 220.067 0 –225 a
SO2 + H2 O  HSO− 3 + H + 26404.29 160.3981 −0:275224 −924:6255 –- b
HSO− +
3  H + SO3
2−
−5421:93 −4:6899 −0:0498769 43.3136 –- b
H2 O  OH− + H+ −13445:9 −22:4773 0 140.932 0 –225 a

ln(H ) = A=T + B ln(T ) + C(T ) + D H (atm:kg=mol) and T (K)


CO2 −6789:04 −11:4519 −0:010454 94.4914 0 –250 a
SO2 −5578:8 −8:76152 0 68.418 0 –100 a

(a) Edward, Maurer, and Prausnitz (1978); (b) Xia, Rumpf, and Maurer (1999).
S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600 3595

0.1 12
SO2 CO2
HSO3- HCO3-
SO3-2 CO3-2
H+ pH 10
0.08

Concentrations (kmol/m 3)
Region 3
Region 1 Region 2 8
0.06
1st.reaction plane 2nd reaction plane

pH
6

0.04
4

0.02
2

0 0
0 0.2 0.4 0.6 0.8 1
Dimensionless distance from the interface

Fig. 4. Concentration pro9les calculated for the absorption of SO2 into aqueous Na2 CO3 solution in the liquid 9lm. Na2 CO3 : 0:0398 kmol=m3 ;
SO2; i : 0:0379 kmol=m3 and CO2; i : 0:0000603 kmol=m3 ;  = 0:00013 m.

the 9rst reaction plane: this particular case. However, the analytical model cannot
be expected to predict the absorption/desorption rates for a
SO2 + SO2−
3 + H2 O → 2HSO−
3 ; (45)
wide range of conditions. At high concentration of Na2 CO3 ,
for example, the two reaction plane model is not realistic be-
SO2 + HCO− −
3 → CO2 + HSO3 ; (46)
cause the assumption of zero concentrations for speci9c re-
actants will not hold. Moreover, the model presented in this
H+ + HCO−
3 → CO2 + H2 O; (47)
study gives a more general solution for chemical absorption
of SO2 at di:erent concentration of aqueous alkaline solu-
H+ + SO2− −
3 → HSO3 : (48)
tion (NaOH and NaHCO3 =Na2 CO3 ) and various SO2 and
At the second reaction plane, the following instantaneous CO2 partial pressures in the gas phase. In addition, the pH
irreversible reaction take place: pro9le is directly calculated in the liquid 9lm.
CO2− − 2− −
3 + HSO3 → SO3 + HCO3 : (49)
In addition, the hydrolysis reaction of SO2 may occur in- 4.2. Scrubber design
stantaneously and reversibly in region 1. Because in regions
2 and 3, CO2 and SO2−3 ions coexist, these two species can The proposed model is applied to the design of a scrubber
react according to for the removal of SO2 from the =ue gas in a power plant. In
the present example, =ue gas from a 600 MW power plant
CO2 + SO2− − −
3 + H2 O  HSO3 + HCO3 : (50)
containing 1000 ppm SO2 (0:1 vol%) is to be puri9ed by
Some of the CO2 liberated at the 9rst reaction plane di:uses absorption into an aqueous NaHCO3 =Na2 CO3 solution. The
towards the bulk of the liquid. The remainder of CO2 dif- inlet =ow rate of gas is 2 × 106 Nm3 =h. The temperature is
fuses towards the gas–liquid interface and desorbs into the 110◦ C and the total pressure is 1:1 bar. The partial pressure
gas phase if the concentration of CO2 at the 9rst reaction of CO2 in the =ue gas amounts 0:14 bar.
plane is greater that at the interface. If incoming gas streams are at elevated temperatures, the
The predictions of the proposed analytical model by 9rst function of the scrubber normally is to saturate the gas
Hikita and Konishi (1983) and our theoretical model for with water and cool the gas. Usually the cooling is adiabatic,
the absorption and desorption rates for SO2 and CO2 that is, the gas is saturated by the scrubbing liquid until the
against the experimental results are shown in Figs. 5a and temperatures of the water and the gas are the same. The ob-
b respectively. It can be seen that the theoretical rates jective of this preliminary step, which may be achieved in
computed with both models are in good agreement with an early stage of the scrubber or in a preliminary satura-
the measured absorption and desorption rates in Na2 CO3 tion chamber, is to reduce the volume of gas entering the
solutions. subsequent stages of the scrubber. Hence, the preliminary
The proposed analytical model by Hikita and Konishi step reduces the equipment size and lower the total capital
(1983) is in good agreement with the experimental data in cost. Herewith, evaporation can be prevented in subsequent
3596 S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600

15 7
Hikita and Konishi Hikita and Konishi
This work This work
6

(NCO2) cal. ×10 (kmol/m2s)


(NSO2) cal. ×106 (kmol/m2s)
5
10
4

6
3
5
2

0 0
0 5 10 15 0 1 2 3 4 5 6 7
(NSO2)exp. ×10 (kmol/m2s) (NCO2) exp. ×10 (kmol/m2s)
6 6
(a) (b)

Fig. 5. SO2 absorption into aqueous Na2 CO3 solutions. Comparison between theoretical absorption/desorption rates and experimental data from Hikita and
Konishi (1983). Na2 CO3 =39:8– 994 mol=m2 ; SO2; i =22:5–48:9 mol=m3 ; CO2; i =0:0548–0:376 mol=m3 . (a) SO2 absorption rate; (b) CO2 desorption rate.

stages where loss of water vapour might cause precipitation Table 3


of unwanted compounds on scrubber surfaces, and subse- Design parameters and design results for the SO2 absorption column
quent stages of the scrubber are protected from the poten- T (K) 328
tially corrosive e:ects of heated gas (McCarthy, 1980). For P (bar) 1.1
these reasons, design calculations were made for saturated
=ue gas at a temperature of 55◦ C. Flow rates
G (m3 =s) 556
The desulfurization methods using sodium compounds L (m3 =s) 1.11
in the absorbent liquid can be generally classi9ed into
spraying, wetted-wall and bubbling systems, depending Inlet gas phase composition
on the particular gas–liquid contacting method. Since the pSO2 (bar) 0.0011
packed and spray column systems are considerably more pCO2 (bar) 0.14
popular and reliable, here the results of model calculations Inlet liquid phase composition
for a packed column are presented. CNaHCO3 (kmol=m3 ) 0.05
The liquid phase enters the column at the top and =ows
in countercurrent with the gas. The height of the column is Mass transfer coeNcients
determined for 95% removal of SO2 from the =ue gas when kG; SO2 (m=s) 0.036
kG; CO2 (m=s) 0.0402
0:05 kmol=m3 bicarbonate solution is used as absorbent. kL; SO2 (m=s) 0.0002852
The choice of the packed column diameter is based on the kL; CO2 (m=s) 0.000303
60% of =ooding condition. The packed column calcula-
tion was based on using a relatively high capacity pack- Interfacial area
ing material (35 mm Pal rings). In practice, the gas =ow a (m2 =m3 ) 84.1
% of =ooding 60
rate should be split and a few smaller columns would op-
erate in parallel. The calculated results are summarized in Calculated column height (m) 2.09
Table 3. Calculated column diameter (m) 19.1

4.2.1. Concentration pro:les in the liquid :lm


The typical calculated pH and concentration pro9les for
the various chemical species in the liquid 9lm in the packed that the substantial reactions take place only at this reaction
column are shown in Fig. 6. SO2 depletion close to the gas– plane. The most important reaction in 9rst region is the hy-
liquid interface corresponds to the very fast reactions, caus- drolysis reaction of SO2 , which occurs instantaneously and
ing a high enhancement of SO2 transfer. It is apparent that a reversibly.
large pH drop occurs close to the gas–liquid interface due to
SO2 absorption. One reaction plane can be recognized via the 4.2.2. Partial pressure and mass transfer rates along the
concentration pro9les in the liquid 9lm. The reaction plane column
is located at x = ', and it divides the liquid 9lm into two re- The partial pressure pro9les of SO2 and CO2 in the gas
gions. The 9rst region is near the interface, where concentra- bulk along the column are shown in Fig. 7. As the gas moves
tion of SO2 is signi9cant, and the second one is near the bulk up the column the partial pressure of SO2 decreases due to
liquid where concentration of SO2 is negligible. It seems absorption. At the top of the column the partial pressure of
S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600 3597

0.03 8 2.E-06
SO2*50 CO2
HSO3- HCO3- NSO2
SO3-2 CO3-2
0.025 H*100 pH
NCO2
7

NSO2, NCO2 (kmol/m .s)


1.E-06
Concentrations (kmol/m3)

2
0.02 Region 1
Region 2

Reaction
plane

pH
0.015 6
0.E+00
0 0.5 1 1.5 2 2.5
0.01
x'=λ/δ
5
0.005 -1.E-06

0 4
0 0.2 0.4 0.6 0.8 1 -2.E-06
Dimensionless distance from the interface
Column height from bottom, z (m)
Fig. 6. Absorption of SO2 and CO2 in 0:05 M NaHCO3 solution
Fig. 8. Mass transfer rate of SO2 and CO2 in the =ue gas along the
in a packed column. Concentration pro9les in the liquid 9lm at
column (NaHCO3 = 0:05; L=G = 2 × 10−3 ).
z = 0 m;  = 1:2 × 10−5 . Concentrations of SO2 and H+ shown in the
graph were multiplied by a factor of 50 and 100 respectively.

9
pH_b
1.4 1.44 pH_i
PSO2 8
1.2 PCO2

1 7
1.435
pCO2×1000 (atm)
pSO2×1000 (atm)

pH

0.8
6

0.6
1.43 5
0.4

0.2 4
0 0.5 1 1.5 2 2.5
Column height from bottom, z (m)
0 1.425
0 0.5 1 1.5 2 2.5
Fig. 9. pH pro9les in the liquid bulk and interface along the column
Column height from bottom, z (m) (NaHCO3 = 0:05; L=G = 2 × 10−3 ).

Fig. 7. Partial pressure of SO2 and CO2 in the =ue gas along the column
(NaHCO3 = 0:05; L=G = 2 × 10−3 ).
Positive values of the component =ux correspond to absorp-
tion, whereas negative values represent desorption.
CO2 is higher than the initial backpressure from the aque-
ous solution. Hence, CO2 is absorbed in the fresh alkaline 4.2.3. pH pro:les along the column
solution at the column top. As the pH decreases from the In Fig. 9, the pH pro9le in the liquid bulk and at the in-
top to the bottom due to the SO2 absorption, the concen- terface along the column is shown. Since the absorption of
tration of CO2 in the liquid increases and the direction of SO2 into aqueous NaHCO3 solutions is accompanied by the
the CO2 =ux at the gas/liquid interface is reversed because desorption of CO2 , there is no strong change in pH of the
the concentration of CO2 in the liquid bulk becomes larger bulk liquid (only 1.5 units of pH). A steep and =ux depen-
than at the interface. CO2 desorption also occurs when the dent pH-gradient in the liquid 9lm layer is the result from
concentration in the reaction plane is larger than that at the the complex absorption of SO2 and absorption–desorption
interface, even if in the bulk it may be lower. This CO2 con- behaviour of CO2 .
centration peak makes therefore possible the apparition of
opposed directions of CO2 di:usion in the liquid 9lm. 4.2.4. Enhancement factor and individual :lm resistances
The presence of both absorption and desorption in the The model results enable us to evaluate the enhance-
column is further illustrated by the interfacial mass transfer ment factor for chemical absorption at assigned gas and
rates of SO2 and CO2 along the column, shown in Fig. 8. liquid bulk compositions. The enhancement factor, E,
3598 S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600

100000
100

Contribution of gas film resistance (%)


10000 95

90
1000
E SO2

85
100
80

10 75

1 70
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
Column height from bottom, z (m)
Column height from bottom, z (m)
Fig. 11. Contribution of the gas 9lm resistance for SO2 transfer along
Fig. 10. Chemical enhancement factor for SO2 absorption along the
the column (NaHCO3 = 0:05; L=G = 2 × 10−3 ).
column (NaHCO3 = 0:05; L=G = 2 × 10−3 ).

considers the e:ect of the chemical reactions on the 4


liquid-side mass transfer, and is de9ned as the ratio between
the actual absorption rate and the rate that would be observed
with the same driving force in the absence of chemical
Column height, z (m)

reactions: 3
NSO2
E= : (51)
(DSO2 (aq) =)(CSO2 (aq) |x=0 − CSO2 (aq) |x= )
The enhancement factor for SO2 absorption (Fig. 10) has
2
its highest value in the top of the absorber, where the par-
tial pressure of SO2 is low and the alkalinity of the liquid
is high. The enhancement factor decreases towards the bot-
tom of the column due to the pH drop. However, a major
part of the scrubber operates at a high enhancement factor. 1
0 0.05 0.1 0.15 0.2 0.25
That means the absorption is strongly enhanced by the very NaHCO3 concentration (kmol/m3)
fast reactions and therefore the gas side resistance to mass
transfer becomes important. Fig. 12. Dependency of the design column height on the bicarbonate
At the top of the absorber the contribution of the rela- concentration in the inlet liquid (NaHCO3 = 0:05; L=G = 2 × 10−3 ).
tive gas 9lm resistance has its largest value as illustrated
in Fig. 11. In the condition used in these calculations,
the contribution of the gas 9lm resistance is about 100% is due to the fact that the mass transfer rate for CO2 is
and 75% in the top and bottom of the absorber respec- not a:ected by chemical reaction and transfer of SO2
tively. The calculations show that the absorption of SO2 is controlled by gas-side resistance. Consequently, the
within a packed scrubber to a large extent is gas side rate of SO2 removal can be best improved by creat-
controlled. ing more turbulence in the gas-phase and thereby higher
gas phase mass transport coeNcients and by increasing
4.2.5. E=ect of bu=er concentration the surface area available for mass transfer.
In Fig. 12 the e:ect of the sodium bicarbonate concen- In the proposed process for SO2 removal, aqueous ab-
tration on the calculated height of the column is shown. sorbent solution from the scrubber is regenerated in a sul-
It is apparent that sodium bicarbonate solution above 9te reduction bioreactor (see Fig. 1). From a biological
0:05 kmol=m3 has enough bu:er capacity for SO2 absorbed, point of view, pH variation of the aqueous solution should
and therefore, a higher concentration of sodium bicarbon- be limited. Increasing bu:er capacity of the aqueous so-
ate does not have a signi9cant e:ect on the mass transfer lution can decrease the pH drop along the column height.
coeNcient of either SO2 or CO2 . Therefore, there is no Even though by using higher concentrations of the bicarbon-
substantial change in the calculated column height when ate solution bu:er capacity can be increased, on the other
bicarbonate concentrations higher than 0:05 kmol=m3 are hand pH in the column will be increased (see Fig. 13),
applied. This invariance of the necessary column height which is unfavourable for microorganisms. Therefore, it
S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600 3599

9 ing the concentration of SO2 in the =ue gas from 1000 to


2000 ppm, the degree of the desulfurization decrease from
8.5 95% to 89.5%. Therefore, it can be concluded that the de-
gree of desulfurization is not very sensitive to SO2 content
8
of the =ue gas.
pH

7.5

7 5. Conclusion

6.5 In this study, a general approach to the modelling and


design of multicomponent reactive absorption/desorption
6
0 0.5 1 1.5 2 2.5 is presented. The complicated absorption process of SO2
Column height from bottom, z (m) into aqueous NaHCO3 / Na2 CO3 solutions accompanied by
the desorption of CO2 is well described by the proposed
Fig. 13. pH pro9le in the liquid bulk along the absorber column; the model. The model developed, in spite of simple description
e:ect of using di:erent bu:ers (NaHCO3 = 0:05; L=G = 2 × 10−3 ).
of the mass transfer, is capable of accurate prediction of
(V)NaHCO3 = 0:1 M; ( ) NaHCO3 = 0:05 M; (◦) NaHCO3 = 0:03 M;
(—) NaHCO3 = 0:05 M, Na2 HPO4 = 0:025 M and NaH2 PO4 = 0:025 M; the transfer rates of absorption/desorption and of the en-
(- - -) NaHCO3 =0:03 M, Na3 HPO4 =0:015 M and NaH2 PO4 =0:015 M. hancement factors. Moreover, the model could predict the
concentrations of all chemical species at any point of the
absorption column. The model is validated using a SO2 ab-
sorption process in aqueous Na2 CO3 solutions. The model
100
presented can be applied to any highly complicated reac-
tive absorption/desorption processes. It should be stressed
that analytical approximations are often oversimpli9ed and
Degree of desulfurization (%)

95 cannot be expected to predict the absorption/desorption


rates for a wide range of conditions and so under practical
conditions.
90

Notation
85
a wetted surface area of packing, m2 =m3
ap total surface area of packing, m2 =m3
C molar concentration in the liquid phase, kmol=m3
80 l
0 1000 2000 3000
CA; b concentration of the component A in the liquid bulk,
SO2 Concentration (ppm) kmol=m3
dp diameter of packing, m
Fig. 14. The e:ect of the =ue gas SO2 concentration on the column D di:usion coeNcient, m2 =s
performance (NaHCO3 = 0:05 M; L=G = 2 × 10−3 ; H = 2:09 m). EA enhancement factor
g gravitational constant, m=s2
G gas volume =ow rate, m3 =s
is interesting to study the e:ect of a non-volatile bu:er so- G gas super9cial mass velocity, kg=m2 s
lution like phosphate bu:er. The numerical model enables He Henry’s law coeNcient, atm m3 =kmol
us to predict the absorption/desorption rates in such a com- I ionic strength, kmol=m3
plex system. In Fig. 13 the e:ect of two bicarbonate and k reaction rate constant
phosphate bu:er on the bulk pH pro9les along the column kg gas side mass transfer coeNcient, m/s
is shown. It is apparent from Fig. 13 that by using phos- kl liquid side mass transfer coeNcient, m/s
phate bu:er not only bu:er capacity can be increased but L liquid volume =ow rate, m3 =s
also the pH in the column can be maintained in the proper L liquid super9cial mass velocity, kg=m2 s
range. N =ux of component A per unit gas–liquid interfacial
area, kmol=m2 s
4.2.6. E=ect of SO2 gas concentration NA;g i interfacial =ux of component A per unit gas–liquid
In power plants the SO2 content of the generated =ue interfacial area, kmol=m2 s
gas may vary strongly depending on the type of coal be- pA partial pressure of the component A, atm
ing burned. The e:ect of di:erent SO2 concentrations was P total pressure, atm
evaluated and the results are shown in Fig. 14. By increas- R gas constant, m3 atm=kmol K
3600 S. Ebrahimi et al. / Chemical Engineering Science 58 (2003) 3589 – 3600

RA reaction rate of the component A, kmol=m3 s Chang, C. S., & Rochelle, G. T. (1985). SO2 absorption into NaOH and
S column cross-section, m2 Na2 SO3 aqueous solutions. Industrial and Engineering Chemistry,
T temperature, K Fundamental, 24(1), 7–11.
Danckwerts, P. V. (1970). Gas–liquid reactions. New York: McGraw-Hill.
x spatial coordinate in the liquid 9lm, m
Ebrahimi, S., Kleerebezem, R., van Loosdrecht, M. C. M., & Heijnen,
z spatial coordinate on the column height, m J. J. (2003). Kinetics of the reactive absorption of hydrogen sul9de
zA electric charge of species A into aqueous ferric sulfate solutions. Chemical Engineering Science,
58(2), 417–427.
Eden, D., & Luckas, M. (1998). A heat and mass transfer model for the
Greek letters
simulation of the wet limestone =ue gas scrubbing process. Chemical
Engineering & Technology, 22(1), 56–60.
 liquid 9lm thickness, m Edward, T. J., Maurer, G., & Prausnitz, J. M. (1978). Vapour liquid
l liquid hold-up m3 =m3 equilibria in multicomponent aqueous solutions of volatile weak
 viscosity of the solution, kg/m s electrolytes. A.I.Ch.E. Journal, 24(6), 966–975.
Gerard, P., Segantini, G., & Vanderschuren, J. (1996). Modeling of
" density, kg=m3
dilute sulfur dioxide absorption into calcium sul9te slurries. Chemical
% surface tension of liquid, N/m Engineering Science, 51(12), 3349–3358.
%c critical surface tension of packing material, N/m Hikita, H., Asai, S., & Tsuji, T. (1977). Absorption of sulfur dioxide
into aqueous sodium hydroxide and sodium sul9te solutions. A.I.Ch.E.
Journal, 23(4), 538–544.
Superscripts Hikita, H., & Konishi, K. (1983). The absorption of SO2 into aqueous
Na2 CO3 solutions accompanied by the desorption of CO2 . Chemical
g in gas phase Engineering Journal, 27(3), 167–176.
l in liquid phase Karlsson, C. B. a. H. T. (1997). Modeling the absorption of SO2 in
a spray scrubber using the penetration theory. Chemical Engineering
Science, 52(18), 3085–3099.
Subscripts Kenig, E. Y., Schneider, R., & GLorak, A. (1999). Rigorous dynamic
modelling of complex reactive absorption processes. Chemical
b in the bulk of the gas or liquid phase Engineering Science, 54(21), 5195–5203.
i at gas–liquid interface McCarthy, J. E. (1980). Flue gas desulfurization: Scrubber types and
selection criteria. Chemical Engineering Progress, 76(5), 58–62.
Onda, K., Sada, E., & Okumoto, Y. (1968a). Mass transfer coeNcients
between gas and liquid phases in packed columns. Journal of Chemical
Acknowledgements Engineering of Japan, 1(1), 62–66.
Onda, K., Takahashi, M., & Okumoto, Y. (1968b). Mass transfer
The support of this project by STW, the Dutch Technol- coeNcients between gas and liquid phases in packed columns. Journal
ogy Foundation and the support of the 9rst author by Iranian of Chemical Engineering of Japan, 1(1), 56–62.
Parkhurst, D. L., & Appelo, C. A. J. (1999). User’s guide to PHREEQC
Ministry of Since, Research and Technology are gratefully (Version 2). U.S. Geological Survey, Denver, CO.
acknowledged. Reid, R. C., Prausnitz, J. M., & Poling, B. E. (1988). The properties of
gases and liquids. New York: McGraw-Hill.
Sada, E., Kumazawa, H., & Butt, M. A. (1980). Absorption of sulfur
References dioxide into aqueous slurries of sparingly soluble 9ne particles.
Chemical Engineering Science, 35, 771–777.
Astarita, G., Savage, D. W., & Bisio, A. (1983). Gas treating with Vanysek, P. (2001). CRC handbook of chemistry and physics (82nd ed.)
chemical solvents. New York: Wiley. (pp. 5 –95 and 6 –194). Boca Raton: CRC Press LLC.
Brogren, C., & Karlsson, H. T. (1997). Modeling the absorption of SO2 in Xia, J., Rumpf, B., & Maurer, G. (1999). Solubility of sulfur dioxide in
a spray scrubber using the penetration theory. Chemical Engineering aqueous solutions of acetic acid, sodium acetate, and ammonium acetate
& Technology, 52(18), 3085–3099. in the temperature range from 313 to 393 K at pressures up to 3:3 MPa:
Chang, C. S., & Rochelle, G. T. (1981). SO2 absorption into aqueous Experimental results and comparison with correlations/predictions.
solutions. A.I.Ch.E. Journal, 27(2), 292–297. Industrial & Engineering Chemistry Research, 38(3), 1149–1158.

Вам также может понравиться