You are on page 1of 7

Ind. Eng. Chem. Res.

1996, 35, 4559-4565 4559

Hydrodynamics and Mass Transfer in a Liquid-Impelled Loop


Reactor

D. M. R. Mateus, M. M. R. Fonseca, and S. S. Alves*,


Secc a o de Biotecnologia and Secc a o de Feno menos de Transfere ncia, Departamento de Engenharia Qumica,
Instituto Superior Te cnico, 1096-Lisboa-Codex, Portugal

The liquid-impelled loop reactor was studied with respect to hydrodynamics and liquid-liquid
mass transfer. The system involved a continuous aqueous phase, a dispersed organic phase
(n-dodecane), and sometimes a solid phase (gel beads). Indole was the transferred solute. Drop
size, dispersed phase holdup, and continuous phase circulation rate were measured. Experi-
mental overall volumetric mass transfer coefficients of indole were taken from an earlier paper.
In the absence of gel phase or up to about 10% gel concentration, circulation and holdup could
be reasonably well predicted from standard expressions for pressure balance and single-phase
friction pressure loss. This was best achieved by a widely used correlation proposed by
Richardson and Zaki. Use of the Zuber and Findlay two-phase drift-flux model was also
successful, with an adjusted parameter. Experimental mass transfer coefficients were compared
with calculations from existing correlations. The best agreement was obtained with correlations
for circulating drops, as expected by the range of drop Reynolds numbers covered in the
experiments.

1. Introduction n-dodecane and water. In neither of these studies was


comparison between experimental data and available
The liquid-impelled loop reactor, or LLR, is a rela- correlations attempted.
tively novel type of bioreactor (Tramper et al., 1987), In this work mass transfer is linked with hydrody-
analogous to the airlift in its hydrodynamics; the main namics to see how well available correlations perform
difference is that continuous liquid phase circulation, in predicting measured overall volumetric mass transfer
rather than being driven by air dispersed in the riser, coefficients.
is driven by a second immiscible liquid phase with a
different density (Figure 1). This kind of reactor may
2. Theory and Modelling
be advantageous if, for some reason, a second liquid
phase is necessary or convenient in the bioreactor, e.g., Drop Size. The mechanism of drop formation at an
when one of the reagents or products has low solubility orifice depends on the regime (Clift et al., 1978). At low
in the aqueous phase, where the reaction takes place. flow rates, there is individual drop formation, whereas
Another major advantage over stirred reactors is the at high flow rates, a jet is formed and drops break up
avoidance of high-shear regions. Examples of applica- from this jet. The single-drop regime occurs if (Meister
tions include the bioconversion of benzene to cis- and Scheele, 1969)
benzenoglycol (van den Tweel et al., 1987), the produc-

[ ( DN
)]
tion of antraquinones by plant cell cultures (Buitelaar 1/2

et al., 1990), the growing of bacteria in tetralin (Vermue vN e 1.73 1- (1)
FdDN d
and Tramper, 1990), and the biosynthesis of L-tryp-
tophan from indole and L-serine (Mateus, 1994). where vN is the average dispersed phase velocity through
Hydrodynamics of the LLR has, to the authors the orifice, d is the drop diameter as calculated for the
knowledge, been studied only by van Sonsbeek et al. single-drop regime, is the interfacial tension, Fd is the
(1990), who modeled it using the Zuber and Findlay two- density of the dispersed phase, and DN is the orifice
phase drift-flux model and a friction coefficient derived diameter.
from one-phase flow theory. The effect of adding a third To estimate drop size in the jet regime, Scheele and
solid phase, important when an immobilized biocatalyst Meister (1968) proposed
is at play, was not studied. Large solid phase holdups
may be necessary in this case to achieve the required d ) (1.5m)1/3DN2/3 (2)
productivity (Mateus et al., 1996).
Liquid-liquid mass transfer in the LLR was inves- where m is the wavelength of the growing disturbance
tigated by van Sonsbeek et al. (1992) and by Mateus et of the jet surface (Meister and Scheele, 1967). In the
al. (1996). The former determined overall oxygen single-drop regime, the following correlation has been
transfer coefficients between tap water and a dispersed proposed (Scheele and Meister, 1968)
phase consisting of an oxygen-enriched perfluorochemi-

[
cal, FC40. A steady state method was used. Mateus
DN 20QDN 4FdQvN
et al. (1996) used a dynamic method to measure overall vd ) F
gF 3gF
+ 2 - +
volumetric mass transfer coefficients of indole between d gF

( Q2DN2Fd
)]
1/3
* To whom correspondence should be addressed. Telephone, 4.5 (3)
+351-1-8417188; Fax, +351-1-8499242. (gF)2
Secc
ao de Biotecnologia.
Secc
ao de Fenomenos de Transferencia. vd is the drop volume after break off, F is the Harkins-
S0888-5885(96)00225-4 CCC: $12.00 1996 American Chemical Society
4560 Ind. Eng. Chem. Res., Vol. 35, No. 12, 1996

Figure 1. Liquid-impelled loop reactor: (a) safety valve, (b) and (c) sampling valves, (d) riser, (e) downcomer, (f) organic phase reservoir,
(g) discharge valve, (h) perforated plate (1 mm diameter holes), (i) recirculating pump.

Brown correction factor generally presented as a plot weight between the riser and downcomer liquid columns
of F vs DN (F/vd)1/3 (Scheele and Meister, 1968), g is the
acceleration of gravity, is the viscosity of the continu- Pdf ) FgH (7)
ous phase, Q is the volumetric flow rate of the dispersed
phase, and F is the density difference between con- where  is the dispersed phase holdup in the riser, F
tinuous and dispersed phases. For both nozzle and is the difference in densities between the continuous and
multinozzle distributors, Vedaiyan (1969) proposed an the dispersed phases, and H is the height of the two-
alternative correlation (in cgs units) to obtain the mean phase column (approximately the height of the riser).
drop size, d The various friction losses may be given, as a first
approximation for low holdups of the dispersed phase,

( )
vN by Darcys equation for single-phase pressure loss

d
(4)
2gDN
1/2
)R

( )
Fg Pi ) KiF
ui2
(8)
2
with R ) 1.592, ) -0.0665, and the following restric-
tion for nozzle size where F is the continuous phase density, ui is the
continuous phase velocity in section i, and Ki is a factor
which in general depends on the Reynolds number and
1 1/2 1/2
( )
2 Fg
< DN <
Fg ( ) (5) on the type of fitting. It may be given by

Leqi
Holdup and Circulation. Circulation results mainly Ki ) fi(Rei) (9)
Rhi
from a balance between driving force and drag at the
walls
where fi is the Fanning friction factor for straight pipe,
Rei, Rhi, and Leqi are respectively the Reynolds number,
Pdf ) i Pi (6) hydraulic radius, and equivalent length of section or
fitting i.
If large holdups and/or a third phase is present, the
where Pdf is the driving force and Pi are the friction situation becomes more complicated, as friction is now
pressure losses in the various reactor sections and a function also of the holdup and properties of the
fittings. Driving force is given by the difference in dispersed phase(s). Two types of model may then be
Ind. Eng. Chem. Res., Vol. 35, No. 12, 1996 4561
Table 1. Continuous Phase Mass Transfer Coefficients in Liquid-Liquid Systems
drop condition expression authors
(i) stagnant Shc ) 2.0 + 0.95Rec0.5Scc0.5 Garner and Suckling (1958)
(ii) stagnant
( ) dg1/3 Hughmark (1967)
0.072
Shc ) 2 + 0.463Rec0.484Scc0.339
Dc2/3
(iii) circulating Shc ) -126 + 1.8Rec0.5Scc0.42 Garner, Foord and Tayeban (1959)
(iv) circulating
[ ( ) ]dg1/3 Hughmark (1967)
0.072
0.484 0.339
Shc ) 2 + 0.463Rec Scc F
Dc2/3
with
F ) 0.281 + 1.615K + 3.73K2 - 1.874K3

()( )
c c ur
1/4 1/6
K ) Rec1/8
d g

used to estimate pressure drops (Holland and Bragg, overall mass transfer coefficient, while kd and kc are the
1995): homogeneous flow models and models for sepa- film coefficients for the dispersed and continuous phases,
rated flow such as the Lockhart and Martinelli model respectively.
(Lockhart and Martinelli, 1949). In the present case, In the general case where both phases control mass
where most of the pressure drop occurs in vertical pipes transfer, the dispersed phase coefficient kd depends on
with no apparent stratification of phases, simpler the magnitude of the continuous phase coefficient, kc,
homogeneous flow models are likely to be applicable. which may be estimated using the empirical correlations
Equations 6-9 relate the continuous phase velocity in Table 1.
to physical properties, geometric and operational vari- To calculate the dispersed phase coefficient, kd, for
ables, and the holdup of the dispersed phase. This may the general case where both phases control mass
be independently related with the slip velocity, us, and transfer, Jakob (1949) derived the following expression
the superficial velocities of the two phases in the riser, for stagnant drops

[ ( )]
uc for the continuous and ud for the dispersed, through
the theoretically derivable formula 4Ddt
d
kd ) -
6t
ln 6 Cn exp -n2 (15)
ud uc n)1 d2
us ) (10)
1-
-

where Dd is the solute diffusivity in the dispersed phase
For many two-phase systems, it has been found that and t is the contact time.
the slip velocity is related to the terminal velocity, ut, For circulating drops, Elzinga and Banchero (1959)
by (Richardson and Zaki, 1954) suggested that the correlation proposed by Kronig and
Brink (1950) for the case in which mass transfer is
us ) ut(1 - )n-1 (11) controlled in the dispersed phase alone be adapted to
the general situation where both phases control. The
where n is an empirical quantity, found to be around 2 Kronig and Brink equation is as follows

[ ( )]
for Re > 500.
d 3 t
An alternative is the two-phase drift-flux theory of
Zuber and Findlay (1965), who proposed the following
kd ) -
6t
ln Bn2 exp -64nDd
8n)1
(16)
d2
expression
Coefficients Bn, Cn, n, and n in eqs 15 and 16 are given,
ub ) c0(uc + ud) + ut (12) for the general situation where both phases control, in
Elzinga and Banchero (1959) as a function of kc.
where c0 may be taken as an empirically adjustable The referred equations for film coefficients have been
parameter and ub is the drop rise velocity. An ap- deduced for clean systems. Impurities, and particularly
proximate relationship between ub and the slip velocity tensioactive substances, may significantly increase or
has been proposed by Ayazi Shamlou et al. (1994) decrease the overall mass transfer kinetics. The phe-
nomena involved are complex, and to date, there is no
ub ) us(1 - ) (13) general treatment of the problem.
Equations 6-10 may be solved together with either
3. Experimental Section
eq 11 or eqs 12 and 13 to give both the velocity of the
continuous phase, uc, and the holdup, , for any given Reactor. The liquid-impelled loop reactor (LLR) is
value of the dispersed phase flow rate, as long as the shown in Figure 1. It consists basically of two glass
drop terminal velocity is known. This may be calculated pipes (riser and downcomer) connected at the top and
using, for example, the graphical correlation given by bottom. It is operated in a way similar to an airlift
Hu and Kintner (1955). reactor, except that the dispersed phase is a liquid
Mass Transfer Coefficient. The overall mass trans- instead of a gas. In the present case, an organic solvent
fer coefficient is given in terms of the partial mass (n-dodecane), with density lower than water, is continu-
transfer coefficients of the two phases by ously dispersed into the aqueous phase through the
perforated plate at the bottom of the riser. The dis-
1 1 m persed organic phase coalesces at the top of the reactor
(14)
K d kd kc
) +
and is recirculated. A sheet of polyurethane foam at
the disengagement section of the LLR helps coalescence
where m is the equilibrium coefficient and Kd is the and avoids the formation of emulsions (Fonseca et al.,
4562 Ind. Eng. Chem. Res., Vol. 35, No. 12, 1996

Table 2. Physical Properties of the Two Liquid Phases


Mutually Saturated with Indole at 37 C
interfacial tension
density viscosity
(mN m- 1)
(kg m- 3) (kg m- 1 s- 1)
n-dodecane 735 0.00163 44a 38b
phosphate buffer 1006 0.00075
(0.1 M, pH 8.0)
a Between 0 and 4 s. b For t f .

1994). The density difference between the riser and


downcomer induces the desired circulation of the con-
tinuous phase, which may include a solid third phase. Figure 2. Drop diameter as a function of flow rate of the dispersed
More details of the LLR and its operation can be found phase: 9 experimental values; s predicted drop diameter using
Scheele and Meister eq 3 for the single-drop regime; - predicted
in Mateus et al. (1994). drop diameter using Scheele and Meister eq 2 for the jet regime;
Physical Properties. Measured physical properties 0 predicted drop diameter using Vedayian eq 4 without adjusted
of the two liquids relevant to this work are presented parameters; - - - predicted drop diameter using Vedayian eq 4 with
in Table 2. The liquids were mutually saturated by adjusted parameters.
mixing and settling, prior to measurement. Viscosities
were determined with a Schott-Gerate micro-Ostwald
viscometer, densities were obtained using a pycnometer,
and interfacial tension was measured by the Harkins-
Brown drop volume technique.
Drop Size. Drop diameters were measured using the
photographic method. To avoid lens effect, photos were
taken through a water-filled transparent cubic chamber
with an engraved scale, which enveloped a section of
the riser. Sauter mean diameters were calculated from
3-4 photos for each experimental condition, each photo
containing 100-200 drop images, depending on operat-
ing conditions. The dispersed phase was dyed with red
oil to improve contrast. This dye is appropriate, since Figure 3. Terminal rise velocity of n-dodecane drops in phosphate
it has a minimal effect on interfacial tension (Hu and buffer (mutually saturated in indole), given by the Hu and Kintner
Kintner, 1955), as was experimentally confirmed. correlation (1955).
Holdup and Circulation. To determine the con-
tinuous phase velocity, a flow-follower technique was mutually saturated in indole, as a function of drop
used. The time a colored gel bead takes to cover a fixed diameter using the correlation of Hu and Kintner (1955).
distance in the downcomer was measured. This mea- Friction Pressure Loss Calculation. In order that
surement was repeated at least 15 times to take into holdup and circulation rate may be calculated, the
account any deviation from a flat velocity profile. The various friction losses must be accounted for (eqs 6, 8,
continuous phase velocity was taken as the difference and 9). Two situations were considered: with and
between the average bead velocity and the experimental without a third, solid phase, consisting of gel beads.
bead terminal velocity. Without solid phase, all pressure losses except those
Holdup was determined by an unconventional tracer in the riser are single phase. Even for the riser, since
injection technique, as follows. At some point in time, the organic phase holdup never exceeded 6%, pressure
the dispersed phase is suddenly dyed. From the time losses could be calculated assuming single phase; two-
the dyed front takes to reach the top of the riser, the phase homogeneous flow calculation gives very similar
known dispersed phase flow rate, and the riser volume, results. Six terms are considered in eq 6: two sections
the holdup can be easily calculated, as long as plug flow of straight pipe (riser and downcomer), two 90 elbows,
is assumed. A similar pulse experiment was enough to and two cross-sectional changes. Values of Ki for vari-
show that the effect of axial dispersion was indeed ous fittings may be found, e.g., in Perry and Chilton
negligible for this determination. (1973), while straight pipe friction factors for smooth
pipe were obtained using the Blasius equation
4. Results and Discussion 0.0781
f) (17)
Drop Size and Terminal Velocity. Both visual Re1/4
observation and the threshold estimated by eq 1 indi-
cated that the prevailing regime was single drop. With the gel phase present, on the other hand, the
Figure 2 compares the prediction of drop size using eqs situation becomes more complicated. Three phases are
3 and 4 with experimental measurements. The result now at play, the continuous aqueous phase and two
is reasonable without parameter adjustment. Adjust- dispersed phases, the organic and gel phases. The
ment of parameters in Vedaiyans equation, (eq 4), gives organic phase, while being the driving phase due to its
R ) 1.753 and ) -0.0787 and very good agreement density, has little effect on friction losses owing to its
with experiment. Thus, for the remainder of the work, small holdup. The gel phase, on the other hand, while
where experimental measurement of drop size was not not contributing to the circulation driving force, because
carried out, Vedaiyans equation with adjusted param- its density is that of water, does affect friction losses
eters was used instead. due to its large holdup. The contribution of gel holdup
Figure 3 presents the terminal rise velocity of n- to friction losses is difficult to calculate, because of the
dodecane drops in 0.1 M phosphate buffer, pH 8.0, complex reactor geometry; hence, experimental values
Ind. Eng. Chem. Res., Vol. 35, No. 12, 1996 4563

Figure 4. Overall friction coefficient as a function of gel holdup Figure 7. Experimental vs calculated volumetric mass transfer
(volume of gel/volume of aqueous phase). coefficients as a function of dispersed phase flow rate. 9 experi-
mental; - kd from eq 15, kc from eq i in Table 1; s kd from eq
15, kc from eq ii in Table 1; kd from eq 16, kc from eq iii in Table
1; - - - kd from eq 16, kc from eq iv in Table 1.

phase holdups with experimental values, for low gel


loads. The use of the two-phase drift-flux model of
Zuber and Findlay (1965), eqs 12 and 13, together with
eqs 6-10 only gives acceptable results for low values of
the adjusted parameter, i.e., c0 ) 0.5 (Figure 5). This
parameter has to do with the continuous phase velocity
profile: it is shown to be equal to 1 for a flat velocity
profile and increases as the velocity profile approaches
a parabola, in which case c0 ) 1.6. However, in the
literature, c0 has taken values as high as 5.0 or as low
Figure 5. Continuous phase velocity (9 experimental, s model) as 0.7, being considered more an empirically adjustable
and dispersed phase holdup in the riser (0 experimental, s model) parameter than a theoretical one (Ayazi Shamlou et al.,
as a function of dispersed phase flow rate. Model predictions were 1994).
made using eqs 6-10 plus the Richardson and Zaki eq 11. Good agreement, on the other hand, without any
adjusted parameter is found when the slip velocity
required for the solution of eqs 6-10 is calculated using
Richardson and Zaki eq 11, as an alternative to the
Zuber and Findlay model (Figure 6).
Mass Transfer. Figure 7 compares experimental
and calculated overall volumetric liquid-liquid mass
transfer coefficients, Kda, referred to the dispersed
phase, a being the interfacial specific area. Experimen-
tal values are from Mateus et al. (1996), obtained in the
absence of gel phase.
Calculated values were obtained using eq 4 with
adjusted parameters for drop size calculation and eqs
6-11 for hydrodynamics calculations ( and uc). Specific
Figure 6. Continuous phase velocity (9 experimental) and area a was calculated assuming spherical drops, hence
dispersed phase holdup in the riser (0 experimental) as a function
of dispersed phase flow rate. Model predictions were made using 
eqs 6-10 plus the Zuber and Findlay model eqs 12 and 13: s a)6 (19)
model predictions using c0 ) 0.5; - - - model predictions using c0 d
) 1.
The slope of the equilibrium curve, m, required in eq
were obtained for an overall K h , on the basis of continu- 14 for the calculation of the overall mass transfer
ous phase velocity in the riser, defined as (from eqs 6-8) coefficient, kd, was taken as an average from experi-
mental equilibrium data in Mateus et al. (1996). The
2FgH dispersed phase film coefficient, kd, was calculated for
K
h ) (18) two different situations, rigid drops and circulating
Fucriser2 drops, using eqs 15 and 16, respectively. The continu-
ous phase film coefficient, kc, was also calculated for
Figure 4 shows the effect of gel phase holdup on K h. rigid and circulating drops, using two expressions for
As expected, the presence of gel beads increases friction, each regime (equations in Table 2).
thus decreasing circulation. Each individual point was Comparison between calculated and experimental
obtained by linear regression of experimental values of overall mass transfer coefficients (Figure 7) shows that
 vs Fucriser2/(2FgH), assuming Kh to be approximately rigid drop correlations give values somewhat lower than
constant for the range of organic flow rates used. The experimental, as would be expected, since drop Reynolds
curve in Figure 4 results from fitting the experimental numbers were in the range 120-200, for which circula-
points. tion usually occurs inside drops. The best agreement
Holdup and Circulation. Figures 5 and 6 compare between calculation and experiment was obtained as-
calculated continuous phase velocities and dispersed suming circulating drops, using eq 16 for the determi-
4564 Ind. Eng. Chem. Res., Vol. 35, No. 12, 1996

nation of kd and Hughmarks correlation (eq iv in Table Rei ) Reynolds number in section i
2) for kc. The correlation of Garner et al. (1959) for kc Rhi ) hydraulic radius, m
in the circulating regime (eq iii in Table 2), on the other t ) contact time, s
hand, gives much too high estimates, reflecting the ub ) drop rise velocity, m s-1
difference in predicted values for the same situation uc ) superficial velocity of the continuous phase in the
from available correlations in the literature and the riser, m s-1
difficulty of a priori choice of appropriate liquid-liquid ud ) superficial velocity of the dispersed phase in riser, m
mass transfer correlations. s-1
ui ) superficial velocity of the continuous phase in section
5. Conclusions i, m s-1
us ) slip velocity, m s-1
Hydrodynamics and liquid-liquid mass transfer in a ut ) drop terminal velocity, m s-1
liquid-impelled loop reactor can be described by a model Va ) volume of aqueous phase
which integrates expressions and correlations available vd ) drop volume after break off, m3
in the literature. For low holdups of the dispersed vN ) average dispersed phase velocity through the orifice,
phase, and in the absence of a third phase, holdup and m s-1
circulation can be calculated from standard expressions Vs ) volume of gel phase
for pressure balance and single-phase friction pressure R ) parameter in eq 4
losses, together with some way of calculating the slip ) parameter in eq 4
velocity between the two phases. This is best done by  ) dispersed phase holdup in the riser
a widely used correlation proposed by Richardson and m ) wavelength of the growing disturbance of the jet
Zaki (1954). The Zuber and Findlay two-phase drift- surface, m
flux model can also be used, with an adjusted param- n ) coefficients in eq 16
eter. The presence of a third, solid phase may be ) viscosity of the continuous phase, kg m-1 s-1
accounted for using an empirical correction to the
F ) density of the continuous phase, kg m-3
friction factor. Mass transfer is well described using
F ) difference in densities between the continuous and
eq 16 for the determination of kd and Hughmarks the dispersed phases, kg m-3
correlation (eq iv in Table 2) for kc.
Fd ) density of the dispersed phase, kg m-3
In brief, this integrated model predicts the mass
) interfacial tension, N m-1
transfer behavior in the LLR solely from knowledge of
n ) coefficients in eq 15
the geometry, physicochemical properties of the system,
and operating conditions (dispersed phase flow rate and
gel load). It thus opens possibilities for optimization of Literature Cited
reactor design and operation. Ayazi Shamlou, P.; Pollard, D. J.; Ison, A. P.; Lilly, M. D. Gas
holdup and liquid circulation rate in concentric-tube air-lift
Acknowledgment bioreactors. Chem. Eng. Sci. 1994, 49, 303.
Buitelaar, R. M.; Susaeta, I.; Tramper, J. Application of the liquid-
The authors thank Teresa Reis for help with inter- impelled loop reactor for the production of antraquinones by
facial tension measurements. The fellowship awarded plant cell cultures. In Progress in Plant Cellular and Molecular
to D.M.R.M. by the CIENCIA/PRAXIS XXI programme Biology; Nijkamp, H. J. J., Van der Plas, L. H. V., Van Artrijk,
J., Eds.; Kluwer: Amsterdam, The Netherlands, 1990; pp 694-
(B.D. 2827) is also gratefully acknowledged.
699.
Clift, R.; Grace, J. R.; Weber, M. E. Bubbles, drops and particles;
Nomenclature Academic Press: New York, 1978.
Elzinga, E. R., Banchero, J. T. Film coefficients for heat transfer
a ) interfacial specific area, m-1 to liquid drops. Chem. Eng. Prog. Symp. Ser. 1959, 55, 149.
Bn ) coefficients in eq 16 Fonseca, M. M. R.; Alves, S. S.; Mateus, D. M. R.; Nunes, I. M. L.
c0 ) parameter in eq 12 Multiphasic bioreactors. In Proceedings of the 6th European
Cn ) coefficients in eq 15 Congress on Biotechnology; Alberghina, L.; Forntali, L.; Sensi,
d ) drop diameter, m P., Eds.; Elsevier: Amsterdam, 1994; p 941.
Dd ) solute diffusivity in the dispersed phase, m2 s-1 Garner, F. H.; Sukling, R. K. Mass transfer from a soluble solid
DN ) orifice diameter, m sphere. A.I.Ch.E. J. 1958, 4, 114.
f ) Fanning friction factor for straight pipe Garner, F. H.; Foord, A., Tayeban, M. Mass transfer from circulat-
g ) acceleration of gravity, m s-2 ing liquid drops, J. Appl. Chem. 1959, 9, 315.
Holland, F. A.; Bragg, R. Fluid Flow for Chemical Engineers, 2nd
H ) height of the two-phase column, m
ed.; Edward Arnold: London, 1995.
K
h ) total number of velocity heads lost by friction based Hu, S.; Kintner, R. C. The fall of single liquid drops through water.
on continuous phase velocity in the riser A.I.Ch.E. J. 1955, 1, 42.
kc ) mass transfer film coefficient for the continuous phase, Hughmark, G. A. Liquid-liquid spray columns, drop size, holdup
m s-1 and continuous mass transfer. Ind. Eng. Chem. Fundam. 1967,
kd ) mass transfer film coefficient for the dispersed phase, 6, 408.
m s-1 Jakob, M. Heat Transfer; Wiley: New York, 1949; Vol. 1.
Kd ) overall mass transfer coefficient for the dispersed Kronig, R.; Brink, J. C. On the theory of extraction from falling
phase, m s-1 drops. Appl. Sci. Res. 1950, A2, 142.
Ki ) number of velocity heads lost by friction in section or Lockhart, R. W.; Martinelli, R. C.; Nelson, D. B. Proposed
fitting i correlation of data for isothermal two-phse, two component flow
Kda ) overall volumetric oxygen transfer coefficient, s-1 in pipes, Chem. Eng. Progr. 1949, 45 (1), 39.
Mateus, D. M. R. Estudos cine ticos e de transfere ncia de massa
Leqi ) equivalent length of section or fitting i, m
num bioreactor agitado pela fase orga nica. M.Sc. Thesis, Insti-
m ) equilibrium coefficient tuto Superior Tecnico, Lisboa, 1994.
Pdf ) pressure driving force for circulation Mateus, D. M. R.; Alves, S. S.; Fonseca, M. M. R. Model for the
Pi ) friction pressure losses in section or fitting i, N m-2 production of L-tryptophan from L-serine and indole by im-
Q ) volumetric flow rate of the dispersed phase, cm s-1 mobilized cells in a three-phase liquid impelled loop reactor.
Re ) Reynolds number Bioprocess Eng. 1996, 14 (3), 151.
Ind. Eng. Chem. Res., Vol. 35, No. 12, 1996 4565
Meister, B. J.; Scheele, G. F. Generalized solution of the Tomotika Biotransformations; Moody, G., Baker, P. B., Eds.; Elsevier:
stability analysis for a cylindrical jet. A.I.Ch.E. J. 1967, 13 (4), London, 1987; p 231.
682. Vedaiyan, S. Hydrodynamics of two phase flow in spray columns.
Meister, B. J.; Scheele, G. F. Prediction of jet length in immiscible Ph.D. Thesis, University of Madras, 1969.
liquid systems. A.I.Ch.E. J. 1969, 15 (5), 689. Vermue, M. H., Tramper, J. Extractive biocatalysis in a liquid-
Perry, R. H.; Chilton, C. H. Chemical Engineering Handbook; impelled loop reactor. Proceedings of the 5th European Congress
MacGraw-Hill, New York, 1973. on Biotechnology; Munksgaard Int. Pub.: Copenhagen, 1990;
Richardson, J. F.; Zaki, W. N. Sedimentation and fluidization: part Vol. I, p 243.
I. Trans. Inst. Chem. Eng. 1954, 32, 35. Walis, G. B. One-dimensional two-phase flow; McGraw-Hill: New
Scheele, G. F.; Meister, B. J. Drop formation at low velocities in York, 1969.
liquid-liquid systems. Part I: prediction of drop volume. Part Zuber, N.; Findlay, J. A. The effect of non-uniform flow and
II: prediction of jetting velocity. A.I.Ch.E. J. 1968, 14 (1), 9. concentration distributions and the effects of local relative
Tramper, J.; Wolters, I.; Verlaan, P. The liquid impelled loop velocity on the average volumetric concentration in two phase
reactor: a new type of density-difference-mixed bioreactor. In flow. Trans. ASME, Ser. C: J. Heat Transfer 1965, 87, 453.
Biocatalyst in Organic Media; Laane, C., Tramper, J., Lilly, M.
D., Eds.; Elsevier: Amsterdam, 1987; p 311. Received for review April 16, 1996
van Sonsbeek, H. M.; Verdurmen, R. E. M.; Verlaan, P.; Tramper,
Revised manuscript received September 29, 1996
J. Hydrodynamic model for liquid-impelled loop reactors. Bio-
Accepted September 26, 1996X
technol. Bioeng. 1990, 36, 940.
van Sonsbeek, H. M.; De Blank, H.; Tramper, J. Oxygen transfer IE960225B
in liquid-impelled loop reactors using perfluorocarbon liquids.
Biotechnol. Bioeng. 1992, 40, 713.
van den Tweel, W. J. J.; Marsman, E. H.; Vorage, M. J. A. W.;
Tramper, J. The application of organic solvents for the biocon- X
Abstract published in Advance ACS Abstracts, November
version of benzene to cis-benzenoglycol. In Bioreactors and 1, 1996.

You might also like