Sunteți pe pagina 1din 14

Airlift Bioreactors: Review of Recent Advances

By Jos C. Merchuk*
Department of Chemical Engineering, Ben-Gurion University of the Negev, Beer-Sheva 84105, Israel

he present review roughly covers the publications that have


appeared since Prof. K. Schgerl delivered a keynote address at
the 4th GLS International Conference in 1997 (Schgerl, 1997). It
includes publications that deal with gas/liquid systems without solids,
but gives an extra weight to applications of airlift bioreactors to
wastewater treatment and to algal growth. Three extensive reviews have
been published lately (Chisti, 1988; Merchuk and Gluz, 1999; Petersen
and Margaritis, 2001). Most of the publications selected for this review
are not considered. The material is presented under several subtitles
fluid dynamic characterization briefly describes some of the reported
data on gas holdup, liquid velocity and mass transfer in airlift reactors. A
section on models follows. This is a field in which several very relevant
papers have been published. These papers are critically reviewed, trying
to enlighten common characteristics and differences among the
approaches adopted. Wastewater treatment is discussed next, because
of the relevance of the subject and of the close relationship to the
modeling section, since some of the most interesting mathematical
models were developed in relation to wastewater. A short review of some
of the design modifications proposed is then presented, followed by a
review of some of the bioprocesses that have been developed choosing
airlift reactors. Among these, algal growth has been given special
consideration.

Fluid-dynamic Characterization
The largest amount of publications related to airlift reactors during
recent years, report on measurements of fluid dynamic characteristics
and mass transfer rates. Some of them are valuable confirmations or
extensions of prior knowledge, and some of them report experimental
results in modified airlift design, which usually cannot be represented by
prior correlations. In any case, the need for collecting and unifying all
those experimental results still exists. We suggested years ago a unified
format for airlift data collection and correlation (Har-Noy et al., 1997).
Al-Masri and Abasaeed (1998) did extensive measurements of liquid
velocity and riser holdup in a series of external loop airlift reactors, and
proposed empirical correlations that fit their data better than other
proposed correlations. Still there is here a lack of a massive bank of data
for validation of those correlations.
One of the most basic characteristics of gas/liquid dispersions is the
bubble size. Couvert et al. (1999) have presented a series of very careful
measurements of liquid velocity, riser and downcomer gas holdup and
Sauter-mean bubble diameter. The last measurements were related to
the gas pressure in the tubular membranes that acted as spargers, and

*Author to whom correspondence may be addressed. E-mail address: jcm@


bgumail.bgu.ac.il

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Airlift reactors are popular in the modern


bioprocess research and development, over a broad
spectrum of processes. Those range from the production
of very expensive biochemicals to wastewater
treatment. In both extremes the reason behind the
selection of the airlift reactor is related to fluid
dynamic characteristics. During the last few years, a
large number of research papers dealing with the
dependence of reactor performance on liquid flow
have been published. In some of them, descriptions of
devices that have been developed in order to improve,
or to gain further control of the fluid patterns in the
reactor are reported.
Les racteurs airlift sont couramment rencontrs
dans la recherche et le dveloppement de
bioprocds modernes, et ce pour une vaste gamme
de procds allant de la production de produits
biochimiques trs chers au traitement des eaux uses.
Dans ces procds, le choix du racteur airlift est li
aux caractristiques dynamiques du fluide. Ces
dernires annes, de nombreux articles de recherche
traitant du lien entre la performance du racteur et
lcoulement des fluides ont t publis. Certains
dentre eux donnent des descriptions de systmes mis
au point dans le but damliorer ou de mieux
contrler les profils dcoulement dans le racteur.
Keywords: airlift bioreactors, fluid dynamics, models,
wastewater treatment, novel design, photosynthesis.

therefore it is possible that the measured results


cannot be extrapolated to coalescing systems with
other type of gas spargers. Nevertheless, their twophase hydrodynamic study is quite complete, and fits
the model that they present.
Bendjaballah et al. (1999) made a study of liquid
velocity and gas holdup in the riser of an external loop
airlift reactor. They used a valve in order to manipulate
liquid velocity independently of gas superficial
velocity, and compared a single-orifice gas sparger
and a multiple-orifice gas sparger. Their results
resemble very closely those published by Stein and
Merchuk (1981), with the difference that no
variations of holdup along the riser were measured

changing both solid density and solid loading. The mass


transfer rate was calculated assuming that the reactor can be
considered a perfectly mixed volume. The correctness of this
assumption can be easily proven by the criterion given in the
next section as Equation (3). Taking as an example a
gas superficial velocity of 0.2 m/s, representative values of their
measured circulation time and volumetric mass transfer coefficient
would be, respectively, 10 s, and 0.01 s1. Thus in Equation (3)
b~10>>1, and the system indeed behaves as a perfectly mixed
vessel. This will be the case in any airlift reactor with a relatively
short draft tube. Their results can be very useful because they
can be considered as obtained in a perfectly mixed system, thus
well defined, and can therefore be used also in the design of tall
bioreactors that require separate consideration of each defined
fluid dynamic region (structured reactor model).
Figure 1. Variation of slip velocity VGL with total gas-liquid velocity for
different downcomer diameters. The minimum in each curve indicates
transition from homogeneous to heterogeneous flow. Adapted from
Bendjaballah et al. (1999).

here. The interesting point in this work is the use of the graphs
displaying gas slip velocity versus overall fluid velocity in order
to make a quantitative determination of the transition from
homogeneous gas flow, where the slip velocity decreases with
the total flow, to heterogeneous gas flow where it increases
with the total flow. Figure 1 shows the transitions for three
different diameters of the downcomer, each giving a different
resistance to liquid circulation. Similar results were found by
partially closing the valve in the downcomer. The same group
attacked the problem of regime identification from a quite
different angle (Vial et al., 2001), proposing the use of
auto-correlation functions of wall-pressure fluctuations for
identification of flow regimes in airlift reactors. This method
seems to be experimentally simple and very promising for this
purpose.
Van Benthum et al. (1999c) studied the transitions between
gas recirculation regimes in airlift reactors that they define as
follows: Regime 1 (no gas recirculation), Regime 2 (stationary
bubble front in the downcomer) and Regime 3 (gas recirculation).
It is apparent that Regime 2 has little interest from an industrial
point of view. The gas that stays practically stationary in the
downcomer will not have much influence on mass transfer,
since it will fast become depleted of oxygen, or other
components of interest. The authors made a systematic study of
the effect of suspended solids and liquid throughflow, and
managed to define criteria for the transition from Regime 2 to
Regime 3. The transition depends on the suspended solids and
on the throughflow of liquid.
See et al. (1999) made a careful study of the effect of drag
and frictional losses on the hydrodynamics of airlift reactors, to
elucidate in which measure the use of available methods and
correlations are applicable to airlift reactors. They recommend a
careful evaluation of the frictional losses in all the elements of
the circuit, including straightening vanes, gas liquid separator,
and riser-wall friction. On the other hand they found that the
use of correlations for fully developed single-phase flow in the
downcomer (their external loop reactor has no gas recirculation)
does not introduce much error.

Effects of Solids on Mass Transfer Rate


The influence of solids on oxygen transfer rate was studied by
Freitas and Texeira (1998, 2001). They studied extensively the
influence of suspended solids on oxygen absorption rate,

Models
Airlift reactors especially the fluid dynamics of airlift reactors
have a strong appeal and there are many researchers that
have tackled the task of describing the flow of liquid, gas
and solids in these bioreactors. In Table 1, some of the models
that have been published during the last six years are
presented. Twelve of them model gas/liquid (G/L) systems,
and eight attempt the representation of three-phase,
gas/liquid/solid (G/L/S) systems. Some overlapping exists, since
several models are presented in a general form that fits both
G/L and G/L/S systems.
The table shows clearly that the liquid velocity and the gas
holdup in the riser are seen as the main variables studied and
appear in all the proposed models. Indeed, those variables are
closely related, since the difference in gas holdup between riser
and downcomer is the only driving force for liquid circulation in
the system. This has been clearly stated by Heijnen et al. (1997),
which expressed the pressure difference per unit height
required as:
DP
= (e Gr - e Gd ) - (e Sr - e Sd )(r S - r L )
gH

(1)

In closed reactors, or as long as the overall solid amount does


not change, the last term is the product of two differences that
have always the same sign, since if the density of the solid is
larger than that of the liquid, its rising velocity will be smaller
and the holdup of solids in the riser will be larger than in the
downcomer, and vice versa. The presence of solids, therefore,
will always diminish the driving force for circulation, independently of their density.
The hydrodynamic model presented by Heijnen et al. (1997)
is one of the most interesting models, and offers a complete
representation of the fluid dynamics in an airlift reactor from a
macroscopic point of view. The model is able to describe both
two- and three-phase flow. The ingenuity of the model consists
in avoiding the need of dealing with four different variables (the
holdups of gas and liquid in both riser and downcomer) by
writing the difference in gas holdups eGr-eGd as a function of the
overall gas holdup eG, the gas superficial velocity and the
circulation liquid velocity. A similar procedure is followed with
respect to solids holdup. One point that may be criticized in this
model is the use of the simplification proposed by Chisti (1989),
which assumes that the ratio of the gas holdups in downcomer
and riser is constant. This is an approximation that seems to be
The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Table 1. Selected models for airlift fluid dynamics and mass transfer
Source
Heijnen et al.(1997)
van Benthum et al. (1999a)
Garcia-Calvo et al. (1999)
Hwang and Lu (1997)
Freitas et al.(1999)
Mudde and Van der Akker (2001)
Camarasa et al.(2001a)
Camarasa et al.(2001b)
Mousseau et al. (1998)

2Phase

3Phase

Gas
recirculation

VL

x
x

x
x

x
x

x
x

x
x
x

x
x

x
x
x
x

eGr

eGd

x
x
x
x
x

x
x
x
x

x
x

x
x

Cockx et al.(1997)
Couvert et al.(2001)
Tobajas et al.(1999)
Camacho-Rubio et al.(2001)

x
x

x
x
x
x

x
x
x
x

x
x

x
x
x
x

x
x
x
x

Sanders et al. (2001)


Shechter et al.(2002)
Orejas (1999)
Steiff et al.(1997)

x
x

close to reality in two-phase systems, though, it contradicts


simple liquid mass balance. This has been recognized by Chisti
himself (Contreras et al., 1998).
In the case of three-phase flow, the point becomes more
complicated. A liquid balance at the top results in:
V A
V A
e Gd = Lr r e Gr - Lr r (1 - e Sr ) + (1 - e Sd ) = ae Gr + b
VLd Ad
VLd Ad

(2)

and the range over which a and b can be considered constants


is still an open question.
Seven of the models listed in Table 1 consider gas recirculation. Since Heinjens model (Heijnen, 1997) circumvents the
need of knowing the gas holdup in riser and downcomer, in
spite of handling the regime of gas recirculation, it does not
provide an evaluation of the fraction of gas recirculated. The
model presented by Freitas et al. (1999) uses a pseudohomogeneous liquid approach, and assumes that the concentration of solids is homogeneous all over the reactor. This is
strictly valid only for solid density that is very close to the density
of the liquid, rS~rL. Indeed, they use alginate beads as solid
carriers, with density close to that of water. Friction coefficients
are calculated from one-phase correlations. The model confirms
that an increase in solids produces a decrease on holdup. The
gas holdup in the downcomer is calculated as in the model by
Heijnen et al. (1997), assuming a linear relationship between
riser and downcomer holdups, or in other words, constant
values for a and b in Equation (2). For the external loop they
assume no gas recirculation. They optimize their parameters (4
parameters), and get excellent fitting of their data.
Tall airlift reactors constitute a special type because the high
liquid velocities that can be reached. Not only can gas be
The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

CFD

Remarks
3 regimes
ALR and
extension
kLa
es constant

kLa,
ammonia

x
x

eG

x
x

eSd

Sez et al. (1998)


Mrquez et al. (1999)

eSr

x
Axial
profiles
x
kLa
O2 axial
profiles

x
x
x

entrapped into the downcomer, but it can be directly injected


in this section, with a considerable gain in gas phase residence
time and in energy required for gas injection. One of the most
promising applications is the disposal of carbon dioxide from
flue gas by injection into deep waters (Kosugi et al., 2001).
Sanders et al. (2001) have presented a model for the prediction
of liquid velocity in such reactors. A basic analysis based on the
Zuber and Findlay (1965) two-phase flow approach seems to
represent satisfactorily the trends of pressure drops and liquid
velocities in a high-recirculation airlift reactor.
Garcia Calvo et al. (1999) presented a thermodynamic model
(based on the first principle), which is simply an extension of a
previous two-phase model (Garcia Calvo, 1989), to the threephase airlift reactor (TPAL). They simply replace the liquid
density in the original model by the liquid-solid pseudohomogeneous phase density. The model requires knowledge of
the gas slip-velocity. They show that it is possible to predict the
transition from packed bed to fluidized bed and to circulating
bed, as well as the hysteresis phenomenon that has been
observed in these flow configurations (Heck and Onken, 1988),
and good concurrence with experimentally observed transitions
is reported. These transitions between flow configurations
classified taking the solid phase as reference, are different from
the transitions between configurations based on the gas
behavior, as those described by van Benthum et al. (1999c).
Mousseau et al. (1998) present a model for a rectangular
section airlift reactor with suspended solids of lower density
than water, used for wastewater treatment: the model allowed
the prediction of the changes in ammonia, dissolved oxygen
and biomass. They use the penetration model and the isotropic
turbulence model for the description of the gas-liquid
mass transfer.
In three of the papers in Table 1, the problem of gas/liquid
mass transfer is addressed and is given more attention than fluid

dynamics: Hwang and Lu (1997), Tobajas et al. (1999) and


Camacho Rubio et al. (2001). Hwang and Lu (1997) present an
analysis of gassing out experiments, where a different value of
the mass transfer coefficient kLa is considered in each of the
three distinct regions: riser, downcomer and separator. This is in
fact an improved version of the original paper by Merchuk et al.
(1992), where this structured analysis of an airlift reactor was
first presented, structured meaning that zones of different fluid
dynamic characteristics are recognized. Those are identified
within the reactor, each of them having potentially a different
gas holdup, liquid dynamics and mass transfer coefficient kLa.
Knowledge of the liquid velocity, gas holdups and liquid dispersion coefficient are required for the use of the Hwang and Lu
(1997) model.
The model by Tobajas et al. (1999) refers to a three-phase
system, since it deals specifically with the biotreatment of
marine sediment. A simple model is proposed combining
Higbies penetration model and Kolmogoroffs theory of
isotropic turbulence, and is used for prediction of mass transfer
rates. No parameters are adjusted, and slip velocity, friction
coefficients, equivalent length, etc. are taken from independent
sources. An excellent match of experimental and predicted
values was found for gas holdup in the riser, liquid velocity and
mass transfer rate.
Camacho Rubio et al. (2001) deal with the axial profiles of
oxygen in tall airlift reactors, which are frequently ignored. They
show that, as a consequence of the hydrostatic pressure
variation, steady-state axial dissolved oxygen concentration
profiles exist even if no oxygen consumption exists. These axial
inhomogenities increase as the gas flow rate increases. It is
obvious, therefore, that the conventional method of kLa
determination, which is based on the assumption of perfect
mixing in the reactor, is not valid in a tall bioreactor. The
problem of inhomogeneity in dissolved oxygen concentration
was touched also by Dhaouadi et al. (2001), and Korpijarvi et al.
(1999), starting from the dynamic response of an airlift reactor
to sudden changes in inlet gas concentration. In both cases they
also made measurements of oxygen concentration and derived
the mass transfer coefficient using the axial dispersion model. In
all those cases, the main point of interest is the recognition that
the common assumption of perfect mixing may not hold in the
case of tall reactors.
This had been previously analyzed by Siegel and Merchuk
(1990). The criterion proposed there was that the parameter b,
ratio of mass transfer characteristic time to liquid circulation
time, should be:

b=

1
>1
k L at c

(3)

In order to be able to consider the system as perfectly mixed.


Tall reactors will have longer circulation times, and the inequality
above will not be fulfilled. In these cases, the calculation of the
mass transfer coefficients from experimental data should be
carried out using models like those proposed by Camacho
Rubio et al. (2001) or Hwang and Lu (1997).
Table 1 lists two papers that deal with the application of
computational fluid dynamic codes to the description of the
flow in airlift reactors. Cockx et al. (1997) presents a simulation
based on one-dimensional two-fluid mass and momentum
balances. Their results produce axial profiles of gas holdup in

both riser and downcomer, and a flow map of the liquid that
includes radial profiles of velocity. Mudde and Van Den Akker
(2001) carried out two- and three-dimensional simulation of
an airlift, trying to minimize the use of ad-hoc closure
terms. Surprisingly, they do not find clear advantage in
the three-dimensional model, since both approaches coincide
with experimental measurements of gas holdup. The two
dimensional model predicts a strong influence of the
sparger geometry, and seems to be better at low gas superficial
velocities.
Camarasa et al. (2001a, and b) developed a model for
external-loop airlift reactors (no gas recirculation) based on a
careful momentum balance, using the drift-flux model (Zuber
and Findlay, 1965) for gas holdup prediction. The model fits
fairly well their experimental results of mean velocity and gas
holdup in the riser, as well as axial variations of the local gas
holdup along the riser. They combined this fluid dynamic
information with a cells-in-series model for mass transfer, which
also addresses the mixing behavior of the phases.
Couvert et al. (2001) present a simple model for the prediction
of liquid velocity and gas holdup in rectangular airlift reactors of
different scales. They do consider gas recirculation, and their
model predicts gas holdups in both riser and downcomer,
which are close to the measured values.
The fundamental and solid model presented a decade ago by
Young et al. (1991) for an external airlift reactor (1991) was
upgraded (Sez et al., 1998), adding to the initial model the
effect of gas buoyancy forces in the gas. Based only on the
physical properties of the gas and liquid phases, the reactor
dimensions and the gas input, the model predicts gas and
liquid velocity and gas holdup along the riser for bubbly flow.
Total disengagement of the gas at the top is assumed.
The seminal model by Ho et al. (1977) has been an inspiration
and a basis for much improved models that represent the airlift
reactor as a sequence of perfectly mixed cells (Camarasa et al.
2001b; Steiff et al., 1997; Orejas, 1999). Those models are
convenient for handling biological or chemical reactions. The
simplification of the mathematical treatment may sometimes
outweigh the unrealistic stepwise change of the variables
along the reactor that is obtained. Steiff et al. (1997) produced
a considerable amount of experimental data that can be used
to calibrate the model parameters. Of special interest, since
these data are scarce in the literature, are the data of gas
recirculation, measured following the technique proposed by
Siegel et al. (1986).
Shechter et al. (2002) presented a three-phase model for the
fluid dynamic description of an airlift unit, part of an overall
process (referred to below). The model is based on a
momentum balance over riser and downcomer, and a series of
macroscopic equations that describe the continuity of liquid
and solid as it passes from the riser to the downcomer. Figure 2
shows the influence of solid loading on the variables of the
system: liquid and solid velocity in the riser (the corresponding
values in the downcomer are not shown to avoid overcharging
the figure), gas and solid holdup in the riser and solids holdup
in the downcomer. The model predicts a small increase in both
gas holdup and solids velocity as the amount of solids increases
at constant gas superficial velocity. The solids holdup in both
the riser and the downcomer increases as expected, and the
increase in the downcomer is much sharper than in the riser.
The most affected variable is the liquid velocity, ULr, which
decreases almost to half due to the change in solids loading.
The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 2. Influence of solids loading on riser gas holdup (er ), riser and
downcomer solids holdups (eSr and eSd), liquid and solid velocity in the
riser (ULr and USr), in a rectangular section airlift reactor of a water
treatment system, as predicted by the model by Shechter et al. (2002).
The superficial gas velocity was 0.006 (m/s), the area ratio (Ar/Ad)=
2.85, the solid density 950 (m/s) and the liquid density 955 (m/s). The
line-identifiers are:  = er (-) ;  = eSr (-) ;  = eSd (-) ; x = ULr (m/s) ; *
= USr (m/s).

Another interesting simulation in Figure 3 shows the effect of


solid density on the behavior of the system: while the gas
holdup in the riser and the velocity of the solids in the riser
remain almost constant when the density of the solid changes
by 5% under and over the density of water, the solids holdup
in the riser and downcomer are affected in opposite ways: while
eSr increases eSd decreases. The lines cross at a particle density
equal to that of water. The difference between riser and
downcomer solids holdup changes thus from negative to
positive as rs increases. The difference in absolute terms is
smaller (approximately 0.03 versus 0.074) at higher particle
density, and the predicted increase in liquid velocity, therefore,
concurs with the general conclusions from Equation (1). This
simulation refers to solids with density close to that of the
liquid. When the density of the particle is substantially higher,
the additional problem of complete fluidization appears. This is
especially critical in systems with short draft tubes, as shown by
Kojima et al. (1999)

Airlift Reactors in Wastewater Treatment


Since one of the recognized characteristics of airlift reactors is
the potential for scaling up and the relatively low power
consumption for agitation and oxygenation, it is only natural
that many processes related to wastewater treatment use this
type of reactor.
Jin et al. (2002) used an airlift reactor in a comprehensive
pilot plant system for starch processing wastewater (SPW)
reclamation. The starch was utilized by Aspergilius orizae.
Simultaneously to a 95% COD, 93% BOD and 98% suspended
solids removal, an important production of a-amylase (~50
EU/ml) was obtained. An interesting point in this paper is the
dependence of fungal morphology on ALR fluid dynamics. In
this type of processes, morphology of the fungal biomass is
extremely important. Free mycelial growth (wild growth)
increases strongly the viscosity, limiting the oxygen transfer rate

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 3. Influence of particle density on riser gas holdup (er ), riser and
downcomer solids holdups (eSr and eSd), liquid and solid velocity in the
riser (ULr and USr), in a rectangular section airlift reactor of a water
treatment system, as predicted by the model by Shechter et al. (2002).
The superficial gas velocity was 0.006 (m/s), the area ratio (Ar/Ad)=
2.85, the solid density 950 (m/s) and the liquid density 955 (m/s). The
line-identifiers are: :  = er (-) ;  = eSr (-) ;  = eSd (-) ; x = ULr (m/s) ; *
= USr (m/s).

from the gas to the culture. The solution that has been almost
universally adopted for this problem which attains to citric
and other organic acids, antibiotics, etc. is to find the
conditions under which the biomass takes the form of fungal
pellets. The advantage in gas-liquid transfer rate, because of the
decrease in viscosity, usually outweighs the added transfer
resistance stemming from the intraparticle diffusion of oxygen.
But the formation of pellets in optimal size and compactness is
a very complex matter. Metz and Kossen (1977) pointed out
the multiplicity of variables and the difficulties of a priori prediction of the operation conditions required. Not much has been
advanced in this matter since. Jin et al. (1999) found empirically
the optimal gas flow rate in their 4.5 L ALR. It is worthwhile to
note that they seem to have been able to scale up these
conditions to their pilot-plant 160 L ALR.
Lazarova et al. (1997) studied experimentally the fluid
dynamics and the performance for wastewater treatment of an
split-vessel airlift with a rectangular section. They studied
carefully the influence of suspended solids on gas holdup both
in the riser and the downcomer, as well as the influence of
the ratio of riser to downcomer cross sectional areas on
liquid velocity. They compare the experimentally measured
velocities for different reactor heights without proposing any
correlation. Their measurements of mass transfer rate do not
reveal important changes with respect to those obtained in
water once the biofilm is developed. The main aspect stressed
by the researchers is the capacity for nitrification observed in
various stages.
Many applications of ALR have been reported in processes
where the point of interest here is simply that the process,
which can take place in a conventional stirred tank, can be run
using an ALR as well, with the consequent savings in energy
requirements, etc. For example, the use of Aspergilius niger for
textile wastewater (biological decoloration) was reported by
Assadi and Jahangiri (2001). Campos et al. (2002) used an ALR
in a combined (microfiltration and biological) treatment of
oilfield wastewater treatment. They obtained satisfactory results
in TOC and COD reduction in a continuous process, using an

ALR with suspended polystyrene particles using hydraulic


retention times from 12 to 48 h. Both the above-mentioned
studies were carried out in small-scale reactors.
Loh and Liu (2001) used an external loop fluidized bed airlift
bioreactor (EIFBAB) for treatment of high strength phenolic
wastewater. To control the oxygen transfer they used the
increase in gas holdup that they got by closing a valve in the
downcomer and restricting liquid circulation. The range of
variation in their device goes in fact from holdup in an airlift
with unrestricted circulation to the holdup in a bubble column,
for similar diameters and gas superficial velocity. Obviously, the
case of a completely closed valve implies that the downcomer
volume will not contribute to the process.
Bakker et al. (1996) immobilized their biomass inside
k-carageenan gel beads, and studied a cascade of two small
scale ALRs to study the oxidation of nitrite to nitrate by
Nitrobacter agilis. This is an important step in the nitrification
process (i.e., the oxidation of ammonia to nitrate via nitrite,
usually followed by a denitrification stage with reduction of
nitrate to N2). They found advantages in the use of two bioreactors
in series, and attributed it to the kinetics of the process
(non-competitive substrate and product inhibition). Because
the density of the beads was close to unity, there was no
problem in fluidizing of the beads in spite of the small scale.
While the basic characteristics of an airlift reactor indicate its
fitness for aeration of large volumes of wastewater, the problem
of ammonia removal calls for special handling. Two approaches
have been presented lately incorporating the nitrificationdenitrification element into a basic airlift arrangement. The first
one is the already mentioned biofilm airlift suspension extension
(BASE) reactor (van Benthum 1999a, 1999b), which present the
very compact design that can be seen in Figure 4. The conventional
airlift with its three phases is enclosed into an additional vessel
(extension) that becomes the anaerobic volume. Part of the
liquid and suspended biofilm coated solids overflows the
aerobic airlift core and enter the top of the extension, reentering at
the bottom. The design allows the control of aerobic/anaerobic
times for the biofilm-coated particles suspended in the system
in order to improve the nitrification/denitrification of
wastewater. The flow of liquid and suspended solids in the
extension, which is the anaerobic volume, can be controlled
manipulating the overpressure in the headspace of the reactor.
A mathematical model was developed and used for the design
of a pilot plant. The experimental results of gas and solid
holdups concur satisfactorily with the model.
Garrido et al., (1997) showed that in the BASE (Heijnen et al.,
1997) they could manipulate the system to obtain higher nitrite
than nitrate, in presence of both Nitrosomonas ( nitrite) and
Nitrobacter strains ( nitrate). They present a simple model
based on diffusion in the biomass film, with homogeneous
distribution of microorganisms. At low O2 tensions, nitrification
is improved. The model assumes zero order reaction for
Nitrosomonas and Blackman kinetics for Nitrobacter.
A different approach to the integration of nitrification/
denitrification in a wastewater treatment process is the one
presented by Shechter et al. (2002). They present an approach
aiming at the upgrade of existing wastewater treatment plants
rather than de novo design. The experiments presented were
obtained in a 230 m3 aeration basin that was divided in 9
sections, as shown in Figure 5. Seven of these were converted
to airlift operation (the fluid dynamic model of which has been
commented above), with two anaerobic sections. In this case,

Figure 4. Scheme of the BASE (biofilm airlift suspension extension)


system. Arrows indicate the flow direction. Liquid, solids and gas flow
from the downcomer to the riser and vice versa; Only liquid and solids
are present in the extension of the airlift. Adapted from van Benthum
et al., (1999a), with permission.

therefore, the system is a once-through aerobic/anoxic/


aerobic/anoxic/aerobic sequence, with options of recirculation
for improved denitrification. The conceptual difference
between this approach and the BASE system is that the biomass
is not recirculated and remains stationary in each of the nine
stages, since solids circulate in each airlift stage, but only liquid
passes from one stage to the other. In the BASE approach, all
the biofilm-coated particles transit cyclically both aerobic and
anaerobic sections.

Design Modifications
We have commented already on some interesting airlift design
modifications, as the BASE shown in Figure 4, where an
extension is added to the reactor providing a region of fluid and
solids flow that is completely different from the usual airlift flow
regions, and still can be controlled, mainly by the headspace
pressure in the main reactor. The GLAD (gas lift advanced
dissolution), a completely different concept, is a method for
sequestration of low purity CO2 emitted by thermal power
plants (Kosugi et al., 2001). The dissolution tube is approximately
200 m long, and the drainpipe reaches more than 1000 m in
depth. In fact, this is not a gas lift reactor, since this is a
once-through system, and recirculation, a basic characteristic of
airlift reactors, does not exist. It is therefore, a gas-lift pump in

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 5. The Agar System (Shechter et al., 2002). The existing


aeration basin was divided into 9 stages, stages 4 and 9 being anoxic,
and the other seven rectangular-section airlift reactors. Sludge is fed
into the first stage, and also into the anoxic sections.

which gas dissolution is especially important. The scheme of the


system can be seen in Figure 6.
A very interesting system, where the drastic change is not in
the reactor design but in the system itself was presented by Sajc
and Vunjak-Novakovic (2000). It consists in an extractive
bioconversion taking place in a four-phase external-loop airlift
bioreactor. The aim is the extractive bioconversion of
anthroquinones by plant cells (Fragnula alnus). The cells are
immobilized in alginate beads suspended in an aqueous
medium. In the riser, gas and a solvent (silicone oil or
n-hexadecane) are injected. The gas bubbles provide the energy
for the circulation of the system, and the solvent drops extract
the anthroquinones produced. Both gas and solvent are
completely separated at the top and only the aqueous medium
and the suspended alginate beads circulate. Figure 7 shows a
scheme of the setup, where the inlet and outlet of gas and
solvent are indicated.
A device that changes sensibly the fluid dynamics in the
reactor is the helical flow promoter (HFP) (Gluz and Merchuk,
1996). The conventional airlift reactor assures a homogeneous
and relatively ordered flow, with all the elements of fluid
circulating in a cyclic pattern from the riser to the downcomer.
However, in case of large devices, the radial mixing in the
downcomer may not be sufficient. The HFP consists in a series
of static baffles that modify the flow in all the reactor (Figure 8).
This modified flow has several consequences, all of them
potentially beneficial to the yield of the process:
a) The helical movement causes secondary flow, which leads to
an enhanced radial mixing, and therefore more homogeneous
distribution of the suspended cells and dissolved substrates.
b) One of the most important characteristics of the HFP is the
enhanced capacity for fluidizing solid particles. Thus, it is
especially useful for a process operating with high cell
densities. It has been shown that the minimal gas flow rate for
complete fluidization of solids in an air-lift reactor may be up
to four times lower when using HFP (Schotelburg et al., 1999).
c) The mass transfer rate from the liquid to the suspended
solids may be enhanced up to 50% due to the higher

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 6. The GLAD (gas lift advanced dissolution), a method for


sequestration of low purity CO2 emitted by thermal power plants.
From Kosugi et al. (2001), with permission.

relative velocity between the particles and the liquid. This


property can be of relevance if the transference of nutrients
or oxygen to a suspended particle becomes rate controlling.
Schlotelburg et al. (1999) made an extensive study of the
influence of the HFP on reactor fluid dynamics and mass transfer
rates, including the effect of liquid viscosity. This was a comparative study of a bubble column, an airlift reactor and an airlift
reactor with HFP. They found that while the fluidization capacity
increased very much because the presence of HFP, the mass
transfer rate diminished. This was attributed to an increased
coalescence due to the corkscrew-like pattern generated in the
riser. On the other hand, the reduction of the mass transfer
rate due to a viscosity increase, which was very strong in the
case of the a bubble column, was milder in the airlift and much
milder in the airlift with HFP. This would indicate some
advantage for use of HFP in processes where a large change in
viscosity is expected.
Wu and Merchuk (2003) conducted measurements of fluid
flow in the downcomer of an internal loop airlift reactor using a
novel optical trajectory tracking system especially developed
with this purpose. Analysis of the experimental results shows
that plug flow exists in this region. Shear stress was also
analyzed and was found homogeneous. On the other hand the
measurements made on a similar airlift reactor with HFP
revealed the existence of secondary flow in the downcomer.
Figure 9 is the flow mapping of the resultant of the axial and
radial components of the liquid velocity, uz and ur in the

Figure 7. Extractive bioconversion in a four-phase external-loop airlift bioreactor. In the riser, gas and a solvent are injected. Both gas and solvent
are completely separated at the top and only the aqueous medium and the suspended alginate beads circulate. From Sajk and Vunjak-Novakovic
(2000), with permission.

downcomer of the ALR with HFP. Inspection of the graphs


reveals that for a given axial position z, the radial component
changes in an almost cyclic way with the azimuth q, being
directed to the exterior and to the axis alternatively. The same
type of fluctuation of ur is seen along z for a fixed q. The
component in the q direction (not shown here) shows a general
trend to decrease slowly from the top to the bottom of the
reactor. It is apparent that the measured velocities are
generated by a vortex-like movement of the liquid. The experimental evidence seems to suggest a stream coiling around itself
as it descends following a helix around the axis. The experimental results are thus a clear indication of the existence of
secondary flow in the downcomer.

Bioprocess Applications
While airlift reactors have become very popular in research
institutes, the stirred tank reactor remains as the undisputed
leader in the realm of industrial bioreactors. That is possibly the
reason many publications that compare airlift and stirred tank
reactors. Kim et al. (1997) compared the production of
b-glucosidase by Aspergillius niger in various bioreactors, and
reported that the best results correspond to an airlift reactor.

Figure 8. The Helical Flow Promoters: The case of HFP located at the
top of the downcomer. The grey arrows indicate the flow of the liquid.

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 10. Comparison of a concentric airlift reactor and a stirred tank


bioreactor for the for the production of Anticarsia gemmatalis multiple
nucleopolyhedrovirus (AgMNPV) in UFL-Ag-286 insect cells.The
cultures were infected synchronically at a high multiplicity of infection
in a serum-free medium. The time is given in days post infection. From
Visnovsky et al. (2003).

Figure 9. Flow mapping of the resultant of the axial and radial


components uz and ur in the downcomer of the ALR with HFP with
superficial gas velocity of 0.015 m/s, and draft tube diameter of 0.12 m.
(Wu and Merchuk, 2003).

However, in some processes, such as the bio-oxidation of


minerals, the airlift seems to be the natural choice because of
the combination of a satisfactory environment for micro-organism
growth and fluidization of solids (Ruitenberg et al., 2001).
In the case of growth of shear sensitive cells, as animal or
plant cells, the airlift reactor has been long ago chosen as one
of the best, and sometimes the only, solutions. Yuan et al.
(1999) presented a modified airlift column that includes an
external resin column through which the liquid is pumped and
returned to the reactor. Alkaloids are secreted by the Cathrantus
roseus cells, which are immobilized in a polyurethane foam in
the riser, and are eliminated from the liquid by adsorption on
the resin column. This enhances noticeably the alkaloid production by the cells.
Su et al. (1996) cultivated plant cells (Anchusa officinalis)
suspended in the medium in an airlift reactor with a calm
sedimentation zone created by a baffle in the downcomer, to
overcome the problem of cells that remain suspended at the

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

top and do not return to the bulk of the liquid. They report to
have doubled the final concentration of the plant cells, and also
the amount of secreted proteins to the medium.
Visnovsky et al. (2003) reported a comparison of a concentric
airlift reactor and a stirred tank bioreactor, for the growth of
UFL-Ag-286 insect cells, and for the production of Anticarsia
gemmatalis multiple nucleopolyhedrovirus (AgMNPV). While no
relevant differences could be observed in the doubling time and
viability of the insect cells, important differences were found in
the kinetics of adsorption of AgMNPV-NOVs (non-occluded
virus) on the insect cells, and in the rate of production of
AgMNPV-OVs (occluded virus). The NOVs were more
quickly and efficiently adsorbed on the cells in the airlift reactor
than in the stirred tank reactor. The onset of OVs production
was earlier in the airlift reactor than in the stirred tank reactor
(Figure 10). In addition, the titers of OVs obtained in the airlift
reactor were slightly higher than those obtained in the stirred
tank reactor for both baculovirus progenies. These results
indicate clearly the influence of reactor fluid dynamics on the
performance of the process.

Airlift Reactors for Algal Culture


Sanchez Miron et al. (2000) have focused their study on airlift
bioreactors for algal culture. They studied extensively the fluid
dynamics (liquid velocity, riser and downcomer holdup and
mass transfer coefficient) in a concentric-tubes airlift reactor,
and a split-column reactor and compared the results with those
corresponding to a bubble column of identical dimensions. The
aim of the paper was to relate algal growth to aeration rate, gas
holdup and liquid velocity. However, their results of
Phaeodactylum tricornutum growth gave essentially the same
results for the three reactors, in spite of the differences in the
fluid dynamic and mass transfer characteristics between the
three bioreactors that they measured.
This is in contrast to the results presented by Merchuk et al.
(2000), which cultivated the red microalga Porphyridium sp. in a
bubble column, an airlift reactor and an airlift reactor with HFP
of similar geometry, and found clear differences among the

results obtained in each reactor. The essential difference


between the reports of these two groups may be attributed to
different kinetics in the photosynthetic growth process of the
two different algae used.
Among the factors controlling growth, light availability is the
most important one. Light flux decreases exponentially with
distance measured from the irradiated side of the photo-bioreactor.
The algae near the irradiation source are exposed to high
photon flux density, which enhances growth rate. The cells at
the core of the reactor receive less light as a result of mutual
shading and will show a lower growth. On the other hand,
excessive light intensity can damage protein D1 in photosystem
II (photoinhibition), and decrease growth rate due to reduction
in the number of active photon traps (Powles, 1984; Krause,
1988). An important feature that must be pointed out is that
growth rate can be influenced not only by intensity but also by
the history of the illumination that the cells experience (Lee and
Pirt, 1981). A periodical change in illumination can also
enhance growth (Marra, 1978). Previous work (Merchuk et al.,
1998) based on the fact that ordered mixing can enhance light
availability and photosynthesis suggested a general approach
for integrating fluid dynamics with the mathematical description of
photosynthesis (Figure 11). In this approach, fluid dynamics is
the key factor determining the history of the illuminations of the
cells.
The kinetics used was based on the three states model of
photosynthetic factories (PSF) originally proposed by Eilers and
Peeters (1988). The PSF is defined as the sum of light trapping
system, reaction centers and associated apparatus, which are
activated by a given amount of light energy to produce a
certain amount of photoproduct. An important assumption is
that the PSF has three states, the resting state (open), called x1,
the activated state (closed), called x2, and the inhibited state,
called x3. The PSF in resting or open state can be stimulated and
transferred to the activated state by the capture of a photon.
The PSFs in activated state may follow one of two possible
paths; either receiving another photon, which produces
temporarily inhibition, or passing the gained energy to
acceptors and starting the photosynthesis, returning then to the
open state. Also the inhibited PSF can recover after some time,
returning to the open state. The kinetic constants for each of
these steps were obtained for Porphyridium species in independent
experiments (Wu and Merchuk, 2001).
The fluid flow pattern in an airlift reactor is relatively
simpler than that in a bubble column, which has been treated
lately (Wu and Merchuk, 2002). The fluid flows in a defined
circulation pattern through the channels defined by the
geometry of the ALR. In an internal loop ALR, the fluid is driven
to pass the downcomer, the riser, and the separator successively
as shown in Figure 12. For the photosynthesis process, the
region of riser can be regarded as dark zone, and the rest as a
light zone with variable illuminance. In the dark zone, the cell
kinetics does not depend on the illuminance, and hence the
flow pattern in this region has no influence on the light utilization.
It was shown elsewhere that perfect mixing can be assumed in
the gas separator (Merchuk and Yunger, 1990, Merchuk et al.,
1996). Since usually the residence time in this region is small,
the light history can be satisfactorily approximated as one of
constant illuminance. The downcomer, however, is the major
region where the cells are illuminated, and detailed flow pattern
in this region is of interest. Wu and Merchuk (2003) studied the
fluid dynamics in this region with an optical trajectories-tracking

10

Figure 11. Schematic description of the integration of fluid dynamics


and photosynthetic growth, from Merchuk et al (1998). Cells in the
illuminated region can capture photons and start the photosynthetic
chain. They may also become inactivated due to photoinhibition. After
being transported to the dark region, only the dark reactions of
photosynthesis and recovery from photoinhibition occur. The cells are
transported cyclically from illuminated to dark zones.

Figure 12. Cyclic light history of cells in the Air Lift Reactor, according
to the model by Wu and Merchuk (2002). Illuminance is taken a nil in
the riser, a mean value is taken for the separator, and in the riser it
varies with the radius of the column.

system (OTTS) developed by them, and found practically plug


flow in the range of interest. This was therefore the reactor
dynamics adopted in the mathematical model. It should be
mentioned that both the experiments and the modeling
mentioned relate to the range of interest for photosynthetic
processes, which corresponds to low gas flow rates. In this
range, homogeneous, bubbly flow is expected in the riser, and
none or minimal gas recirculation in the downcomer.

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 13. Simulation of the effect of column diameter on cell growth


in an ALR. The initial cell concentration was x0 = 8 106 (cel/mL), the
height of the column 1 (m), the gas superficial velocity was 0.00331
(m/s), and the light intensity I0=250 mEm2s1. Wu and Merchuk
(2003).

According to the discussion above, the light history of a cell


in the ALR can be represented by the scheme shown in Figure
12. The downcomer is divided into several radial regions
according to the prevailing illuminance. The algal cells are
assumed to be homogeneously distributed in the downcomer,
and the fraction of the cells in each light zone can be calculated
according to the geometry of the light zones while the gas
separator can be considered as a perfectly mixed volume, and
the riser as a dark zone.
The model allows the simulation of the photosynthetic
growth in an airlift bioreactor. Figures 13 to 15 show those
effects of airlift configuration parameters on the growth. In this
simulation, the illuminance is constant. Figure 13 is the profile
of growth working with different column diameters, at HT = 1.0 m.
It shows that the increase in biomass concentration (Dx = xf x0)
diminishes as the column diameter increases. This decrease in
the biomass concentration gain is very sharp at low column
diameter (approximately D < 0.2 m) and much slower for large
diameters (D > 0.4 m). For example, if the diameter of the
column id increased from 0.5 to 1 m, the gain in concentration
would decrease in 40%. The volume, on the other hand,
wouold increase four-fold. This clearly indicates that a proper
analysis of the system should include the costs of biomass
separation.
The value of biomass gain (Dx = xf x0) decreases also as the
ratio of Ar/Ad increases. At Ar/Ad=1.46, the increase in biomass
is zero, i.e., no net positive growth can be obtained at this
condition. In Figure 14, the effect of the column height on
growth is presented. Here the column diameter is fixed to be
0.2 m, as in the actual experiments. The growth decreases as
the draft tube height increases, due to the increase in the
residence time in the riser. As the column height increases, the
cell stays in the dark riser longer, and this leads to a decrease in
growth. Another effect is that as the height increases, the shear
stress increases also. In Figure15, the effect of Ar/Ad on growth
is presented. The column diameter and column height are fixed
to be 0.2 m and 1.0 m, respectively. It shows that the growth

The Canadian Journal of Chemical Engineering, Volume 81, June-August 2003

Figure 14. Simulation of the effect of draft tube height on cell growth.
The initial cell concentration was x0 = 8 106 (cel/ml), the gas superficial velocity was 0.00331 (m/s), and the diameter of the column 0.2
(m) and the light intensity I0 = 250 mEm2s1. Wu and Merchuk
(2003).

Figure 15. Simulation of the effect of the ratio of cross sectional areas
Ar/Ad on the cell growth. The initial cell concentration was x0 = 8 106
(cel/ml), the gas superficial velocity was 0.00331 (m/s), The height of
the column was 1 (m), the diameter of the column 0.2 (m) and the
light intensity I0 = 250 mEm2s1. Wu and Merchuk (2003).

first increases slowly as Ar/Ad increases, and reaches the highest


point at value of Ar/Ad = 0.8. When Ar/Ad goes beyond 1.0, the
growth drops dramatically. This point should be regard as a
critical ratio (Rc). When at smaller ratio (Ar/Ad< Rc), the liquid
velocity in the downcomer is smaller than in the riser, and cells
receive more light. This may cause some photoinhibition. As the
ratio increases, the photoinhibition is tempered because a
longer dark period is introduced. However, a increase in dark
time may lead to shortage of photon capture and the negative
growth effect of the dark period is enhanced. When the dark
time is further increased, the overall growth drops and finally
reaches negative values. The profile indicates that when designing
an airlift photobioreactor, the ratio of Ar/Ad, must be chosen
carefully to avoid losses in biomass growth. An additional
conclusion that can be drawn from the numerical results shown
on Figure 15 is that in spite of the apparent monotony in the
influence of Ar/Ad in Figures 13 and 14, a maximum of the
increase in biomass exists, at Ar/Ad slightly below 1. This
maximum is not sharp, and is masked there because of the
round values chosen for the parametric representation.

11

The simulations presented here are only qualitative in nature,


since they are based on extrapolations. Nevertheless, the trends
show the potential of this model for simulation and design of a
larger scale airlift bioreactors for algal biomass production in
closed systems.

Selection Strategy
A point that usually appears at the early stages of the development
of a system where an ALR is considered is the selection of the
type of bioreactor. Should an external circulation ALR be
adopted, or rather an internal circulation ALR?
In the case of bioreactors (except biological leaching of
minerals) fluidization is not a serious problem. The solids have
usually a density that is not far apart of that of water, and
fluidization is attained at relatively low liquid velocities.
Therefore, the higher liquid velocities that can be attained in
external loop ALR are not a decisive point. On the contrary,
excessive liquid velocity may translate into damages to the cells
due to shear stress. The main difference remaining is therefore,
the presence of gas in the downcomer. In processes where
the oxygen concentration in the liquid may be depleted in
the downcomer, recirculation of gas will alleviate the
problem. Obviously, a wise consideration of the geometric
design, in order to shorten the residence time in the
downcomer will also contribute.
Our present understanding of the dynamic behaviour of the
ALR system indicates that rather than maintaining the classic
classification of those reactors into internal and external
recirculation devices, attention should be given to the design of
the top of the reactor where gas separation takes place. Since
most of times the main element to take into consideration is gas
disengagement, the elements must be balanced will be: a) The
time that a bubble must spend in the gas separator before
entering the downcomer (function of the liquid velocity in the
gas separator and the length of path of the liquid element), and
b) The ascending velocity of the bubble in the liquid. A clear
example would be a split-vessel ALR, that may be seen as an
external recirculation reactor on one hand, but offers also a
short residence time in the gas separator which provides the
basic characteristics of an internal recirculation ALR.
It should be taken into account that this aspect constitutes a
serious scale-up problem. As the diameter, or equivalent
diameter of the reactor increases, the path of the liquid element
traveling from riser to downcomer becomes longer and the
chances of bubble disengagement increase too. Ingenuity in
the geometric design has an important role here.

Recapitulation
The distinctive characteristics of airlift reactors are conferred by
the fluid dynamics of the gas/liquid or gas/liquid/solid systems
circulating. These characteristics are usually expressed as gas
holdup, liquid and solid velocities and mass transfer rate. It is
important for the design engineer to recognize whether there is
a need to consider those variables separately for each of the
distinct zones of the reactor (structural model). The decision on
this can be made comparing the circulation time and the mass
transfer characteristic time (kLa1). In any case, only a correct
understanding of the behaviour and interconnection of riser,
separator, downcomer and bottom of the reactor will allow the
reliable scale up from the laboratory to pilot or industrial size.
Several models that allow the simulation of the fluid dynamics
of airlift reactors have been presented during the last years. It

12

seems that the mechanisms begin to be clear, and the main


differences among the models remains in the closure relationships,
and in the validity of using in airlift reactors prediction methods
of drag and frictional losses that are generally accepted in
developed flow.
Several modifications in the design airlift reactors have been
proposed during the last years, and these novel configurations
improve the performance of the reactors for specific processes,
like nitrification/denitrification steps in wastewater treatment,
or improve the capacity of fluidization or the radial mixing in
the system.
An interesting application of ALRs is their use as
photo-bioreactors, taking advantage of the ordered fluid
circulation to monitor the light/dark cycles of the photosynthetic
cells. The influence of the gas input and the geometry of the
ALR on the performance of the photo-bioreactor can now be
predicted, qualitatively at least.
Many correlations are available for the prediction of gas
holdup, liquid velocity, solid circulation and mass transfer rates,
and several new ones have been added lately. However, those
are usually limited to a certain design type, and no generalized
equation with wide range validity exists. A joint analysis of all
data available would probably be of help, but this has not been
done. The designer confronting a scale up or a de novo design
must therefore be extremely careful and analyze the validity of
the correlations chosen.

Nomenclature
Ar
Ad
b
D
g
H
I
I(t)
k
kLa
r
P
t
tc
T
ur
uz
ULr
UGr
USr
uq
VGL
VLr
VLd
x1
x2
x3
x0
xf
Dx
z

cross-sectional area of the riser, (m2)


cross-sectional area of the downcomer, (m2)
ratio of characteristic time for mass transfer and circulation time
reactor diameter, (m)
gravitational constant, (m/s2)
draft tube height, (m)
illuminance, or photon flux density, (m.E/m2.s)
illuminance history of a photosynthetic cell, (m.E/m2.s)
yield of photosynthesis production to the transition of x2 x1
volumetric mass transfer coefficient, (1/h)
radial distance in the reactor, (m)
pressure drop, (Pa)
time, (s)
liquid circulation time, (s)
liquid residence time, (s)
radial component of the liquid velocity, (m/s)
axial component of the liquid velocity, (m/s)
liquid superficial velocity in the riser , (m/s)
gas superficial velocity in the riser , (m/s)
solids superficial velocity in the riser , (m/s)
circular component of the liquid velocity, (rad/s)
gas slip velocity, (m/s)
liquid velocity in the riser, (m/s)
liquid velocity in the downcomer, (ms1)
fraction of PSF in open state
fraction of PSF in close state
fraction of PSF in inhibited state
initial biomass concentration, (kg/m3)
final biomass concentration, (kg/m3)
increase in biomass concentration (xf- x0), (kg/m3)
axial distance in the reactor, [m]

Greek Symbols
eG
eGd
eGr
eS

gas volumetric fraction


gas volumetric fraction in the downcomer
gas volumetric fraction in the riser
volume-averaged solid volumetric fraction

The Canadian Journal of Chemical Engineering, Volume 81, August 2003

eSd
eSr
q
rL
rS
Dr

solid volumetric fraction in the downcomer


solid volumetric fraction in the riser
circular distance in the reactor, (radians)
liquid density, (kg/m3)
solid density, (kg/m3)
(rS-rL), (kg/m3)

Subscripts
d
r
s

downcomer
riser
separator

Abbreviations
ALR
BOD
CFD
COD
G/L
G/L/S
HFP
NOV
OV
PFD
PSF
STR
TOC

airlift reactor
biological oxygen demand
computational fluid dynamics
chemical oxygen demand
gas/liquid
gas/liquid/solid
helical flow promoter
non-occluded virus
occluded virus
photon flux density
photosynthetic factory
stirred tank reactor
total organic carbon

References
Al-Masri, W.A. and A.E. Abasaeed, On the Scale-up of External Loop
Airlift Reactors, Newtonian Systems, Chem Eng. Sci 53, 40854094
(1998).
Assadi, M.M. and M.R. Jahangiri, Textile Wastewater Treatment by
Aspergilius niger, Desalination 141, 16 (2001).
Bakker, W.A.M., P. Kers, H.H. Beefting, J. Tamper, and C. D. de Gooijer,
Nitrite Conversion by Immobilized Nitrobacter agilis in an Airlift
Loop Bioreactor Cascade: Effects of Combined Substrate and Product
Inhibition, J Ferm. Bioeng. 81, 390393 (1996).
Bendjaballah, N., H. Dhaouadi, S. Poncin, N. Midoux, J.M. Hornut, and
G. Wild, Hydrodynamics and Flow Regimes in External Loop Airlift
Reactors, Chem. Eng. Sci. 54, 52115221 (1999).
Camacho Rubio, F., J.L. Garcia, E. Molina, and Y. Chisti, (2001). Axial
Inhomogeneities in Steady-State Dissolved Oxygen in Airlift
Bioreactors, Predictive Models, Chem. Eng. J. 84, 4355 (2001).
Camarasa, E., E. Carvalho, L.A.C. Meleiro, R. Maciel Filho, A.
Domingues, G. Wild, S. Poncin, N. Midoux, and J. Bouillard, A
Hydrodynamic Model for Air-Lift Reactors, Chem. Eng. Proc. 40,
121128 (2001a).
Camarasa, E., E. Carvalho, L.A.C. Meleiro, R. Maciel Filho, A.
Domingues, G. Wild, S. Poncin, N. Midoux, and J. Bouillard,
Development of a Complete Model for an Air-Lift Reactor, Chem.
Eng. Sci. 56, 493502 (2001b).
Camarasa, E., E. Carvalho, L.A.C. Meleiro, R. Maciel Filho, A.
Domingues, G. Wild, S. Poncin, N. Midoux, and A. Bouillard,
Complete Model for Oxidation in Airlift Reactors, Computers &
Chem. Eng. 25, 577584 (2001c).
Campos, J.C., R.H.M. Borges, A.M. Olivera Filho, R. Nobrega, and
G.L. Sant Anna Jr., Oilfield Wastewater Treatment by Combined
Micro Filtration and Biological Precess, Water Research 36, 95104
(2002).
Chisti, Y., Airlift Bioreactors. Elsevier Applied Science, New York, NY
(1989).
Chisti, Y., Pneumatically Agitated Bioreactors in Industrial and
Environmental Bioprocessing: Hydrodynamics, Hydraulics and
Transport Phenomena, Appl. Mech. Rev. 51, 33112 (1998).
Cockx, A., A. Line, M. Roustan, Z. Do-Quang, and V. Lazarova,
Numerical Simulation and Physical Modeling of the Hydrodynamics
of an Airlift Internal Loop Reactor, Chem. Eng. Sci. 52, 37973793
(1997).

The Canadian Journal of Chemical Engineering, Volume 81, August 2003

Contreras, A., Y. Chisti, and E. Molina, A Reassessment of Relationship


between Riser and Downcomer Gas Holdups in Airlift Reactors,
Chem. Eng. Sci. 53, 41514154 (1998).
Couvert, A., M. Roustan, and P. Chatellier, Two-Phase Hydrodynamic
Study of a Rectangular Air-Lift Loop Reactor with an Internal Baffle,
Chem. Eng. Sci. 54, 52455252 (1999).
Couvert, A., D. Bastoul, M. Roustan, A. Line, and P. Chatellier,
Prediction of Liquid Velocity and Gas Holdup in Rectangular
Airlift Reactors of Different Scales, Chem. Eng. Proc. 40, 113119
(2001).
Dhaouadi, H., S. Poncin, N. Midoux, and G. Wild, Gas Liquid Mass
Transfer in an Airlift Reactor-Analytical Solution and Experimental
Confirmation, Chem. Eng. Proc. 40, 129133 (2001).
Eilers, P.H.C. and J.C H. Peeters, A Model for the Relationship between
Light Intensity and the Rate of Photosynthesis in Phytoplankton,
Ecological Modeling 42, 199215 (1988).
Freitas, C. and J.A. Texeira, Hydrodynamic Studies in an Airlift Reactor
with Enlarged Degassing Zone, Biprocess Eng. 18, 267279 (1998).
Freitas, C., M. Fialova, J. Zahradnik, and J.A. Teixeira, Hydrodynamic
Model for Three-Phase Internal- and External-Loop Airlift Reactors,
Chem. Eng. Sci. 54, 52535258 (1999).
Freitas, C. and J.A. and Texeira, Oxygen Mass Transfer in High Solids
Loading Three-Phase Internal-Loop Airlift Reactor, Chem. Eng. J. 84,
5761 (2001).
Garcia Calvo, E., A Fluid Dynamic Model for Airlift Reactors, Chem.
Eng. Sci. 46, 29472951 (1989).
Garcia Calvo, E., A. Rodriguez, A. Prados, and J. Klein, A Fluid Dynamic
Model for Three-Phase Airlift Reactors, Chem. Eng. Sci. 54,
23592370 (1999).
Garrido, J.M., W.A.J. van Benthum, M.C.M. van Loosdrecht, and J.J.I.
Heijnen, Influence of Dissolved Oxygen Concentration on Nitrite
Accumulation in a Biofilm Airlift Suspension Reactor, Biotechnol.
Bioeng. 53, 168178 (1997).
Gluz, M.D. and J.C. Merchuk, Modified Airlift Reactors: The Helical
Flow Promoters, Chem. Eng. Sci. 51, 29151920 (1996).
Har-Noy, G., M. Gluz, P. Shoval, U. Onken, and J.C. Merchuk, in
Preprints of the 4th Japanese/German Symposium Bubble
Columns97 S.C.E.J., Kyoto, Japan (1977), pp. 240242.
Heck, J. and U. Onken, Characteristics of Solid Suspension in a Bubble
Column with and without Draft Tube, Chem. Eng. Sci. 44,
17431745 (1988).
Heijnen, J.J., J. Hols, R.G.J.M. van der Lans, H.L.J.M. van Leeuwen, A.
Mulder, and R. Weltevrede, A Simple Hydrodynamic Model for
Liquid Circulation Velocity in a Full-Scale Two- and Three-Phase
Internal Airlift Reactor Operating in the Gas Recirculation Regime,
Chem. Eng. Sci. 52, 25272540 (1997).
Ho, C.S., L.E. Erickson, and L.T. Fan, Modeling and Simulation of
Oxygen Transfer in Airlift Fermenters, Biotechnol. Bioeng. 19,
15031522 (1977).
Hwang, S.-J. and W.-J. Lu, Gas-Liquid Mass Transfer in an Internal Loop
Airlift Reactor with Low Density Particles, Chem. Eng. Sci. 52,
853857 (1997).
Jin, B., H.J. van Leeuwen, B. Patel, and K. Yu, Utilization of Starch
Processing Wastewater for Production of Microbial Biomass Protein
and Fungal a-Amylase by Aspergilius orizae, Bioresource Technol. 66,
201206 (1998).
Jin, B., H. J. van Leeuwen, B. Patel, and K. Yu, Mycelial Morphology and
Fungal Protein Production from Starch Processing Wastewater in
Submerged Cultures of Aspergilius orizae, Proc. Biochem. 34,
335340 (1999).
Jin, B., Y. X.Q., K. Yu, and H.J.A. van Leeuwen, Comprehensive Pilot
Plant System for Fungal Biomass Protein Production and Wastewater
Reclamation, Adv. Environ. Res. 6, 179189 (2002).
Kim, S.W., S.W. Kang, and J.S. Lee, Cellulase and Xylanase Production
by Aspergilius Niger KKS in Various Bioreactors, Bioresource Techn.
96, 6367 (1997).
Kojima, H., J. Sawai, H. Uchino, and T. Ichige, Liquid Circulation and
Critical Gas Velocity in Slurry Bubble Column with Short Size Draft
Tube, Chem. Eng. Sci. 54, 51815185 (1999).

13

Korpijarvi, J., P. Oinas, and J. Reunanen, Hydrodynamics and Mass


Transfer in an Airlift Reactor, Chem. Eng. Sci. 54, 22552262
(1999).
Kosugi, S., K. Niwa, T. Saito, and K. Hoamaoji, Design Factors in GasLift Advanced Dissolution (Glad) System for CO2 Sequestration in the
Ocean, Chem. Eng. Sci. 56 56, 62056210 (2001).
Krause, G.H., Photoinhibition of Photosynthesis. An Evaluation of
Damaging and Protective Mechanism, Physiol. Plant. 74, 566574
(1988).
Lazarova, V., J. Meyniel, L. Duval, and J.A. Menem, Novel Circulating
Bed Reactor, Hydrodynamics, Mass Transfer and Nitrification
Capacity, Chem. Eng. Sci. 53, 39193927 (1997).
Lee, Y. and S.J. Pirt, Energetics and Photosynthetic Algal Growth.
Influence of Intermittent Illumination in Short (40s) Cycles, J. Gen.
Microbiol. 124, 4352 (1981).
Loh, K. and J. Liu, External Loop Inversed Fluidized Bed Airlift
Bioreactor (Eifbab), Chem. Eng. Sci. 56, 61716176 (2001).
Mrquez, M.A., A.E. Sez, R.G. Carbonell, and G.W. Roberts, Coupling
of Hydrodynamics and Chemical Reaction in Gas-Lift Reactors,
AIChE J. 45, 410423 (1999).
Marra, J., Phytoplankton Photosynthetic Response to Vertical
Movement in a Mixed Layer, Mar. Biol. 46, 203210 (1978).
Merchuk, J.C. and R. Yunger, The Role of the Gas-Liquid Separator of
Airlift Reactors in the Mixing Process, Chem. Eng. Sci. 45,
29732976 (1990).
Merchuk, J.C., G. Osemberg, M. Siegel, and M. Shacham, A Method
for Evaluation of Mass Transfer Coefficients in the Different Regions of
Air Lift Reactors, Chem. Eng. Sci. 47, 22212226 (1992).
Merchuk, J.C., N. Ladwa, A. Cameron, M. Bulmer, A. Pickett, and I.
Berzin, Liquid Flow and Mixing in Concentric-Tube Airlift Reactors,
J. Chem. Technol. Biotechnol. 66, 174182 (1996).
Merchuk, J.C., M. Ronen, S. Giris, and S. Arad (Malis), Light-Dark
Cycles in the Growth of the Red Microalga Porphyridium sp.
Biotechnol. Bioeng. 59, 705713 (1998).
Merchuk, J.C. and M. Gluz, in Encyclopedia of Bioprocesss Technology:
Fermentation, Biocatalysis, Bioseparation, M. C. Ficklinger and W. D.
Stephen Eds., John Wiley & Sons, New York, NY (1999), Vol. 1, pp.
320353.
Merchuk, J.C., M. Gluz, and M.I., Comparison of Photobioreactors for
Cultivation of the Red Microalga Porphyridium, J. Chem.Technol.
Biotechnol. 75, 11191126 (2000).
Mousseau, F., S.X. Liu, S.W. Hermanowicz, V. Lazarova, and J. Menem,
Modeling of Turboflow- a Novel Reactor for Wastewater Treatment,
Water Sci. Technol. 37, 177181 (1998).
Mudde, R.F. and H.E.A. Van Den Akker, 2-D and 3-D Simulations of an
Internal Airlift Loop Reactor on the Basis of a Two-Fluid Model,
Chem. Eng. Sci. 56, 63516358 (2001).
Orejas, J.A., Modelling and Simulation of a Bubble-Column Reactor
with External Loop: Application to the Direct Chlorination of
Ethylene, Chem. Eng. Sci. 54, 513522 (1999).
Petersen, E.E. and A. Margaritis, Hydrodynamic and Mass Transfer
Characteristics of Three-Phase Gaslift Bioreactor Systems, Crit. Rev.
Biotechnol. 21, 233294 (2001).
Powles, S.B., Photoinhibition of Photosynthesis Induced by Visible
Light. Ann. Rev. Plant Physiology 35, 1544 (1984).
Puteh, M., K. Shimizu, T. Kanai, and Y. Kawase, in Preprints of the 4th
Japanese/German Symposium Bubble Columns97 S.C.E.J, Kyoto,
Japan (1997), pp. 365371.
Ruitenberg, R., C. E. Schultz, and C. J. N. Buisman, Bio-Oxidation of
Minerals in Airlift-Loop Bioreactors, Int. J. Mineral Proc. 62, 271278
(2001).
Sez, A.E., M.A. Mrquez, G.W. Roberts, and R.G. Carbonell,
Hydrodynamic Model for Gas-Lift Reactors, AIChE J. 44,
14131423 (1998).
Sajc, L. and G. Vunjak-Novakovic, Extractive Bioconversion in a FourPhase External-Loop Airlift Bioreactor, AIChE J. 46, 13681375 (2000).
Sanchez Miron, A., F. Garcia Camacho, A. Contreras Gomez, E. Molina
Grima, and Y. Chisti, Bubble-Column and Airlift Photobioreactors for
Algal Culture, AIChE J. 46, 18721887 (2000).

14

Sanders, D A., H. Cawte, and A.D. Hudson, Modeling of the Fluid


Dynamics Processes in a High-Recirculation Airlift Reactor, Int. J.
Energy Research 25, 487500 (2001).
Schlotelburg, C., M. Gluz, M. Popovic, and J.C. Merchuk,
Characterization of an Airlift Reactor with Helical Flow Promoters,
Can. J. Chem. Eng. 77, 804810 (1999).
Schgerl, K., Three-Phase-BiofluidizationApplication of Three-Phase
Fluidization in the Biotechnology-a Review, Chem. Eng. Sci. 52,
36613668 (1997).
See, K. H., R.G.W., and A.E. Sez, Effect of Drag and Frictional Losses
on the Hydrodynamics of Gas-Lift Reactors, AIChE J. 45, 24672471
(1999).
Shechter, R., J.C. Merchuk, and T. Ronen, presented at the WEFTEC
2002, Chicago, IL, Sep.28-Oct 2, unpublished (2002).
Shechter, R. and J.C. Merchuk, Modeling of an Airlift Reactor with
Floating Solids for Wastewater Treatment, Can. J. Chem. Eng.
(2003).
Siegel, M., J.C. Merchuk, and K. Schgerl, Gas Recirculation in Air-Lift
Reactors, AIChE J. 32, 15851596 (1986).
Siegel, M. and J.C. Merchuk, presented at the APBioChEC 90,
unpublished (1990).
Steiff, A., S. Gran-Heedfeld, S. Schlter, and P.M. Weinspach, presented
at the Preprints of the 4th Japanese/German Symposium Bubble
Columns97, Kyoto, Japan unpublished (1997).
Stein, Y. and J.C. Merchuk, Local Hold-up and Liquid Velocity in Air-Lift
Bioreactors, AIChE J. 27, 377388 (1981).
Tobajas, M., E. Garcia-Calvo, M.H. Siegel, and S.E. Apitz, (1999),
Hydrodynamics and Mass Transfer Prediction in a Three-Phase Airlift
Reactor for Marine Sediment Biotreatment, Chem. Eng. Sc. 54,
53475354 (1999).
van Benthum, W.A.J., R.G.J.M. van der Lans, M.C.M. van Loosdrecht,
and J. J. Heijnen, Bubble Recirculation Regimes in an Internal-Loop
Airlift Reactor, Chem. Eng. Sci. 54, 39954006 (1999a).
van Benthum, W.A.J., R.G.J.M. van der Lans, M.C.M. van Loosdrecht,
and J. J. Heijnen, The Biofilm Airlift Suspension Extension Reactor.
Part Ii: Three-Phase Hydrodynamics, Chem. Eng. Sci. 54,
19091924 (1999b).
van Benthum, W A.J., J.H.A. van den Hoogen, R.G.J.M. van der Lans, M.
C. M. van Loosdrecht, and J. J. Heijnen, The Biofilm Airlift
Suspension Extension Reactor. Part I: Design and Two-Phase
Hydrodynamics, Chem. Eng. Sci. 54, 19091924 (1999c).
van Metz, B. and N.W.F. Kossen, Biotechnology Review, the Growth
of Molds in the Form of Pellets- a Literature Review, Biotechnol.
Bioeng. 16, 781799 (1977).
Vial, C., S. Poncin, G. Wild, and N. Midoux, A Simple Method for
Regime Identification and Flow Characterization in Bubble Columns
and Airlift Reactors, Chem. Eng. Proc. 40, 135151 (2001).
Visnovsky, G., J.D. Claus, and J.C. Merchuk, Cultivation of Insect Cells
in Bioreactors: Influence of Reactor Configuration and Superficial
Velocity, Latin American Applied Research 33, 117121 (2003).
Wu, X. and J.C. Merchuk, A Model Integrating Fluid Dynamics in the
Photosynthesis and Photoinhibition Process, Chem. Eng. Sci. 56,
35273538 (2001).
Wu, X. and J.C. Merchuk, Simulation of Algae Growth in a Bench Scale
Bubble Column, Biotechnol. Bioeng. 80, 15668 (2002).
Wu, X. and J.C. Merchuk, Measurement of Fluid Flow in the
Downcomer of an Internal Loop Airlift Reactor Using an Optical
Trajectory Tracking System, Chem. Eng. Sci. in press (2003).
Young, M.A., R.G. Carbonell, and D.F. Ollis, Airlift Bioreactor: Analysis
of Local Two-Phase Hydrodynamics, AIChE J. 37, 403412 (1991).
Yuan, Q., H. Xu, and Z. Hu, Two-Phase System for Enhanced Alkaloid
Synthesis and Release in a New Airlift Reactor by Catarantus Roseus,
Biotechnology Techniques 13, 107109 (1999).
Zuber, N. and J.A. Findlay, Average Volumetric Concentration in Two
Phase Flow Systems, J. Heat Transfer 87, 453468 (1965).

Manuscript received December 4, 2002; revised manuscript received


June 4, 2003; accepted for publication June 5, 2003.

The Canadian Journal of Chemical Engineering, Volume 81, August 2003

S-ar putea să vă placă și