Vous êtes sur la page 1sur 8

1310 Ind. Eng. Chem. Res.

2008, 47, 1310-1317

Dynamic Measurement of Carbon Dioxide Volumetric Mass Transfer Coefficient


in a Well-Mixed Reactor Using a pH Probe: Analysis of the Salt and
Supersaturation Effects
M. Kordač and V. Linek*
Department of Chemical Engineering, Institute of Chemical Technology, Prague, Technická 5,
166 28 Prague 6, Czech Republic

This paper presents a new method for the evaluation of volumetric mass transfer coefficients (kLa) using a
pH probe response recorded during the absorption of carbon dioxide in a well-mixed reactor. For this method
of evaluation, it is not necessary to know the reaction equilibrium constant, the experiment start time, or the
initial and final steady-state pH probe readings. The experimental procedure is based on Hill’s recent publication
(Ind. Eng. Chem. Res. 2006, 45, 5796). Hill reported a decrease in kLa following the addition of salt during
carbon dioxide absorption in a well-mixed reactor and explained it as the result of ionic charge effects reducing
the ability of carbon dioxide molecules to diffuse away from the surface. Generally, oxygen absorption from
air is preferred because the solubility of oxygen is low and, thus, so will be the gas-phase depletion. In such
experiments, the mass transfer coefficients are less affected by gas-phase residence time distribution and by
bubble-size distribution. The solubility of carbon dioxide is 26 times higher than that of oxygen, which can
lead to significant gas concentration changes. Thus, the kLa values, measured using absorption of diluted
carbon dioxide, are more likely to be distorted by driving force errors caused by the use of an inappropriate
gas-mixing model. Using a stirred cell, the interfacial area of which is known, the mass transfer coefficients
of oxygen and carbon dioxide in water and in a salt solution were compared. The mass transfer coefficients
obtained for oxygen agreed with, or were slightly superior to, the coefficients obtained for carbon dioxide,
which corresponds with the lower molecular diffusivity of CO2. The mass transfer coefficients of carbon
dioxide obtained in salt solution were not significantly lower than those obtained in pure water, which is in
strong agreement with the literature but contradicts the results reported by Hill. It is shown that Hill’s findings
may be the result of his use of both an inaccurate gas-mixing model (no depletion of gas) and an imprecise
reaction term in his equations.

Introduction serious underestimation of kLa values in the case of carbon


dioxide absorption, (ii) to propose a new method that allows
The steady increase in carbon dioxide concentration in the
kLa to be calculated without knowledge of the experiment start
atmosphere, and the belief that it is a significant contributing
time or of the initial and final steady-state readings of the pH
factor to global warming,1 has recently led to attempts (e.g.,
probe (knowledge of which are essential to Hill’s method), and
refs 1 and 2) to develop photobioreactors that use atmospheric
(iii) to critically reassess the effect of salt on the absorption
carbon dioxide as a source of carbon for cell cultures. Because
rate of carbon dioxide presented by Hill.
the solubility of carbon dioxide, compared with other substrates,
is not high enough to ensure an adequate carbon supply at the
start of fermentation, carbon dioxide must be continuously fed Theory of the Method
into a fermentation broth by gas absorption. The rate of carbon Carbon dioxide is dissolved in the aqueous phase as four
dioxide absorption is a limiting factor in the process, and thus, different compounds (carbon dioxide, CO2; carbonic acid, H2-
the volumetric mass transfer coefficient is an important param- CO3; bicarbonate ion, HCO3-; and carbonate ion, CO3)), the
eter in photobioreactor design. Typically, engineers estimate the equilibrium concentrations of which are pH-dependent. Assum-
mass transfer rate of carbon dioxide using oxygen mass transfer ing equilibrium reactions to be instantaneous, Hill deduced how
coefficients that factor in a different diffusion coefficient of to recalculate a pH profile into a profile of dissolved carbon
carbon dioxide. Recently, this procedure has been questioned dioxide and, therefore, how to evaluate the volumetric mass
by Hill,3 who reported notably lower carbon dioxide mass transfer coefficients of carbon dioxide.
transfer coefficients compared to those of oxygen. The dissociation of carbon dioxide in water can be described
Hill3 used a dynamic method to measure carbon dioxide mass using the following relation set
transfer coefficients in a well-mixed reactor, with the coefficients
being evaluated from pH profiles of the aqueous phase after K0 K1
the introduction of bubbles containing carbon dioxide. In this H2O + CO2 798 H2CO3 798 H+ + HCO3- (1)
procedure, Hill assumed that the gas-phase concentration of
carbon dioxide in the vessel was constant and equal to the inlet From the value of the dissociation constant K0 (7 × 10-7 M,
concentration. Danckwerts4), it follows that the concentration of dissolved CO2
The aims of this paper are as follows: (i) to demonstrate greatly exceeds that of H2CO3 and, thus, that the dissolved CO2
that the use of an incorrect absorption driving force results in and H2CO3, the neutral compounds, may be considered as one
component, which, following Hill’s notation, is called carbonic
* Corresponding author e-mail: linekv@vscht.cz. acid, H2CO3, in the following text. Danckwerts and Sharma5
10.1021/ie0711776 CCC: $40.75 © 2008 American Chemical Society
Published on Web 01/08/2008
Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008 1311

provided the following expression for calculation of the solubil- the interface. Assuming both an ideal mixing of the gas phase
ity of carbon dioxide and a negligible mass transfer resistance in the gas phase,
(cH2CO3)/2 is in equilibrium with the carbon dioxide concentra-
(cH2CO3)/ ) pCO2/H ) pCO2 × 10-(5.3-1140/T) (2) tion in the gas leaving the vessel. The latter assumption holds
for the absorption of sparingly soluble gases, such as oxygen
where H is Henry’s coefficient, pCO2 is the partial pressure of and carbon dioxide, but the former assumption is questionable
carbon dioxide above the aqueous phase, and (cH2CO3)* is the at some situations when absorption takes place in a mechanically
concentration of physically dissolved CO2 and H2CO3 in agitated gas-liquid contactor. A detailed study7 on the correct
equilibrium with pCO2 in the gas phase. The same authors5 also use of the dynamic method for the determination of kLa in
provided the following expression for the equilibrium constant aerated agitated vessels has shown that these assumptions hold
of the second reaction for the measurement of low values of kLa (<∼200 h-1), which
was the case in Hill’s experiments. Furthermore, the same
cH+cHCO3- authors have shown7 that, because it avoids mixing interchanged
K1 ) ) 10-(3404.7/T+0.03279T-14.832) (3) gases in a reaction vessel, a gas interchange start-up technique
cH2CO3
considerably broadens the applicability of the dynamic method,
By analyzing equilibrium reactions, Hill3 deduced that the when used with an ideal mixing of the gas phase. The start-up
concentration of carbonate ion is insignificant and that, as the technique is as follows: after saturating the batch with a gas
solution becomes acidic (pH < 6), the concentration of (e.g., air or nitrogen), stirring and the gas supply are interrupted
hydroxide ion may also be ignored. In this case, the principle for the time necessary for the bubbles to escape from the batch;
of electroneutrality requires that the concentrations of hydrogen next, the stirring is restarted simultaneously with the introduction
and bicarbonate ion be equal of another gas containing the solute component (carbon dioxide).
In the subsequent evaluation, it is necessary to be aware of the
cH+ ) cHCO3- (4) fact that the instantaneous value of an interfacial area of gas
bubbles in a dispersion increases linearly over the start-up period
Danckwerts and Sharma5 also provided the following rate θ. The start-up period ends when gas holdup Hu reaches a steady
equation of the hydrolysis (eq 1) state at time θ ) VLHu/[QG(1 - Hu)].

( )
The bicarbonate concentration in relation 7 can be replaced
dcH2CO3 cH+cHCO3- using the differentiated eq 3 [dcHCO3- ) d(K1cH2CO3)0.5 )
B1 cH2CO3 -
) -k 0.5(K1)0.5(cH2CO3)-0.5 dcH2CO3] leads to
dt K1 (5)

[ ]
log B
k 1 ) 329.85 - 110.54 log T - 17265/T dcH2CO3 2
) kLa((cH2CO3)/2 - cH2CO3) (8)
For absorption from gas bubbles in dispersion, the lowest value
dt 2 + (K1/cH2CO3)0.5
estimated for the physical mass transfer coefficient koL is 2 ×
10-4 m/s (e.g., Calderbank and Moo-Young6). Using eq 5, Assuming a well-mixed gas phase, the molar fraction of carbon
together with the relation given in ref 4, the enhancement factor dioxide in the gas leaving the reactor, y2, can be deduced from
of absorption accompanied by a first-order irreversible chemical the mass balance of carbon dioxide. If changes in the rate of
reaction (eq 1) at 27.5 °C can be estimated as volumetric gas flow through the reactor, and CO2 accumulation
in the gas holdup, are both omitted, the balance has the following

E)
x(k ) o 2
L +B
k 1DCO2
)
form

koL p Q GH
QG (y1 - y2) ) ((cH2CO3)/1 - (cH2CO3)/2) )
x(2 × 10 -4 2
) + 0.0311 × 2 10 -9
) 1.000777 (6)
RT RT
kLa((cH2CO3)/2 - cH2CO3)VL (9)
2 × 10-4
This value is on the conservative side compared to the factor which can be rewritten as
for the reversible version of the same reaction. From this value
of E, it follows that the absorption of carbon dioxide in water A 1
(cH2CO3)/2 ) (c )/ + c
is not enhanced by a reversible reaction (eq 1). 1 + A H2CO3 1 1 + A H2CO3
Model of Instantaneous Reversible Reaction (Eq 1) Taking
Place in Bulk Liquid. In this model, it is assumed that carbon Q GH
dioxide is physically absorbed and, in bulk liquid, undergoes a A) (10)
kLaVLRT
reversible reaction (eq 1) that is sufficiently fast to keep the
bulk concentrations of carbonic acid, bicarbonate ion, and
By substituting (cH2CO3)/2 from eq 10 into eq 8, we obtain the
hydrogen ion in equilibrium. Assuming equilibrium (given by
following differential equation for describing the dynamic
reaction 1) is reached in bulk liquid, the differential balance of
response of the concentration of carbon dioxide in the liquid
carbon dioxide in the liquid phase during gas absorption is as
phase
follows
dcH2CO3 dcHCO3- dcH2CO3 A
kLa((cH2CO3)/2 - cH2CO3) ) + (7) ) kLa ((c )/ - cH2CO3) ×
dt 1 + A H2CO3 1

[ ]
dt dt
2
(11)
where (cH2CO3)/2 is the concentration of carbonic acid (i.e., the 2 + (K1/cH2CO3)0.5
sum of physically dissolved CO2 and H2CO3) in the liquid at
1312 Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008

Model of Moderately Rate Reversible Reaction (Eq 1)


Taking Place in Bulk Liquid. In this model, it is assumed that
carbon dioxide is physically absorbed and undergoes a reversible
reaction (eq 1) in bulk liquid, the reaction rate of which is given
by eq 5. In this case, the balances have the following forms

dcH2CO3
) kLa((cH2CO3)/2 - cH2CO3) -

( ) ( )
dt
(cHCO3-)2 dcHCO3- (cHCO3-)2
k1 cH2CO3 - ) k1 cH2CO3 - (12)
K1 dt K1

where (cH2CO3)/2 is given by eq 10 for a well-mixed gas phase in


a vessel. Figure 1. Concentration profiles obtained by integration of eqs 10 and
Model of Absorption Presented by Hill.3 To factor in the 11. Conditions: (cH2CO3)initial ) 1.38 × 10-5 M; (cH2CO3)/1 ) 2.81 × 10-3
M; K1 ) 8.15 × 10-7 M-1; QG ) 1.1 L/min; VL ) 2.45 L; HCO2 ) 32.09
changes in carbonic acid concentration due to reaction system atm/M; T ) 300.5 K.
1, Hill, instead of using the differential relation dcHCO3- )
d(K1cH2CO3)0.5, used a simple relation between the concentrations tions of CO2 gas in the vessel are only 3.8% and 6.9%,
of bicarbonic ion and carbonic acid cHCO3- ) (K1/cH2CO3)0.5cH2CO3, respectively. However, in eq 13, Hill assumed that the concen-
which follows from eq 3 and is valid for steady-state concentra- tration of carbon dioxide in the gas phase was essentially
tions only. He presented the mass balance of dissolved carbon constant, equaling 10% during the whole experiment. Such an
dioxide in the following form overestimation of the absorption driving force leads to serious
underestimation of kLa.

[ ]
dcH2CO3 1
) kLa((cH2CO3)/2 - cH2CO3) (13)
dt 1 + (K1/cH2CO3)0.5 Evaluation of pH Profiles
Hill reported problems in the evaluation of the pH profile
Comparison of Relations 11, 12, and 13 generated by each run of his experiment. To solve these
problems, caused by time delays after gas switching, Hill used
Equations 11, 12, and 13 were integrated to produce the an empirical equation to estimate both “the time to the midpoint
concentration profiles compared in Figure 1. Integration was of the rise in concentration” and “the span of the rising curve”,
performed by applying the conditions presented in Figure 2 of thus allowing him to determine the experiment start time. To
Hill’s paper: (cH2CO3)initial ) 1.38 × 10-5 M; (cH2CO3)/1 ) 2.81 avoid using this unnecessary procedure, another method of
× 10-3 M; K1 ) 8.15 × 10-7 M-1; QG ) 1.1 L/min; VL ) evaluation is proposed here, in which it is not necessary to know
2.45 L; HCO2 ) 32.09 atm/M; and T ) 300.5 K. The the initial and final steady-state readings of the pH probe nor
concentration profiles obtained from eqs 11 and 12 did not differ, the start time.
their agreement validating Hill’s assumption that reaction In order to solve eq 11 analytically, it is necessary to analyze
equilibrium is instantaneously achieved in bulk liquid. However, the reaction term in square brackets. In Figure 3, this term is
the profiles obtained from eqs 11 and 13 differed significantly; plotted as a function of the equilibrium partial pressure of carbon
the higher the kLa value, the greater the difference. This was dioxide. The equilibrium constant K1 (given in ref 9) is as
due to the following: (i) the different reaction terms used for follows,
the conversion of carbonic acid into bicarbonate ion (compare
the terms in square brackets in eqs 11 and 13) and (ii) the use pK1 ) 3670.7/T - 62.008 + 9.7944 ln T -
of an incorrect driving force for absorption. A well-mixed gas
phase is assumed in models 11 and 12, and thus the outlet gas 0.0118S + 0.000116S2 (14)
concentration is used, but Hill assumed that the concentration
where salt concentration is defined by salinity S in units of g/kg.
of carbon dioxide in the gas phase is constant during its
From Figure 3, it follows that the reaction term is approximately
absorption and, consequently, used the inlet gas concentration
constant and equals 0.99 for pH < 5. Therefore, because pH is
to determine the driving force in eq 13.
usually <5 during the main part of carbon dioxide absorption
Because the solubility of carbon dioxide is relatively large,
experiments, the following approximation can be used in the
changes in the concentration of carbon dioxide in the gas phase
evaluation of absorption experiments conducted in both water
due to absorption must be taken into account. This fact is
and salt solution
illustrated in Figure 2, in which outlet gas concentrations, for

[ ]
the absorption of both 10% CO2 and oxygen (from air) into
2
water, are plotted as a function of kLa, as well of the saturation ≈ 0.99 (15)
(R) of the liquid phase with the solute gas. Saturation R is 2 + K1/(cH2CO3)0.5
defined as the ratio of dissolved gas concentration to saturated
solution (e.g., for carbon dioxide, R ) cH2CO3/(cH2CO3)/1). Pre- The differential eq 11 then can be reduced to
cisely because the solubility of carbon dioxide is 26 times higher
than that of oxygen (ref 8: HO2 ) 819 atm/M), the reduction dcH2CO3 A
for CO2 absorption remains large, even when the decrease in ) kLa 0.99((cH2CO3)/1 - cH2CO3) )
dt 1+A
gas-phase concentration for air absorption is insignificant. For
example, Figure 2 shows that, for kLa ) 60 h-1, the y2/y1 ratio B((cH2CO3)/1 - cH2CO3) (16)
is 0.37 at the start of the dynamic experiment (R ) 0) and equals
0.68 at a saturation of R ) 0.5. Correspondingly, the concentra- where
Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008 1313

A caused by the fact that the diffusion coefficient of oxygen in


B ) k La 0.99
1+A water being 20% higher that that of carbon dioxide. However,

[ ]
-1 (17) this does not fully explain a difference in kLa of ∼40% (relative
0.99 VLRT to the maximum value obtained by Robinson and Wilke10) at
kL a ) -
B QGH (kLa)(13) 120 h-1. On the other hand, 40% corresponds well with
the difference between Hill’s model (eq 13) and the model
Integration of eq 16 with the initial condition cH2CO3 (0) ) presented in this work (eq 11). We can use these equations to
(cH2CO3)0 leads to the following time profile for the concentration estimate the kLa ratio as follows,
of H2CO3
2A 1 + (K1/cH2CO3)
0.5
(kLa)(13)
cH2CO3(t) ) (cH2CO3)/1 - [(cH2CO3)/1 - (cH2CO3)0] e-Bt (18) ) (23)
(kLa)(11) 1 + A 2 + (K /c )0.5
1 H2CO3
Using the following rearrangements,
In Figure 5, this ratio is plotted as a function of kLa for the
cH2CO3(t + ∆t) ) (cH2CO3)/1 - [(cH2CO3)/1 - (cH2CO3)0] e-B(t+∆t) experimental ranges described above. The experimental mean
value of cH2CO3 was calculated using [(cH2CO3)initial + (cH2CO3)/1]/2.
cH2CO3(t) - cH2CO3(t + ∆t) ) ∆cH2CO3 ) (19)
Other authors7,21-24 have reported lower kLa values in
e-Bt[(cH2CO3)/1 - (cH2CO3)0](e-B∆t - 1) noncoalescent batches at a significant depletion of solute
component in the gas phase, exactly the conditions described
we obtain by Hill for his experiments in salt solution. In all these cases,
it is likely that there was serious overestimation of the absorption
ln(K1∆cH2CO3) ) -Bt + const1 (20) driving force caused by the use of an incorrect gas-mixing
model. The increased likelihood of incorrectly estimating the
To determine the concentration of carbonic acid, we use the absorption driving force under such conditions is a result of
following simple inverse relationship between the concentrations the following process. Each bubble, according to its size and
of carbonic acid and hydrogen ion given by eqs 3 and 4 residence time in the batch, exhibits different gas-phase deple-
tion and, thus, a different absorption driving force. The
10-2pH differences are more pronounced in dispersions in solutions of
cH2CO3 ) (21) electrolytes and in viscous solutions, where the bubble coales-
K1
cence rate and, thus, the mutual mixing of their contents are
For K1∆cH2CO3, it can thus be deduced from eq 21 that hindered compared with dispersions in pure water. As a result,
estimates of the absorption driving force based on unstructured
K1∆cH2CO3 ) 10-2pH(t) - 10-2pH(t+∆t) (22) models (plug flow, perfect mixing) may differ from the real
situation by an order of magnitude. The consequent underesti-
For pH < 5, the volumetric mass transfer coefficient kLa can mation of kLa based on these models is serious even for
be calculated from B in eq 20. Note that B is dependent neither sparingly soluble gases, such as oxygen.7,21,23
on the initial nor final steady-state values of the pH probe signal
nor on the equilibrium constant K1. Integrating eq 11 using kLa Analysis of the Salt and Supersaturation Effects
) 110 h-1, the highest value measured by Hill, we obtained It is well-known7,10,11 that the addition of a structure-making
the concentration profile presented in Figure 4. Evaluating the electrolyte (such as NaCl) reduces the rate of bubble coales-
profile using ∆t ) 30 s, we obtained B ) 0.00733 s-1. Using cence, which, in turn, results in reduced bubble size and an
eq 17, the volumetric mass transfer coefficient was calculated increase in kLa values compared to those measured in pure
as water. However, while Hill visually observed bubble size to be

[ ]
-1 noticeably smaller when salt (2.85% solution NaCl in water)
0.99 VLRT
k La ) - ) was present, he actually reported a decrease in kLa values (e.g.,
B QGH from a mean value of 34 h-1 down to 28 h-1 at 375 rpm and

[ ]
-1
0.99 2.45 × 0.08205 × 300.5 an aeration rate of 1.1 L/min). Hill3 explained this as follows:
- )
0.00728 0.018333 × 32.09 “the ionic charge effect at the bubble surface may have reduced
the ability of carbon dioxide molecules to diffuse away from
0.03002 s-1 ) 108 h-1 the increased bubble surface area, resulting in no net change in
This agreement shows that the simplifications used have a the overall volumetric mass transfer coefficient”.
negligible effect on the resulting kLa value. Because the highest To reach an unambiguous conclusion about the effect of salt
value measured by Hill was used in this example, it can be on the mass transfer coefficient of carbon dioxide, we measured
expected that this small difference may be valid in the whole the mass transfer coefficients kL of oxygen and carbon dioxide
range of kLa values. in water and in 2.85% NaCl solution in a stirred cell with a flat
gas-liquid interface (interfacial area and absorption driving
force well-defined). The cell was a cylindrical vessel (inner
Comparison of Hill’s Data with Literature Data
diameter ) 0.117 m) equipped with six baffles, together with,
Hill3 compared his kLa data with those of Robinson and on a common shaft, a vaned disc impeller (diameter ) 0.06 m)
Wilke,10 who used a similar vessel to measure oxygen mass in the liquid phase and a two-paddle agitator in the gas phase.
transfer in both water and saltwater. While both studies used The impeller shape enabled the liquid to be agitated without
similar mixing and aeration rates, Hill’s kLa data ranged from appreciable surface wave motion. The surface remained flat
20 to 120 h-1, whereas those of Robinson and Wilke ranged across the whole range of agitator frequencies used in our
from 40 to 200 h-1. Hill has explained that this difference is experiments (100, 200, and 300 min-1). The temperature was
1314 Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008

Figure 2. Molar fraction of solute gas component in outlet gas. Conditions: (cH2CO3)initial ) 1.38 × 10-5 M; (cH2CO3)/1 ) 2.81 × 10-3 M; K1 ) 8.15 × 10-7
M-1; QG ) 1.1 L/min; VL ) 2.45 L; HCO2 ) 32.09 atm/M; T ) 300.5 K.

Figure 3. Reaction term as a function of equilibrium partial pressure of carbon dioxide. Conditions: K1 from eq 14; T ) 300.5 K.

Figure 4. Concentration and pH profiles obtained by integration of eq 11


for kLa ) 110 h-1 and evaluated using eq 20. Conditions: (cH2CO3)initial )
1.38 × 10-5 M; (cH2CO3)/1 ) 2.81 × 10-3 M; K1 ) 8.15 × 10-7 M-1; QG
) 1.1 L/min; VL ) 2.45 L; HCO2 ) 32.09 atm/M; T ) 300.5 K.
Figure 5. Comparison of relative decrease in kLa values obtained using
maintained at 20 °C by circulating liquid through the cell jacket. Hill’s model (eq 13) and our model (eq 11), as follows from eq 23.
At the start of each experiment, the vessel was filled with 1/2 L Conditions: K1 ) 8.15 × 10-7 M-1; VL ) 2.45 L; HCO2 ) 32.09 atm/M;
T ) 300.5 K.
of liquid (i.e., specific liquid surface a ) 19.5 m-1). Experiments
were carried out batchwise in the liquid phase and continuously
in the gas phase. Pure oxygen, carbon dioxide, and nitrogen, the batch. Next, the oxygen atmosphere in the cell was quickly
all taken from the cylinders, were used as the gas phase. A flushed with nitrogen. After the interchanges, (O2 f N2)
membrane-covered polarographic probe of our own construction followed by (N2 f O2), oxygen desorption and, subsequently,
(described in detail by Fujasova et al.12) and the glass pH probe absorption were monitored. Then, each experiment was repeated
(Sentix 41, WTW) were used for monitoring solute gas using carbon dioxide instead of oxygen. In all experiments, a
concentration in the liquid batch. gas flow rate of 100 mL/min was fed above the liquid surface
Mass transfer coefficients were evaluated from concentration to ensure an insignificant concentration of the solute in the gas
profiles of oxygen and carbon dioxide in the liquid batch during above the agitated liquid. This implies that the value of A tends
absorption/desorption of the gas into/out of the batch. At the to infinity, and thus, A/(1 + A) ) 1 in these experiments. In
beginning of each experiment, pure oxygen saturated with water the case of the linear dependence of probe signal G on the
vapor was fed above the liquid surface until the gas saturated oxygen concentration [G ) rcO2 + q; r and q are arbitrary
Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008 1315

Figure 6. Oxygen probe responses measured for oxygen absorption/desorption in stirred cell and evaluated according to eq 24. Batch: water, 20 °C.

Figure 7. pH probe responses measured for carbon dioxide absorption/desorption in stirred cell and evaluated using eq 20 for pH < 5. Batch: 2.85% NaCl
solution, 20 °C.

constants], oxygen absorption can be evaluated by reducing eq Table 1. Mass Transfer Coefficients of Oxygen and Carbon Dioxide
20 to in Water and in 2.85% Salt Solution Measured in a Stirred Cell at
20 °C; Values in Italic Were Measured in Conditions under Which
Supersaturation May Occur
ln(∆G) ) -kLa‚t + const2 (24)
kL × 105, m/s
Probe responses (pH and oxygen) evaluated using relations Agitator,
100 200 300 100 200 300
frequency,
20 and 24 are presented in Figure 6 for water and in Figure 7 rpm abs. des. abs. des. abs. des. ave.a ave.a ave.a
for 2.85% NaCl solution. ∆G represents the difference between
water, O2 2.09 2.14 3.57 3.48 5.67 5.39 2.18 3.50 5.29
oxygen probe readings measured at times t and t + ∆t (∆t ) 2.25 2.21 3.60 3.51 5.41 5.18
144 s). Using eq 22, K1∆cH2CO3 was calculated from pH probe water, CO2 4.63 4.65 2.06 3.21 4.69
readings measured at times t and t + ∆t (∆t ) 100 s). Table 1 2.15 1.93 3.27 3.44 3.95b 4.40
presents the mass transfer coefficients evaluated from the probe 1.97 2.06 3.14 3.37 4.75 4.87
responses during absorption and desorption. It also displays the kL,CO2 calculated from eq 25 1.98 3.19 4.81
2.85% NaCl, 2.22 2.15 3.57 3.48 5.49 5.31 2.12 3.45 5.21
kL values of carbon dioxide obtained by the recalculation of O2 2.18 2.09 3.49 3.41 5.24 5.11
oxygen mass transfer coefficients using the following ratio of 2.85% NaCl, 1.77 2.62 3.52 4.53 5.48 6.71 1.89 3.36 5.44
diffusion coefficients of the gases in water [ref 19 for O2, ref CO2 2.03 2.57 3.44 4.38 5.55 6.15
20 for CO2] 1.87 2.10 3.13 4.25 5.28 6.26
a Average values. b Not considered in the average value.
kL,CO2 ) kL,O2(DCO2/DO2)1/2 )
Our results show that the kLa of oxygen absorption (N2 f
kL,O2(1.90 × 10-9/2.30 × 10-9)1/2 ) 0.91kL,O2 (25) O2) and carbon dioxide desorption (CO2 f N2) (data in italic
in Table 1) are systematically higher (by 4% for oxygen and
Distortions of the responses, caused by the dynamics of the by up to 25% for carbon dioxide in the salt solutions) than the
pH probe, had negligible impact on the evaluation as, in both corresponding desorption (O2 f N2) and absorption (N2 f CO2)
this and Hill’s work, the time constant of the electrode was much values. Some authors13-17 have reported that, all other experi-
larger (see Figure 8, k > 3500 h-1) than the kLa (<110 h-1). mental conditions being equal, kLa measurement methods
1316 Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008

carbon dioxide molecules to diffuse away from the bubble


surface. We therefore conclude that the lower kLa values
reported by Hill in salt solution are an artifact caused by his
use of an incorrect model for calculating the absorption driving
force.

Conclusions
Relation 13, derived by Hill3 for the rate of carbon dioxide
absorption in water in a well-mixed reactor, gives substantially
lower kLa values than those evaluated using relation 11 in this
paper. We have shown that this is due to Hill’s use of an
incorrect absorption driving force in the model (eq 13), which
itself is a result of the omission of the concentration changes
induced by the absorption of highly soluble gases, such as
carbon dioxide.
Figure 8. Responses of pH probe measured for pH step changes: pH A new method for the evaluation of kLa, based on pH
differences are measured for time interval ∆t ) 1.03 s.
probe responses to the absorption of carbon dioxide, was
involving the temporary or permanent supersaturation of liquid presented. The advantage of this method is that it is not
by dissolved gases produce higher kLa values than methods not necessary to know the reaction equilibrium constant, the
involving supersaturation. The supersaturation of liquid by gas experiment start time, or the initial and final steady-state pH
occurs if it holds that18 probe readings.
By comparing the mass transfer coefficients of carbon dioxide
∑j pL j
in water and in a salt solution in a stirred cell, we have shown
that Hill’s finding that the addition of salt (NaCl) reduces kLa
s) -1>0 (26) values compared to those in pure water is not caused by ionic
pt charge effects at the bubble surface reducing the ability of
carbon dioxide molecules to diffuse away from the surface but
i.e., the combined content of dissolved gases in the liquid by the incorrect absorption driving force used by Hill to evaluate
is higher than its complete saturation. Where simultaneous kLa. Our experiments have shown that volumetric mass transfer
physical absorption-desorption occurs, supersaturation may coefficients evaluated from responses occurring after CO2 f
occur at the constant total pressure of the system pt as a result N2 interchanges are enhanced by the effect of liquid supersatu-
of the different component diffusivities. During an (N2 f O2) ration.
interchange, oxygen has a higher diffusivity in the liquid
phase than nitrogen, whereas during a (CO2 f N2) inter-
change, nitrogen has a higher diffusivity in the liquid phase Acknowledgment
than carbon dioxide (ref 19: DO2water ) 2.30 × 10-9 m2/s, Support from the Ministry of Education (MSM 6046137306)
DN2water ) 2.0 × 10-9 m2/s; ref 20: DCO2water ) 1.90 × 10-9 and from the Grant Agency of Czech Republic through Project
m2/s, all at 20 °C). Oxygen, in the former case, and nitrogen, No. 104/05/P203 is gratefully acknowledged.
in the latter case, enter the liquid faster than nitrogen and
carbon dioxide, respectively, are able to leave it. During these Nomenclature
interchanges, supersaturation (s > 0) develops in the liquid
phase, and, where supersaturation is followed by spon- A ) QGH/(kLaVLRT)
taneous bubble nucleation, this can increase the mass transfer c ) concentration (mol/L)
coefficients for these processes in comparison with the reverse D ) coefficient of molecular diffusivity (m2/s)
ones. Bubble creation enhances mass transfer but depends E ) enhancement factor of absorption
on conditions that are not irreproducible, such as the presence G ) reading of oxygen probe (volt)
of trace contaminants that act as nuclei for new bubbles. H ) Henry’s coefficient ((atm L)/mol)
This explains why we found a more pronounced enhancement Hu ) steady-state gas holdup, volumetric fraction of gas
in electrolyte solutions than in pure water (see Table 1), a K0,1 ) equilibrium constant (mol/L)
finding also reported in the literature.14,15 Therefore, in k1 ) first-order kinetic constant (s-1)
what follows, we only consider the mass transfer coefficients koL ) mass transfer coefficient for physical absorption (m/s)
of oxygen and carbon dioxide for responses obtained from kLa ) volumetric mass transfer coefficient (h-1 or s-1)
gas interchanges unhampered by this almost irrepro-
pt ) barometric pressure (atm)
ducible supersaturation effect, namely, (O2 f N2) and (N2 f
pCO2 ) partial pressure of carbon dioxide (atm)
CO2).
A number of conclusions can be drawn from the results pLj ) equilibrium partial pressure of gas component j (atm)
presented in Table 1. The mass transfer coefficients obtained QG ) aeration rate (L/s)
for oxygen agree with, or are slightly superior to, the coef- R ) gas constant ((atm L)/(mol K))
ficients obtained for carbon dioxide, which corresponds with S ) salinity (g/kg)
the lower molecular diffusivity of CO2. The mass transfer s ) supersturation criterion defined by eq 24
coefficients of carbon dioxide obtained in salt solution T ) temperature (K)
were not significantly lower than those obtained in pure water, t ) time (s)
thereby contradicting Hill’s statement that the ionic charge VL ) liquid volume in reactor (L)
effect of NaCl at the bubble surface reduces the ability of y ) molar fraction
Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008 1317

R ) saturation of liquid phase with carbon dioxide [)cH2CO3/ dynamic pressure method in individual sections of the vessel. Chem. Eng.
(cH2CO3)/1] Sci. 2007, 62, 1650.
(13) Vivian, J. E.; King, C. J. The mechanism of liquid-phase resistance
θ ) start-up period [)VLHu/{QG(1 - Hu)}] (s) to gas absorption in a packed column. AIChE J. 1964, 10, 221.
(14) Hikita, H.; Konishi, Y. Desorption of carbon dioxide from
Literature Cited supersaturated water in an agitated vessel. AIChE J. 1984, 30, 945.
(15) Hikita, H.; Konishi, Y. Desorption of carbon dioxide from aqueous
(1) Sik, S. H.; Ryong, C. S. Continuous photo bioreactor for carbon electrolyte solutions supersaturated with carbon dioxide in an agitated vessel.
dioxide removal to inhibit global warming and mass-production of mi- AIChE J. 1986, 31, 697.
croalgae. Korean Patent No. KR 2005-081766. Korean Patent Appl. KR (16) Linek, V.; Havelka, P.; Sinkule, J. Supersaturation effect in steady-
2004-10141, 2004.02.16. state and dynamic methods for measuring kLa in gas-liquid dispersions.
(2) Chae, S. R.; Hwang, E. J.; Shin, H. S. Single cell protein production Chem. Eng. Sci. 1996, 51, 5223.
of Euglena gracilis and carbon dioxide fixation in an innovative photo- (17) Kordač, M.; Linek, V. Mechanism of enhanced gas absorption in
bioreactor. Biores. Technol. 2006, 97, 322. presence of fine solid particles. Effect of molecular diffusivity on mass
(3) Hill, G. A. Measurement of Overall Volumetric Mass Transfer transfer coefficient in stirred cell. Chem. Eng. Sci. 2006, 61, 7125.
Coefficients for Carbon Dioxide in a Well-Mixed Reactor Using a pH Probe.
(18) Brian, P. L. T.; Vivian, J. E.; Matiatos, D. C. A criterion for
Ind. Eng. Chem. Res. 2006, 45, 5796.
supersaturation in simultaneous gas absorption and desorption. Chem. Eng.
(4) Danckwerts, P.V. Gas Liquid Reactions; Harvard University Press:
Sci. 1967, 22, 7.
Cambridge, MA, 2003.
(5) Danckwerts, P. V.; Sharma, M. M. The Absorption of Carbon (19) Verhallen, P. T. H. M.; Omen, L. J. P.; Elsen, A. J. J. M.; Kruger,
Dioxide into Solutions of Alkali and Amines. Chem. Eng. 1966, 202, CE244. A. J.; Fortuin, J. M. The diffusion coefficients of helium, hydrogen, oxygen
(6) Calderbank, P. H.; Moo-Young, M. B. The Continuous Phase Heat and nitrogen in water determined from the permeability of a stagnant liquid
and Mass Transfer Properties of Dispersions. Chem. Eng. Sci. 1961, 16, layer in the quasi-state. Chem. Eng. Sci. 1964, 39, 1535.
39. (20) Tamimi, A.; Rinker, E. B.; Sandall, O. C. Diffusion Coefficients
(7) Linek, V.; Vacek, V.; Beneš, P. A. Critical Review and Experimental for Hydrogen Sulfide, Carbon Dioxide, and Nitrous Oxide in Water over
Verification of the Correct Use of the Dynamic Method for the Determi- the Temperature Range 293-368 K. J. Chem. Eng. Data 1994, 39, 330.
nation of Oxygen Transfer in Aerated Agitated Vessels to Water, Electrolyte (21) Midoux, N.; Laurent, A.; Charpentier, J. C. Limits of the chemical
Solutions and Viscous Liquids. Chem. Eng. J. 1987, 34, 11. method for the determination of physical mass transfer parameters in
(8) Linek, V.; Vacek, V. Chemical engineering use of catalyzed sulfite mechanically agitated gas-liquid contactors. AIChE J. 1980, 26, 157.
oxidation kinetics for the determination of mass transfer characteristics of (22) Heijnen, J. J.; Van’t Riet, K.; Wolthius, A. J. Influence of very
gas-liquid contactors. Chem. Eng. Sci. 1981, 36, 1747. small bubbles on the dynamic kLa measurement in viscous gas-liquid
(9) Mook, W.G.; deVries, J.J. EnVironmental isotopes in the hydrological systems. Biotechnol. Bioeng. 1980, 22, 1945.
cycle: Principles and applications. Volume 1: Introduction, theory, (23) Keitel, G.; Onken, U. Errors in the determination of mass transfer
methods, reView; IAEA Publication: Vienna, Austria, 2005; pp 143-166. in gas-liquid dispersions. Chem. Eng. Sci. 1981, 36, 1927.
(10) Robinson, C. W.; Wilke, C. R. Oxygen Absorption in Stirred (24) Cents, A. H. G.; de Bruijn, F. T.; Brilman, D. W. F.; Versteeg, G.
Tanks: A Correlation for Ionic Strength Effects. Biotechnol. Bioeng. 1973, F. Danckwerts-plot technique by simultaneous absorption of CO2 and
15, 755. physical desorption of O2. Chem. Eng. Sci. 2005, 60, 5809.
(11) Van’t Riet, K. Review of measuring methods and results on
nonviscous gas-liquid mass transfer in stirred tank. Ind. Eng. Chem. Process ReceiVed for reView August 30, 2007
Des. DeV. 1979, 18, 357. ReVised manuscript receiVed October 26, 2007
(12) Fujasova, M.; Linek, V.; Moucha, T. Mass transfer correlations Accepted October 30, 2007
for multiple-impeller gas-liquid contactors. Analysis of the effect of axial
dispersion in gas and liquid phase on “local” kLa values measured by the IE0711776

Vous aimerez peut-être aussi