You are on page 1of 49

Chemical Engineering Communications

ISSN: 0098-6445 (Print) 1563-5201 (Online) Journal homepage: http://www.tandfonline.com/loi/gcec20

AIRLIFT REACTORS: CHARACTERISTICS,


APPLICATIONS AND DESIGN CONSIDERATIONS

M.Y. CHISTI & M. MOO-YOUNG

To cite this article: M.Y. CHISTI & M. MOO-YOUNG (1987) AIRLIFT REACTORS:
CHARACTERISTICS, APPLICATIONS AND DESIGN CONSIDERATIONS, Chemical Engineering
Communications, 60:1-6, 195-242, DOI: 10.1080/00986448708912017

To link to this article: http://dx.doi.org/10.1080/00986448708912017

Published online: 03 Apr 2007.

Submit your article to this journal

Article views: 240

View related articles

Citing articles: 119 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gcec20

Download by: [Cornell University Library] Date: 06 October 2016, At: 23:27
Chem. Eng. Comm. 1987, vol60, pp. 195-242
Photocopying permitted by license only
0 1987 Gordon and Breach Science Publishers S.A.
Printed in the United States of America

AIRLIFT REACTORS: CHARACTERISTICS,


APPLICATIONS AND DESIGN CONSIDERATIONS
M.Y. CHISTIT and M. MOO-YOUNG
Industrial Biotechnology centre
Department of Chemical Engineering
Uniuersiry of Waterloo
Waterloo, Ontario
Canada N2L 3GI
(Received January 14, 1987; in final form June 5, 1987)

Bioreactors of the airlift type are a promising design for aerobic fermentations. The basic knowledge
required for understanding and predicting the performance of these reactors is only now beginning to
emerge. In this review we present our observations and those of other investigators in an attempt to
build up a coherent picture of airlift devices. All the major aspects-mixing and hydrodynamics, mass
and heat transfer-in these reactors are considered. Comparisons between bubble columns and airlift
systems are made where analogies, similarities and/or differences between them provide insight into
airlift systems. Throughout, the areas of particular concern and those in need of further research in
this field are mentioned. Extensive work on all forms of airlift reactors, particularly in non-Newtonian
media-homogeneous and suspensions-remains to be done. Current knowledge does not permit
airlift reactor design with a high degree of confidence. However, the technical feasibility of all types of
fermentations-dant cell, tissue culture, bacterial, fungal,
- and those utilizing yeasts-in airlift vessels
has been demonkrated.
KEYWORDS Airlift reactor Hydrodynamics Mass transfer Heat transfer

1. INTRODUCTION

Bioreactors are the core of any bioprocess. Gas-liquid or gas-slurry reactors are
used in aerobic fermentations. The most common aerobic reactors used in
commercial productive fermentations are of the stirred tank variety' and the
design of these dates from 1940's' when they were used for the first modern
industrial fermentation-that of the antibiotic penicillin. Pneumatic reactors, in
which all agitation is due to bubbling gas, are a relatively recent invention. An
exception is the simple bubble column reactor which, despite its long history of
application in fermentation, is not widely used. Airlift reactors are also pneumatic.
Other completely new designs of fermentation vessels and modifications of some
of the earlier types have been proposed. These include bubble columns with
many forms of internal^,^^ pulsed bubble column^,^ airlifts with stirrers,' tubular
loop which may be suitable for small volume fermentations, and
aeration using a downward directed two-fluid nozzle.'0." Spouted beds1' also find

t Author to whom all correspondence should be addressed.


196 M.Y. CHISTI AND M. MOO-YOUNG
applications in aerobic fermentations. In this review we deal principally with
airlift systems.

2. A I R L I l T REACTORS: GENERAL CONCEPTS

Airlift reactors consist of a liquid pool divided into two distinct zones only one of
which is usually sparged by a gas. The different gas holdup in the gassed and
ungassed zones results in different bulk densities of the fluid in these regions
which causes circulation of the fluid in the reactor by a gas-lift action. The part of
the reactor containing the gas-liquid upflow is the riser and the region containing
the downflowing fluid is known as the downcomer (Figure 1).
Theoretically, airlift reactors may be employed for any gas-liquid or gas-slurry
contacting process. Practical application depends on the ability to achieve the
required rates of momentum, heat and mass transfer at acceptable capital and
operating costs. The technical and economic feasibility of using airlift devices has
been conclusively established for a number of processes and these reactors find
increasing use in aerobic fermentations, treatment of wastewater and other
similar operations. The simplicity of their design and constr~ction,'~better
defined flow patterns in them and comparatively low power inputs for requisite
transport rates, make them very attractive. Low shear fields, good mixing and
extended aseptic operation made possible by elimination of stirrer shafts, seals
and bearings are important advantages of airlifts in fermentation applications.
Continuous production of beer, vinegar, citric acid, and biomass from yeasts,
bacteria and fungi has been carried out in airlift vessels at different working
capacities.13 The USSR and Eastern Europe have extensively employed airlift
vessels for SCP yeast c~ltivation'~ and in England 1CI has operated at 1500 m3
(working volume) airlift fermenter for the PRUTEEN process.15 Ho et a1.I6 have
mentioned several currently used industrial scale applications of airlift fermen-
ters. Malfait et a[.'' claimed a more than 18% (by weight) enhancement in yield
of a filamentous mould Monascus purpureus in an external loop airlift (0.055 m3,

-
DOWNCOMER ( DOWNFLOW I
GAS SPARGED RISERS
I U P FLOW )

FIGURE 1 Schematic of an airlift reactor


AIRLIFT BIOREACTORS 197
0.15 m diameter riser, 0.05 m diameter downcomer) relative to that in a stirred
tank (0.100 m3 volume, operated at 1VVM air flow, 3- standard turbines
(6-blades) operated at 300rpm, 3000-4000 ~ m agitation - ~ power). This yield
improvement was achieved with a 50% reduction in power input leading to a
more than 50% reduction in the cost of biomass produced in the airlift. Improved
productivity of the airlift reactor relative to the stirred vessel was associated with
the higher overall mass transfer coefficient obtained in the airlift." Although this
was not mentioned, possible shear damage to cells in the stirred vessel may have
been a contributory factor to its poor performance. Similarly, Erickson et a1."
quoted a report where novobiocin production using Streptomyces niveus in 0.1
and 0.2m3 airlift fermenters was found to be as good or better compared to
results obtained in stirred tanks. Excellent results have been reportedlg for
penicillin fermentation in an external loop device. Several other examples of
successful cultivation of mycelial fungi in airlift reactors have been presented by
Erickson et al. l8
Fermentations of Candida lipolyticaZOand Candida intermediaZ1on n-paraffin
in internal loop airlifts have been demonstrated and continuous cultivation of
another yeast (Candida utilis) in a concentric draught-tube airlift has been shown
to be possible.z2 A hydrocarbon fermentation in airlift of external loop design has
also been reported.14 Smart and Fowle? demonstrated the feasibility of
extended continuous cultivation (214 days) of plant cell suspensions of
Catharanthus roseus in a small (=0.010 m3) external loop airlift reactor. Success-
ful use has been made of ,very deep airlifts of the internal loop type (i.e., split
cylinder or concentric draught-tube reactors) both for municipal and industrial
aerobic sewage treatment.24 Such airlifts may be 100-200 m deep and tend to be
extremely small in volume compared to conventional activated sludge systems
due to the much higher oxygen transfer rates achieved in them. The power
economy for the deep shaft plants has been estimated at three times that of a
conventional plant.24 A value of 2.2 kg O2 transferredlkwh has been quotedz4for
the former. Our primary interest, however, is in the use of similar systems for
productive fermentations rather than in sewage treatment. Further details of
airlift reactors for sewage treatment are provided in the paper by Hines et al."
and in the CIL p u b ~ i c a t i o n More
. ~ ~ recently airlift reactors have been put to use
for hybridoma culture for monoclonal antibody production." The relatively low
shear tolerances of mammalian cells and their low oxygen demands mean that
very different reactor operation regimes must be used for these fermentations
than is conventional for microbial processes. Extensive experimental work is
needed in this area. Shear stress that may be tolerated by animal cells rangesz7
from less than 0.05 Nm-' to 500 Nm-'. It has been estimated2' that at shear
stresses of 0.05 ~ m - 'oxygen transfer rates of between 0.6 and 1.0 mmol L-' h-'
may be achieved in airlift devices.
Despite numerous successes the industrial application of airlift reactors remains
limited principally because the basic knowledge needed for their design is not
widely available in the open literature. The limited information which is
accessible frequently shows wide variations and conflicting claims. Consequently a
reliable design basis is still far from established for airlift reactors. Important
M.Y. CHISTI AND M. MOO-YOUNG
parameters such as induced liquid circulation rates for example, cannot at present
be satisfactorily predicted except in low viscosity Newtonian fluids. Notwithstand-
ing the difficulties we attempt here a more coherent picture of airlift reactors and
the main aspects of their design. Where possible, unifying concepts are
emphasized, possible causes of variations are identified and problems requiring
investigation are mentioned.

2.1 Classification
A large variety of configurations of airlift reactors have been investigated and
occasionally confused terminology is encountered in the literature. We distinguish
two basic classes of airlifts: (i) the internal loop airlifts where what would
otherwise be a simple bubble column has been split into a riser and a downcomer
by an internal baffle; and (ii) the external or outer loop airlift reactors where the
riser and the downcomer are two quite separate tubes connected by horizontal
sections near the top and bottom. Internal and external loop reactors may be
further subdivided depending on the peculiarities. ~ n t e r n a lloop airlifts, for
example, may be of the split-cylinder type (Figure 2a) or they may have a
concentric draught-tube configuration (Figure 2b). In the latter either the
draught-tube or the annulus may be gas sparged. The draught-tube and the
vertical baffle (in split-cylinder mode) may themselves be divided into vertical
sections to increase communication between the riser and the downcomer (Figure
2c). Multiple concentric draught-tubes have also been studied.= External loop
reactors (Figure 2d) have lesser variety, but several designs of horizontal
connections between riser and downcomer, particularly the top connection, may
be appropriate. Further modifications can be introduced into the headspace
region of the reactor where gas-liquid separation takes place, and, depending on
the separator efficiency the reactor performance may be significantly i n f l ~ e n c e d . ~ ~
Internal and external loop airlift reactors usually have circular cross-sections, but
rectangular and square cross-sections which have practical applications in industry
are also a definite option and have indeed been Gas sparger types,
both in the static and dynamic34 sparger classes, and their location in the riser
and/or downcomer may be altered to give different performances for different
purposes. Multiple gas injection points in the reactor in addition to the primary
sparger may also be beneficial. In addition, the downcomer and/or riser may
contain internals such as sieve plates" and baffles projecting from the walls."
In summary, a myriad of possible variations on basic airlift design exist and
may be advantageously utilized for different applications. Influences of some of
the design modifications on performance are examined in later parts of this
review.

2.2 Reactor Hydrodynamics and Flow Regimes


The hydrodynamics of multiphase flow in gas-liquid contactors have a controlling
influence on both the interphase and bulk transport phenomena and on the
fluid-wall heat transfer. Beginning with the introduction of gas into a reactor and
AIRLIFT BIOREACTORS

RISER -

FIGURE 2 Airlift reactor types: (a) split cylinder internal loop; (b) concentric draught-tube internal
loop; ( c ) concentric draught-tube (vertically split) internal loop; (d) external loop.

as gas flow is increased, several different flow regimes may be observed. At low
gas inputs the gas bubbles rise almost straight up the reactor with little interaction
between them. This is the bubble flow regime known also as unhindered bubble
flow and homogeneous bubble flow (Figure 3a). With further increases in the gas
flow rate the bubble density in the fluid gradually rises and there is greater
interaction between them-bubble collision frequencies increase-and a coal-
esced bubble flow regime characterized by greater turbulence ensues as shown in
Figure 3b. This is a transitional regime which leads eventually to fully developed
churn turbulent flow where, in addition to many small bubbles, larger bubbles
occur frequently. Because of the very high turbulence fields the large bubbles
often have little definition to their shape which fluctuates quite randomly. Figure
3c depicts churn turbulent flow. Spherical caps or bullet nosed bubbles which rise
rapidly may form at higher gas flows. Frequency and size of spherical caps
M.Y. CHISTI AND M. MOO-YOUNG

0 1 BUBBLE FLOW b ) COALESCED


BUBBLE FLOW

(Smoll dio. I Lorge d i o .


column I column 1

C) CHURN-TURBULENT d l SLUGGING

FIGURE 3 Reactor flow regimes.

increases with gas flow rate and in reactors with small diameters, particularly in
viscous fluids, these bubbles may attain dimensions approaching those of the tube
through which they rise and this is the fully developed slug flow shown in Figure
3d. Annular film flow and spray flow can be obtained at very high gas velocities,
but these are of little interest for aerobic fermentation purposes. The range of gas
flows over which a particular flow regime occurs and whether it exists at all in a
given application is dependent on the properties of the fluid being handled and on
the physical configuration of the reactor. Earlier transition to slug flow occurs, for
example, in tubes of small diameter than in vessels of larger size; and spherical
cap bubbles form more readily in highly viscous fluids and mycelial media than in
water-like systems. Identification of the flow regime existing in a reactor is
simplified by the use of flow regime maps two examples of which are shown in
Figure 4 for air-water. Figure 4a is more suitable for bubble columns with no net
liquid flow, while 4b may be applied to the riser or downcomer of an airlift when
the gas and liquid flows through these sections are known, and to bubble columns
with large fluid throughputs. Other similar maps, more or less suited for
particular purposes, are available in texts and handbooks on multiphase flow.
The hydrodynamic behaviour of bubble columns and airlift reactors is very
different. The main distinction between cocurrent or counter-current bubble
columns and airlift reactors is that in the latter the rate of liquid circulation
depends on, and is determined by, the gas flow rate, whereas in the former types
the liquid flow is independent of gas flow. Because of the long residence times
that are typically necessary in bioreactors, large liquid throughputs are not
possible in bubble columns without significant recycle rates. In airlift reactors, on
the other hand, quite high linear liquid velocities may be generated without the
AIRLIFT BIOREACTORS

HOMOGENEOUS
I BUBBLY I

FIGURE 4a Approximate dependency o f flow regime on gas velocity and reactor diameter for
bubble columns with no liquid flow (water and dilute aqueous solutions). Based on Shah el 01."

/
/
/

ANNULAR WISPY-
;" 10' I ANNULAR
E I
CHURN 7------
---I
CI,
N YI
,
, I
I
\ BUBBLY
\
\

/
/
BUBBLY -SLUG
'
\

FIGURE 4b Flow pattern map for vertical gas-liquid flow for low viscosity Newtonian liquids.
Adapted from Collier."

need for any external recirculation mechanism. The consequent turbulence


postpones the incipient slugging in the airlifts to higher gas velocities than is usual
in bubble c o ~ u m n s . ~Also,
~ , ' ~ the gas velocities for liquid blow out condition are
lower in the bubble columns than in airlift reactors. As a result of these effects
the operating range of airlift reactors in terms of the possible gas and liquid
M.Y. CHISTI A N D M. MOO-YOUNG

AIRLIFT LOOP REACTORS

C ,f-BUBBLE COLUMNS

FIGURE 5 Operating ranges of gas and liquid velocities in bubble column and airlift reactors. Based
on Weiland and O n k e r ~ . ) ~

superficial velocities in them is broader3"han for bubble column reactors (Figure


5).

2.3 Power Inputs Into Pneumatic Devices


The degree of turbulence in a reactor is a function of the power imparted to the
fluid in addition to the momentum transport characteristics of the fluid itself. The
power input to a pneumatic reactor containing a batch of liquid results solely
from the gas fed to it. The work done during the isothermal expansion of the gas
as it moves up the reactor and the kinetic energy transferred to the fluid by jet of
gas entering the reactor are the two contributors to the total power input which
may be shown to be

where P,, and P,, are the pressures at the bottom and at the top (headspace
presure) of the fluid in the vessel. Q, in Eq. (1) is the molar flow rate of the gas,
Uois the gas velocity in the orifice of the sparger which relates to the gas velocity
just above the orifice through the efficiency q (generally 0.06),'"' and M is the
molar mass of the gas. The kinetic energy term in Eq. (1) (second term on the
right hand side) would generally be found to be negligible. Discrepancies in the
literature have arisen due to underestimation of power input when the kinetic
energy term in Eq. (1) was ignored without justification. Interesting debate^^'.^'
on this subject exist. It should be noted that even when the molar gas flow into a
reactor is constant the power input to the fluid declines with increasing reactor
AIRLIFT BIOREACTORS 203
headspace pressure and the reactor becomes markedly less turbulent. Thus when
pneumatic reactors are operated in a pressurized state-to improve the gas-liquid
mass transfer driving force, for example-the turbulence intensity and parameters
dependent on it (e.g. mixing, overall mass transfer coefficient, and gas holdup)
are sacrificed to various degrees. Pressurized operation, however, is less common
for biotechnological processes because of the added costs and problems as-
sociated with carbon dioxide toxicity.

2.4 Gas Velocity


All the main indicators of reactor performance: gas holdup, mixing or blending
time, axial dispersion coefficients for gas and liquid phases and the overall mass
transfer coefficient, have commonly been correlated with gas throughput in
pneumatic devices and are discussed in other sections of this review. Here we
comment on the nature of gas velocity and its associated terminology.
The term "superficial" gas velocity remains the source of some discrepancies in
the literature on airlift vessels. The superficial gas velocity in airlifts may be based
either on the cross-section of the riser or on the total cross-section, i.e. riser and
downcomer, of the reactor. The superficial velocity based on riser cross-section is
the only one which has real physical meaning in airlift devices. However, when an
internal loop airlift is compared with an equivalent bubble column (same outer
tube diameter or equivalent hydraulic diameter, same liquid and unaerated liquid
height) then it is incorrect to base the gas velocity in the airlift on riser cross-section
and that in bubble column on the entire column cross-section. Instead, both
systems should be compared on the identical basis of superficial velocities based
on the entire cross-sectional area of the column. Only under this situation is the
equal power input in the two reactors correctly calculated. In fact, in general
reactors should be compared only on the basis of equal power inputs per unit
total liquid volume in the vessel.
The other important consideration in evaluating gas velocities is due to the
axial variation in volume flow of gas in a reactor because of changes in the
hydrostatic pressure. The mean superficial gas velocity for a reactor with
unaerated liquid height L, operating at headspace pressure Ph (usually atmos-
pheric) is given by:

where T is the reactor temperature and A its cross-sectional area. From Eqs. (1)
and (2) it can be shown that for a reactor of uniform cross-sectional area the
power input per unit liquid volume is

which applies to airlifts when Us, is based on the total riser and downcomer
cross-sectional area. If either the superficial mass velocity of gas or its velocity
under reactor inlet (or outlet) conditions of temperature and pressure are used in
M.Y. CHISTI AND M. MOO-YOUNG

1 r G =4.69x I O - ~ ( IP ~V ~ ) ~ ( . b ~ ~ ~

7 A
Headspace
o 145
101 k P o
Pressure

187 kPo

FIGURE 6 Gas holdup in a draught-tube internal loop airlift reactor in air-water at diflerent
reactor headspace pressures. The apparent reduction in gas holdup with increasing pressure seen in
(a) disappears ( b ) when true superficial gas velocity (Usg) is used for correlation instead of the
superficial gas velocity under inlet conditions as in (a).

correlating gas holdup or gas-liquid mass transfer, for example, then apparent
effects of reactor headspace pressure or of unaerated liquid height on these
parameters may seem to exist. An example of this is shown in Figure 6 which
presents our gas holdup data obtained in air-water in a large concentric
draught-tube airlift vessel (d, = 0.762 m, dci= 0.355 m, total height of draught-
tube = 2.06 m (split vertically into 3 sections: from top to bottom 0.434,0.914 and
0.559 m, respectively, separated by gaps of 0.077 m), draught-tube clearance from
bottom of reactor = 0.202 m, unaerated liquid height = 2.32 m; annulus sparged
by perforated pipe (total holes = 128, hole diameter = 0.002 m) ring (3 concentric
rings) sparger) at reactor headspace pressures (absolute) of 101, 145 and 187 kPa.
In Figure 6a the total holdup obtained by the level displacement technique is
shown as a function of superficial gas velocity at reactor inlet conditions and an
apparent pressure effect on holdup is seen. The same data plotted against actual
superficial gas velocity (Eq. (2)) in Figure 6b all falls on the same line. The
superficial air velocities in Figure 6 are based on the entire cross-section of the
airlift reactor.

3. GAS-LIQUID DISPERSIONS

3.1 Gas Sparger Types and Properties


Either static or dynamic gas spargers may be used for gas injection into airlift
reactors. Perforated plates and pipes, porous plates and single hole spargers are
examples of the static type. Dynamic spargers are exemplified by injector nozzles
and venturies which disperse gas due to the kinetic energy of a liquid jet.34 Figure
AIRLIIT BIOREACTORS
STATIC SPARGERS

gas
POROUS PLATE SINGLE NOZZLE PERFORATED PLATE
OR PERFORATED PIPE

DYNAMIC SPARGERS
I

INJECTOR NOZZLE EJECTOR NOZZLE VENTURl SLOT


INJECTOR
FIGURE 7 Gas spargers.

7 shows some of the spargers that may be used. The choice of sparger for a
particular application depends on several factors. Perforated plates and pipes are
cheap to install and operate. Porous plates are more expensive and have higher
operating costs due to greater pressure drops through them. In addition they are
prone to blockages and can be a source of contamination. We have, however,
successfully used porous plate aerators (100 p m pore size) in small (=2 L) airlift
devices during cultivation of mammalian cells for monoclonal antibody produc-
tion. Cell growth on aerator surface or into the pores posed no problems.
Dynamic gas spargers are not commonly used in fermentation practice. They are
more complex to design and build and require some external liquid circulation
mechanism, usually a pump. High shear rates in the dynamic spargers and in
associated pumping machinery rule them out for shear sensitive microorganisms
and tissue cultures. Dynamic sparger design is briefly treated by Z l ~ k a r n i k . ~ '
Apart from those considerations the required gas bubble size in dispersion can
also be influenced to some degree by the type of gas sparger used.
206 M.Y. CHlSTI AND M. MOO-YOUNG
For porous plate spargers operating in air-water the effect of various degrees
of sparger blockage on gas holdup may be computed" from:

where f is the fraction of total free sparger area relative to the cross-sectional area
of riser, and Ub is a characteristic bubble rise velocity. The mean bubble rise
velocity may be estimated using
+
Ub = 0.284 2.7U, (5)
(ms-') (ms-') (ms-')
Equation (5) is applicable to bubbles rising in swarms in air-water dispersions in
bubble columns. This equation was obtained by us using gas holdup data from
several sources including our own4' and it described all the data within f 2 0 %
over a gas velocity range of 0.005 to 0.4 ms-I. In the risers of airlift reactors
bubbles rise faster than in bubble columns due to liquid circulation and the liquid
velocity should be added to the value obtained from Eq. (5) to get the correct rise
velocity. More accurate predictions of mean bubble rise velocity are possible
using the equation

which applies to bubble flow regime (Us, < 0.05 ms-') and

for coalesced bubble flow (U, > 0.05 ms-I). Equations (6) and (7) are for bubble
columns. The use of these equations in Eq. (4) requires an iterative procedure.
Another correlation which may be suitable for the estimation of mean bubble rise
velocity in the riser of an airlift reactor is
+
Ub = 0.24 + 1.35(Usg UL,)0.93
(ms-I) (ms-I) (ms-I) (8)
Equation (8) was obtained by Hills46 for air-water in a 0.15 m diameter bubble
column for UL, > 0.3 ms-'.

3.2 Dispersion Characteristics


Under the highly turbulent conditions typically found in bioreactors the bubble
size in dispersion is generally independent of the size at birth and it is controlled
by the equilibrium between the dynamic pressure forces which work to break the
bubble and the surface tension force which attempts to preserve its size and
shape. There are exceptions, however. In strongly non-coalescing media aerated
with porous plates, for example, the very fine bubble size originating at the
AIRLIFT BIOREACTORS 207
sparger persists in dispersion. In homogeneous power law fluids the mean bubble
size in gas-liquid dispersion may be obtained4' using

which is said to apply when the scale of the energy containing primary eddies is
much greater (>200-fold) than the Kolmogoroff scale of energy dissipating
terminal eddies. In Eq. (9) V I is not the liquid volume in the reactor, but it is the
volume contained in the riser. Another expression for bubble size is

which was reported by Azbel."' Dussap and Gros4' proposed the following
expression for gas-liquid interfacial area in aqueous sodium sulphite in a
concentric cylinder airlift (volume = 0.015 m3, d, = 0.11 m, dci = 0.0756 m, total
height = 1.8 m):

which applied for approximately 200 5 PG/VD ( W ~ I - 5


~ )1500. Equation (11)
predicts a Sauter mean bubble diameter in the form:

Volume-surface or Sauter mean bubble diameter is the diameter of a sphere with


the same surface to volume ratio as the bubble. In gas-liquid dispersions where a
distribution of bubble sizes exists, the Sauter mean bubble diameter is given by

where ni is the frequency of occurrence of bubbles with diameter dB;.


In non-Newtonian carboxymethyl cellulose solutions containing 0.8 k m ~ l r n - ~
sodium sulphate (n = 0.440-0.697, K = 7.683-0.095 Pas") Godbole et al."
recommended:
a, = 19.2@: p ~ $ ~ ~
(m-') (ms-') (Pas)
for U, < 0.25 ms-' in bubble columns. Under identical conditions these inves-
208 M.Y. CHISTI AND M. MOO-YOUNG
tigators correlated their gas holdup data by the equation

EG = 0 . 2 5 5 e E p;l9
(ms-') (Pas)
Substitution of the relation

and Eq. (15) into (14) yields the following equation for Sauter mean bubble
diameter:
dB = 0.08Gsl3 ;::p
(17)
(m) (ms-') (Pas)
The apparent viscosity in Eq. (17) was determineds0 from

where K and n are the consistency and flow behaviour indices, respectively, of
power law fluids. When Eq. (18) is substituted into (17) and the result is
expressed in terms of the power input using Eq. (3) we get equations of the form

where (Y and /? are 0.096 and -0.19, respectively for the solution with
K = 7.683 Pasn and n = 0.440. Similarly for the fluid of K = 0.095 Pasn and
n =0.697, a and /? are 0.007 and -0.04, respectively. A much weaker
dependence of bubble size on power input is indicated in pseudoplastic
homogeneous fluids than in salt solutions.
The nature of gas-liquid dispersion of water, salt water, and CMC-in-water
solutions has been studieds1 in a split-cylinder (Figure 2a) airlift vessel (d, =
0.15 m, A,/A, = 0.5, height = 1.36 m, perforated plate sparger with 38 holes of
0.0016 m diameter, baffle clearance from the bottom = 0.055 m). Results showed
that for air-water system in the riser the bubble size distribution was dependent
on position along the vertical a x k 5 ' The bubble size distribution in air-water was
observeds1 to shift gradually toward the small bubble region with increasing
distance from the sparger; furthermore, the greater was the distance from the
sparger the more uniform was the bubble size distribution. This behaviour was
explaineds1 by the continuous breakage of primary bubbles as they left the
sparger and were exposed to the turbulent flow field. The data of Glasgow et ~ 1 . ~ '
indicated that at a fixed distance up the riser the bubble size distribution was
influenced by superficial gas velocity (or the power input). Increased power inputs
lead to smaller bubbles; however, coalescence also became important at higher
bubble densities and tended to offset bubble breakup resulting from the increase
in dynamic pressure forces. The dispersion behaviour in CMC solutions (0.5 and
0.8%, K and n not given) was very different from that in air-water dispersions.
AIRLIFT BIOREACTORS
In the viscous systems the survivability of large bubbles in the upflow was
e n h a n ~ e d . ~Which
' was explained as being due to increased coalescence and
decrease in available disruptive turbulent energy at a given gas flow rate. Our
observations have shown that the distribution of bubble size in the downcomers
of airlift reactors is narrower than in the risers. In the downcomer, too, there is
an axial gradation of bubble size. This is discussed elsewhere in this work.
Notice that nearly all the empirical equations for gas holdup dependent
parameters such as interfacial area and bubble diameter presented in this section
followed the general form: parameter = constant . Usg? Similar dependence of
gas holdup on superficial gas velocity (or specific power input) has been well
known, although it was only recently that some theoretical insight into this
dependence became availab~e."~
At present there is insufficient knowledge for reliable estimation of gas-liquid
interfacial areas and bubble sizes in homogeneous power law fluids and even less
so in broths containing solids which simulate mycelial suspensions.

3.3 Sparger Location and Flow Patterns


In keeping with the normal practice in bubble columns, the gas sparger in airlift
reactors is usually located at the base of the riser. This is not the best location,
however, and the recirculating fluid flowing from the downcomer over the gas
stream emerging from the sparger causes a maldistribution of the gas to the far
wall of the riser as shown in Figure 8a and b for internal and external loop
reactors, respectively. The gas thus concentrated along the wall of the riser
coalesces into large bubbles and channels up the wall for some distance before it
is redispersed by turbulent forces over the entire cross-section of the riser. The
channeling distance is usually no more than 1 m from the base of the riser except
when the gas flow rates, and hence the fluid turbulence, are quite low
(U, < 0.005 ms-'). In small reactors, which are not uncommon for industrial
production of speciality products, the poor gas distribution has a marked
deleterious affect on performance. Even in tall airlift vessels when low gas flow
rates are used as, for example, in animal tissue culture and plant cell fermenta-
tions, the effect of gas distribution is still significant. Our observations have
shown that gas can be better distributed by placing the sparger just inside the
riser (Figure 8c and d). By proper filling in of the dead zone, by plastic inserts in
our work, very smooth flow over the bottom of the reactors was obtained as
shown in Figure 8c and 8d. No problems with solids settling were encountered
with the redesigned reactor bottoms even for solid contents as high as 30 kg dry
solids per m-3 suspension for Solka-Floc cellulose fibre solids which
s i m u ~ a t e d ' ~the
. ~ ~behaviour of pulp form of mycelial fermentation media very
well.
An unusual mode of airlift reactor operation is that in which nearly all the gas
is injected at some location in the downcomer instead of at the base of riser
during most of the operational cycle. This is the aeration technique used in the
Imperial Chemical Industries' 100-300m deep airlift reactors described by
Kubota et Because the sparger is located about halfway or a greater distance
M.Y. CHISTI AND M. MOO-YOUNG

GAS GAS
(a (b)

t
GAS
t
GAS
(c) ( d)

FIGURE 8 The influence of sparger location on gas distribution in airlift reactors: poor distribution
o f gas in internal (a) and external (b) loops. Proper sparger positioning (c and d) for improved gas
distribution. Hatched areas indicate filled-in zones to improve liquid flow and to prevent biomass
settling.

above the base of the downcomer the hydrostatic head on it is lower than if it
were positioned conventionally and hence the power demands are correspond-
ingly low. In the operation of this type of reactor initially all gas is injected at the
base of the riser until a high rate of fluid circulation has been established, then
most of the air flow is gradually transferred to the downcomer sparger. The
hydrodynamic metastability of the flow pattern produced in this manner can be a
serious problem5hnd flow reversal (i.e., downcomer becoming the riser) may
occur leading to operational difficulties. In relatively small sizes of airlift reactors
usually examined in academic institutions stable fluid circulation during down-
comer gas injection is extremely difficult to obtain and maintain and consequently
this mode of operation has not been investigated. Kubota er aLS2have proposed
an algorithm, based on the balance of hydrostatic and frictional pressure drops,
for predicting the gas injection rate necessary to maintain liquid circulation in
these reactors. An additional drawback of this form of operation is that even for a
low oxygen demand gas throughput may not be reduced sufficiently because it is
dictated by the need to maintain a stable liquid circulation. Consequently, high
mass transfer economies would occur only when a large oxygen demand exists.s2
Another novel strategy, not examined so far, would be to position the gas sparger
some distance up the riser. This would reduce power consumption and the ease of
startup and stable operation would remain. The region of the riser below the
AIRLIFT BIOREACTORS 211
sparger could be aerated either by recirculating gas from the downcomer or by
supplementary spargers located either in the downcomer or further down in the
riser.

3.3.1 Substrate injection points The problem of location of the substrate feed
points in an airlift vessel is similar to that of gas injection. For rapidly utilized
substrates, the concentration of which must be kept low for reasons such as
substrate toxicity or substrate inhibition, the microorganisms in a tall airlift may
be starved of the substrate only a short distance downstream of the point of
substrate i n j e ~ t i o n . ~Thus,
' multiple substrate feed points may be necessary
axially up a reactor if product yield reduction due to substrate starvation is to be
avoided. The recent work of Fields and SlaterS5which involved cultivation of the
bacterium Methylophilus methylotrophus on methanol further highlighted this
point. The substrate balance for a differential volume of the riser may be written
as:

where S is the substrate concentration at any vertical position x, UL is the true


liquid velocity, EL the axial dispersion coefficient of the liquid phase and Rs the
rate of substrate consumption. When the substrate concentration must not fall
below a critical minimum value S,,, and it should not exceed a maximum of S,,,,.
because of inhibition considerations, then Eq. (20) may be solved with appropri-
ate reaction kinetic expression to determine the axial distance at which fresh
substrate addition becomes necessary. In highly aerobic fermentations in tall
airlift bioreactors a similar consideration will apply to oxygen injection. In fact
the possibility exists of near complete oxygen depletion close to the bottom of a
tall d o ~ n c o m e r . ' ~Multiple oxygen injection points in the riser and, to some
extent, in the downcomer may be needed to overcome this. The location of these
supplementary injection points will depend primarily on the rate at which oxygen
is depleted by the microorganisms and on the velocity of liquid circulation in the
reactor. A practical application of the knowledge of liquid circulation in airlifts is
immediately obvious. The ability to predict liquid circulation is of definite
importance in design.

4. FLOW VISUALIZATION STUDIES

In addition to the general description of flow regimes given in an earlier section a


more detailed examination of flow is required for better understanding of
reactors. Flow visualization studies such as those in classical stirred reactors have
provided tremendous insight into their functioning. Circulation patterns, impeller
flooding, impeller velocities for suspension and complete dispersion of solids,
identification of dead zones and regions of intense turbulence, optimum impeller
location for flow distribution and numerous other details were obtained on the
basis of visualization studies. Similar observations in airlift reactors are lacking.
M.Y.CHISTI AND M. MOO-YOUNG

FIGURE 9 Rectangular internal loop airlift vessel. All dimensions in metres.

The few existing studies (e.g. 57) have treated only the overall pattern of fluid
circulation mostly in water or salt solutions. Here we report some of our
observations in aqueous solutions (water, 0.15 kmolm-' NaCI) and non-
Newtonian media (CMC solutions K = 0.03-0.11 Pas", n = 0.68-0.84, a = 80-
83mNm-') in an internal loop airlift of rectangular cross-section (Figure 9)
sparged with air via a perforated plate (20 holes of 0.001 m diameter) located at
the bottom of each riser. Influence of sparger location on flow behaviour has
already been discussed and will not be repeated here.
At very low gas flow rates the bubbles rose in the riser only and the rate of
liquid circulation was quite low. As the gas velocity increased the liquid
circulation rate also increased and some gas bubbles were dragged by the liquid
into the downcomer. The bubble size in the downcomer was lower than in the
riser since the liquid flowing into the downcomer was incapable of sweeping
larger gas bubbles with high rise velocities into it. The depth of penetration of a
bubble into the downcomer depended on its size and the tiniest bubbles
penetrated the deepest. Having attained a certain penetration depth the gas
bubbles remained more or less static and in this condition the buoyancy force on
the bubble was exactly balanced by the downward drag force due to the flowing
liquid. As the gas velocity increased further, progressively larger bubbles were
also swept into the downcomer and the depth of bubble penetration in it
increased. Eventually a recirculation of most bubbles, i.e. complete flow through
the downcomer, was obtained. Throughout, the bubble size distribution in the
AIRLIFT BIOREACTORS
downcomer was narrower and the mean bubble size smaller than in the risers.
The extent of liquid turbulence in the downcomer was also significantly lower.
There were two reasons for this: (i) some energy was used up in forcing the
bubbles against the buoyancy-dictated direction of flow, and, (ii) the net loss of
energy from the liquid to the gas which was used up in gas compression. Both
riser and downcomer turbulence increased with gas flow. The residence time of
gas bubbles in the downcomer was longer than their residence time in the riser.
The bubble density was lower in the downcomer than in the riser and because of
this bubbly flow existed in the downcomer even when the flow in the risers had
changed to coalesced bubble or churn-turbulent type.
As the liquid exited the downcomer and turned into the risers the downward
component of its velocity reduced so much so that the bubbles were not ejected
to any significant depth in the region between the end of the downcomer and the
reactor base. Instead most of the bubbles which came out of the bottom of the
downcomer entered the riser just under the edge of the downcomer baffle.
Consequently a small zone of the reactor located mainly between the lower end
of the downcomer and the reactor base remained largely free of bubbles (Figure
10a). When the downcomer clearance from the reactor base was increased the
volume of gas free zone also increased. Based on these observations we
recommend that the total cross-sectional area for fluid flow from the downcomer
to the riser, i.e. the area just below the baffles separating the downcomer and the
riser in internal loop devices or the cross-sectional area of the bottom horizontal
connection in the external loop units, should not exceed 1.65 times the
downcomer cross-section. This is not so critical for external loops, however.
The reactor shown in Figure 9 was designed as a two dimensional device-a
section through a concentric draught-tube vessel. Stagnant zones were observed
in the corners on the base of the reactor particularly at low gas flow rates. Filling
in of the corners eliminated this problem. A plastic prism was found to smoothen
the turn around of the liquid from the downcomer into the riser. This is
recommended in annulus sparged concentric draught-tube internal loops in which
it takes the shape of a cone as was used in the work of Kawase and ~ o o - Y o u n g . ~ *
The cone should be used even in reactors with otherwise elliptical bottoms. It is
interesting to notice that concern is being expresseds9 about the shapes of the
bottoms of stirred tanks also. The geometric parameters of stirred tank bottoms
so successfully employed by ChudacekS9 may be used for the concentric
draught-tube airlift vessels by replacing the propeller diameter with the diameter
of the draught-tube. Figure lob shows the liquid circulation currents which were
visualized by following the movement of gas bubbles and by the injection of dye.
In the risers there was some liquid downflow and circulation near the walls, but
this was never as strong as has been observed in bubble columns with batch liquid
under identical conditions where large circulation cells have been claimedm to
exist. Figure 10c shows the flow cellsm in a bubble column for comparison with
Figure lob as observed in our airlift. Studies with dyes showed the backmixing in
the riser to be far stronger relative to the downcomer where the flow
approximated to plug flow except near the downcomer entrance where there was
strong mixing. Good mixing was also observed in the reactor headspace region
M.Y. CHISTI AND M. MOO-YOUNG

STRONG
- ~
DOWNFLOW
N E A R WALLS - - -,,
LlOUlD CIRCULATION
IN DOWNCOMER ENTRANCE

EDDIES FLOWING DOWN


N E A R THE W A L L S

- Z O N E O F LlOUlO CIRCU-
LATION I SPARGER EFFECT)

FIGURE 10 Flow visualizations in the rectangular airlift: (a) gas free region below the downcomer;
(b) liquid flow pattern in the reactor; (c) chaotic circulation cells in a bubble column.

particularly above the downcomer where the fluid streams from the two risers met
and reversed their flow direction. As will be seen later this was in keeping with
our observation, confirmed also by others," that mixing times in concentric
draught-tube airlifts decline with increasing liquid volume in the headspace
region.
In CMC solutions the general flow was the same as that described for water and
salt solution. There were important differences, however. Under otherwise
identical conditions the turbulence intensity in any zone of the reactor was less
than was observed for water. The initial bubble size from the sparger was slightly
larger in CMC solutions than in water or salt water. Even at very low gas flow
rates the bubble coalescence in the riser lead to the formation of large bubbles
(>0.025 m in diameter) which increased in size still further by coalescence as they
moved up the riser. Many very small bubbles were also present. The bubbles
from the spargers were displaced toward the far walls (Figure 8a) of the risers due
to the fluid flowing from the downcomer. However, the intense circulation zones
observed at the entrances of the riser in water (Figure lob) were now so much
calmer that even very small gas bubbles rose straight up through them instead of
circulating. Bubbles which moved down the downcomer were generally smaller
than in water. Unlike in water or salt water however, bubble coalescence
occurred in the downcomer and the bubble size increased as the bubbles flowed
down until it became so large that the direction of flow of bubble reversed and it
moved up the downcomer against liquid flow. Further entrappment of smaller
bubbles increased their size even more until small spherical caps (0.02-0.03 m
AIRLIFT BIOREACTORS 215
diameter) formed. The latter did not rise very stably in the centre of the down-
comer due to the higher liquid velocity in the centre and consequent higher resis-
tance to upflow of bubbles, and they moved towards the walls of the downcomer
and rose up along the walls. As the gas velocity in the riser increased further large
spherical caps formed in it. In the downcomer, too, the size and frequency of the
spherical caps increased. However, never did these bubbles attain the dimensions
of the riser or the downcomer and fully developed slug flow was not obtained.
With increasing gas flow the depth of the downcomer where the formation of
spherical caps commenced moved progressively downward. The upper entrance
region of the downcomer was quite turbulent and as a result when the spherical
caps moving up the downcomer reached this zone they were broken up into
smaller bubbles once again and carried down with the flow. Periodically one or
two spherical cap bubbles escaped the downcomer via its entrance. The reactor
became increasingly turbulent at higher gas velocities. These observations pointed
to the distinct possibility, particularly in highly viscous media, of oscillation of
liquid circulation due to large spherical cap slugs momentarily lodged in the
downcomer of an airlift reactor. T o avoid this the minimum downcomer diameter
should exceed the value in which fully developed slugging is possible. An
equivalent hydraulic downcomer diameter of more than 0.1 m may be satisfac-
tory. The maximum stable size of spherical caps in the riser was apparently lower
at the higher gas flow rates due to greater turbulence. In viscous systems the gas
free region between the end of the downcomer and the reactor base (Figure 10a)
which was observed in water no longer existed. This was because gas bubbles now
experienced a greater drag due to higher fluid viscosity and penetrated this zone
with liquid eddies. More violent agitation of the gas-liquid dispersion surface was
noticed in viscous systems due to frequent bursting of the spherical cap bubbles
on the surface. In CMC solution it was noticed at all gas flows that near the inside
edges of the downcomer entrance (Figure 11) there was a zone of trapped gas
which grew in size with gas flow and, despite high turbulence, lead to rapidly
fluctuating separation of liquid flow from the downcomer wall.
The foregoing flow visualization studies covered a very broad range of
superficial gas velocities: 0.01-0.40 ms-' based on the cross-sectional area of the
risers.

LARGE GAS SPACE


FLUCTUATES RAPIDLY
DUE TO TURBULENCE

I I DOWNCOMER

FIGURE 11 R o w separation and gas attachment to the inside edge of the downcomer entrance in
the rectangular internal loop at (a) low, and (b) high gas flow rates.
216 M.Y. CHISTI AND M. MOO-YOUNG
5. REACTOR DESIGN PARAMETERS

The design and performance evaluation of gas-liquid-solid reactors depends on a


knowledge of the following principal parameters: gas holdup, gas-liquid interfa-
cial area, overall volumetric mass transfer coefficient, mixing or blending time of
liquid, liquid circulation rate, liquid and gas phase axial dispersion coefficients
and fluid-wall heat transfer coefficients. In particular cases liquid-solid and
intraparticle mass transfer become important.
The experimental techniques used to gain information on gas-liquid reactor
design parameters are well known and will not be treated here except in those
instances where the specificities of airlift reactors require significant modifications
to them. Commonly used experimental procedures have been discussed by
Schiigerl,6' and Y ~ s h i d a ; ~chemical
' methods of measurement of gas-liquid
interfacial area and mass transfer properties have been examined in detail by
Sharma6' and other^;^ hydrodynamic parameter measurements have been
described by Nottenkamper er and Cheremisinoff," and tracer methods were
examined by Shah et Some general aspects of bioreactor design were
presented by Moo-Young and Blanch" and more recently by Kargi and
~oo-~oung.~'

5.1 Gas-Liquid Mass Transfer and Gas Holdup


The rate of mass transfer of a gas such as oxygen or carbon dioxide from gas to
liquid phase (or vice versa) may be expressed in terms of a true mass transfer
coefficient k,, the specific gas-liquid interfacial area a, through which the
transfer occurs, and a concentration driving force for the transfer:

where C, is the concentration of the transferring component in the liquid phase


and C* is its maximum possible concentration (i.e. saturation or equilibrium
value) in the liquid. Equation (21) is written for gas diffusing into liquid. Mass
transfer performance of gas-liquid reactors including airlift devices has frequently
been stated in terms of an overall volumetric mass transfer coefficient kLaL or
kLaD, based on unaerated liquid volume and gas-liquid dispersion volume,
respectively. All direct interfacial area (a, or a,) measurement techniques have
well known shortcomings and most of the directly obtained data on interfacial
areas is subject to severe criticism; making the calculation of k, from readily
measured kLaL and experimental a, a less reliable approach. On the other hand,
the use of overall coefficients, kLaLor kLaD, and their equivalents means that all
the information concerning both the kinetics of the transport process and the
interfacial area is combined in these coefficients and such a phenomenological
approach does not permit an evaluation of the factors which influence mass
transfer from one phase to another. An alternative approach is proposed by us
for evaluating the true mass transfer coefficient k, from the experimental kLaL
AIRLIFT BIOREACTORS 217
and gas holdup (E,) data which are both easily determined with a high degree of
confidence. Thus, it is known that for gas bubbles of Sauter mean diameter d , in
a dispersion with gas holdup E, the interfacial contact area is given by

Multiplication of Eq. (22) by k L followed by rearrangement leads to

The k J d , ratio for water and Solka-Floc (grade KS-1016) cellulose fibre
suspensions is shown in Figure 12 as a function of the superficial air velocity
(based on the entire column cross-section) in the rectangular airlift depicted in
Figure 9 and in an identical bubble column (the removal of the downcomer from
the airlift in Figure 9 and replacement of the spargers by a single perforated plate
with holes of 0.001 m diameter converted it to a bubble column). Over a wide
range of gas velocities spanning bubble flow and coalesced bubble flow regimes
the k L / d , ratio was constant in any given fluid irrespective of the type of
pneumatic contactor used (Figure 12). This revealed a direct relationship between

--',* I0
8 -
Airlift I

FIGURE 12 kJd, ratio in 0.15 kmolm-' NaCl (@), I wt./vol.% Solka-Floc (SF)(O), 2 wt./vol.%
SF(O), and in 3 wt./vol.% SF ( A ) in the rectangular airlift and the bubble column at various
superficial gas velocities.
218 M.Y. CHISTI AND M. MOO-YOUNG
the mass transfer coefficient kL and the bubble diameter. The results showed3'
that for air-water dispersions and for suspensions where the suspending fluid was
water-like in its rheology, kL could be calculated using the equation

where C, (dry wt./vol.%) is the fraction of solids in suspension. Clearly for a


bubble of given diameter, kL was strongly influenced by the solid contents of the
three-phase system. A possible e ~ p l a n a t i o nfor
~ ~this was that the solids at the
gas-liquid interface reduced the rate of bubble surface renewal by lowering the
turbulence intensity in the interfacial film. Film thickness may also have been
increased. For air-water dispersions Figure 13 presents a comparison of our
equation (Eq. 24) with that proposed by Akita and Y ~ s h i d a : ~ ~

Equation (25) was obtained by direct determination of interfacial areas and it


predicts a dLn dependence of k , on bubble diameter. It does not indicate any
influence of liquid viscosity on kL even though such influence is expected. Both
Eqns. (24) and (25) show identical dependence of k L on liquid diffusivity.
Experimentally obtained k J d , was plotted against the values predicted by Eq.
(24) for various fluids in our two contactors (Figure 14) and a good fit of the data
was observed.
The effects of liquid viscosity on kL in solid-free .liquids have been examined in
detail by Erickson e t al. I n

FIGURE 13 Variation of true mass coefficient k , with bubble diameter d , in air-water dispersions
according to (a) Eq. (24), and (b) Eq. (25).
AIRLIFT BIOREACTORS

FIGURE 14 Comparison of experimental k , / d , with values obtained using Eq. (24): (@)
0.15kmolm-"aCl. 1 wt./vol.% SF ( A ) , 2 wt./vol.% SF (m), and 3 wt./vol.% SF (V)in the airlift
and bubble column reactors.

5.1.1 Overall mass transfer and hydrodynamic performance The overall mass
transfer coefficients, k,aL or kLa,, are more useful for practical reactor design
purposes. These have been studied by a number of investigators,28~31-33~38~57~58~7w79
and ourselves, among others, in external and internal loop airlifts of various
types. All the main correlations available for mass transfer and for gas holdup in
airlift vessels are summarized in Table I. Notice that apart from the gas holdup
correlation of Miyahara et al." all other equations in Table I may be reduced, for
a given fluid and reactor geometry, to simple power law type functions of gas
velocity. Similar equations were reported by other investigators.= Expressions of
this form are now recognized to have more fundamental hydrodynamic basis.45
In general, for draught-tube internal loop airlift reactors there is no influence of
either the presence of a draught-tube or of its relative area with respect to that of
the annulus on the overall gas holdup. We confirmed this by a reanalysis of the
data presented by bell^'^ for the internal loop vessels examined by Bello er aL7'
Thus, the total gas holdup in the internal loop airlifts was exactly the same as in a
TABLE I

Gas holdup and mass transfer correlations for airlift reactors

Sr. Reactor type K


No. and reference Equation Parameter range 4
-
C]
1. External loops; draught- A Air-water or aqueous salt solution
tube internal loom k a ~ +
~ = 0.76(1 2)-% (26) (0.15 kmol m-"aC1); d, oc d, = 0.152 m;
\ r.,, cn
(annulus sparged); 5-I ms-I h, = 1.8 m; U , , = 0.0137 to 0.086 msC1
bubble column. (bubble flow only); Ad/A, (external =!
Bello er a[." or in terms of power input: loops) = 0.11 to 0.69; Ad/A, (internal
loops) = 0.13,0.35 and 0.56; bubble
+z
(27) column AdIA, = 0. b

s-I Wm-'
and

EG, = 3.4 x lo-3(1+ ~ ) - ' ( p G l v D ) ~

Wm-'
2. External and internal Liquid velocity effects: As above.
loops as in 1
Bello er
(29)

for water and salt solution; and

a = 0.56 (water)
o = 0.58 (salt solution)
3. Draught-tube internal Water; sodium sulphate. glycerol and iso-
loops (draught-tube (31) butyl alcohol solutions; U,8, = 0.015 to
spar ed) Chakravarty 0.20ms-'; d,=0.10m;dc,=0.074, 0.059,
% .
ec a/. and 0.045 m; L, = 0.40 rn; L, = 0.026 m.

pL, p, in mPas; a in mNm-I; U,K, in cms-I


4. Rectangular internal loop Bubble flow: Water; aqueous salt solution (0.15 kmol m-3
(Fig. 9). Chisti era/." NaCI); salt solution 1, 2 or 3 dry
+
E~ = (1.488 - 0.496C,)@~YZ*0.075 (33) wt./vol.% Solka-Floc (KS-1016) cellulose
coalesced bubble flow: fibre; geometric details in Figure 9.
E~ = (0.371 - 0.089C,)@~M*0~0'5 (34) L
U, based on total reactor cross-section
5. External loops Chisti ef Same fluids as in 4; 0.026 5 U,8,(ms-') 5
a/. 33 (35) 0.21; d, = 0.152 m; L = 1.75 m; Ad/A, =
0.25, 0.44; sparger: perforated plate, 52
holes, d o = 0.001 m. 0
and gd

(36)
F
9
a
gd
6. Draught-tube internal Equation 37: water, pseudoplastic fluids; V1
loops; bubble columns. (37) 0.008 S U , (ms-I) 5 0.285; 0.14 5 d,(m) 5
Kawase and Moo- 0.35; 1 z n z O . 2 8 ; 0.001 5 K(Pas")S 1.22.
Youngs8 and Equation 38: Pseudoplastic fluids, water;
0.008< U,8(ms-L) < 0.084; 0.14 < d,(m) <
(38) 0.305; 1 > n > 0.543 and 0.00089 < K
(Pas") < 2.82.
TABLE I (Continued)
h)

8
Sr. Reactor type
No. and reference Equation Parameter range

7. Draught-tube internal Newtonian fluids: pL = 0.9 to 13 mPas; o = 51


loops (annulus to 73 mNm-'; DL=0.186 x to 2.42 X
sparged). Koide et m2 s-'.
applied within f 12% for: O.OO98 5 U,g(ms-') 5 0.156, based on total
column cross-section. K
3.71 x l d a p L / p L D L a 6 . 0 0x 1@ d, = 0.10 to 0.30 m;
1.18 x 1@apLa31gp:=5.93 X 10" d, =0.06 to 0.19 m; 3
L, = 0.70 to 2.10111; C)
0.471 5 d,;ld, a 0.743 L = 0.84 to 2.24 m;
d, = 0.001 to 0.004 m.
7.41 x a dJd, a 2.86 x lo-' E1
I
0.03M a E, a 0.305
8. Draught-tube internal Reactor geometry, operating conditions and %
loops (draught-tube fluids similar to those in 7. 0
sparged). Koide el 01." Spargers: (single n o d e ) do = 0.001 to
0.015 m; (perforated plate) do = 0.001 m. 7
holes; porous plate of unspecified pore size.
K
applied within f 13% for: 0
9
4
0
C
Z
0

9. Draught-tube internal Fluids: water, ethanollwater.


loops (draught-tube (41) glycerinelwater, millet jellylwater; pL =
sparged). 952 to 1168 kgm-'; p,. = 1.0 to 14.9 mPas;
Miyahara el a[." a = 34.1 to 72.0 mNm-I; d, = 0.148 m;
L, = 1.0m; Maximum h, = 1.20 m; d, =
0.0005 to 0.0015 m; AJA, = 0.128 to
and 0.808.
AIRLIFT BIOREACTORS
224 M.Y. CHISTI AND M. MOO-YOUNG
corresponding bubble column when the superficial gas velocity based on the
entire column cross-section was used for comparison. Exactly the same finding
was reported by Kawase and Moo-Youngss in non-Newtonian media, by Piggott3'
in a rectangular internal loop and by Koide et in draught-tube internal loops
for 0.471 5 dcildca 0.743. An explanation for this lay apparently in the mag-
nitude of liquid circulation. Increased liquid circulation in the airlift may have
reduced the riser gas holdup because the liquid velocity was now superimposed
on the bubble rise velocity which increased relative to that in the absence of
liquid flow; the downcomer gas holdup may have increased by a corresponding
amount because the liquid flow here was a negative influence on bubble rise.
Overall, then there was little net effect on gas holdup in the internal loop vessels.
have also been reported, however, probably due to variations in
gas-liquid separating ability of different reactor headspace configurations. In
another report:' the riser gas holdup was found independent of the relative areas
of the riser and downcomer in concentric draught-tube airlift vessels, but the
downcomer holdup varied with (A~/A,)-'.'~. The influence of induced liquid
circulation on the riser gas holdup and on overall mass transfer were reflected in
the correlations proposed by Bello el aL7' and Miyahara et aLs1 (Eqs. (29), (30),
(41) in Table I). All these equations also contained gas velocity as a parameter
but these two velocities were not independent of each other.
The difference between the downcomer and riser gas holdups is the cause of
liquid circulation in airlift devices. Bello et using data of their own and
additional data from other sources proposed the following linear equations to
relate the downcomer and riser gas holdups:

for the external loop devices, and

for the concentric draught-tube internal loops. Equations (47) and (48) applied to
air-water only. For water and for rheologically complex cellulose fibre suspen-
sions Chisti et al.33 confirmed a similar relationship:

This equation was obtained in external loop reactors identical to those used by
Bello et From these results we conclude that even though the absolute value
of overall gas holdup in external loops is fluid property and reactor geometry
dependent, the relationship between the riser and downcomer holdups is
independent of these factors. This is to be expected unless the reactor headspace
configuration is changed.
The lower downcomer gas holdups in the external loop airlifts relative to the
concentric-tube reactor^'^ were due to the geometric peculiarity of the external
loop devices in which the horizontal connection between the riser and the
downcomer allowed most of the gas to separate from the liquid and little gas was
carried to the downcomer. In general then while the gas holdups for bubble
column and concentric draught-tube internal loop airlifts tend to be similar under
AIRLIFT BIOREACTORS
0.15

BUBBLE
COLUMNS
AND
INTERNAL
LOOPS

FIGURE 15 Comparison of overall volumetric mass transfer coefficients in salt solutions in bubble
columns, concentric draught-tube internal loops and external loop reactors. Based on Weiland and
Onken."'

identical conditions, the gas holdups in external loop airlifts are usually lower. As
illustrated in Figure 15 an identical trend has been observed for the overall
volumetric mass transfer coefficient. According to Bello et a1.,72the contribution
of the downcomer volume to oxygen transfer in airlift reactors was negligible and
most of the transfer took place in the riser. This was said to be due to the lower
gas-liquid relative velocities (slip velocities) in the downcomer. Others have
made similar claims.83 Declining gas holdup and k,aL with increasing apparent
viscosities of homogeneous power law fluids have been ~n
' airlift
reactors and in bubble column^.'^^^^ Introduction of an apparent viscosity in
correlations (e.g. references 58, 73) makes them questionable for bioreactor
design purposes. This is because for many biological fluids, particularly suspen-
sions such as the fungal fermentation broths, flow indices K and n depend quite
strongly on the shear rate range used in their determination. Furthermore, the
value of the shear rate in the bioreactor should also be known if the n value
determined under given conditions of shear is to apply to it. Shear fields in
bioreactors are position dependent and the estimation of either local or global
shear rate experienced by a fluid in a reactor is not only extremely complex, but it
is beyond analytical treatment with the currently available knowledge. As Table I
shows, the effect of airlift reactor geometry on gas holdup and kLaL has
frequently been expressed in terms of the riser-to-downcomer cross-sectional area
ratio. Some investigator^^^,^^.^^ have found the outer column diameter, d,, to be a
significant influence on performance in internal loops. For example, the gas
holdup equation (Table I) reported by Kawase and M ~ o - Y o u n gand , ~ ~which was
claimed to apply also to bubble columns, contained d, as a variable. The columns
226 M.Y. CHISTI AND M. MOO-YOUNG
used by these investigatorsss were all larger than 0.14 m in diameter and for them
there is substantial accumulated evidences7-" that holdup should be diameter
independent. For otherwise identical conditions in concentric tube internal loop
airlifts we found, based on the results reported by Koide et a1.,75.76inexplicable
differences in the dependence of kLaL on reactor geometry for annulus and
draught-tube sparged modes of operation. For the latter, kLaLshowed a strong
dependence on the diameter ratio of the inner and outer tubes while the influence
of outer tube diameter was very pronounced: kLaLa dkos9 (draught-tube sparged)
versus kLaLa dzw' (annulus sparged). In a draught-tube internal loop no effect of
draught-tube clearance (L,) from the column base on gas holdup or kLaL was
observed7"or 1 / 8 5 Lh/dCi5 1. Similarly, no effect of draught-tube length, L,,
on holdup or kLaL was found75 for 5 5 LJd, s 15 in a reactor with d, = 0.140 m
and dCi= 0.082 m. Most of these geometric variables have not been studied in
enough detail to justify definitive conclusions.
It seems that while there is little influence of the area or diameter ratios of
downcomer and riser on gas holdup in internal loop airlift reactors in water, salt
solutions and other low viscosity media such as iso-propanol solutionss7 this may
not always be true in viscous systems. For example: up to 60% higher gas holdups
were obtained by W e i l a r ~ din~ ~glycerol solution (63.4 wt%), which was 14-times
more viscous than water, when the ratio of draught-tube to reactor diameter was
increased from 0.59 to 0.74. For further rise in this ratio the influence on holdup
was slight.
The gas holdup values of interest in design are the total holdup in the reactor as
a whole, in the riser and in downcomer, and those are the values most often
quoted in the literature. In reality there are positional variations in holdup.
Radial gas holdup variations in the riser of airlift reactors have been
documentedw~" in a fashion similar to that observed in bubble columns.6s Higher
holdups near the core than closer to the riser walls were found.w Axial variations
in riser and downcomer gas holdups are also known to o c c ~ r . ~ ' ~ ~ ~
It should be noted that most gas holdup expressions for airlift reactors correlate
the holdup with superficial gas velocity based on the fresh gas input to the riser.
The additional contribution of the recirculated gas is ignored. Because the extent
of gas-liquid separation in the reactor headspace varies widely depending on
design, the amount of recirculated gas also varies. This would explain at least
partly2y the large difference between gas holdups observed by various inves-
tigators in otherwise similar reactors.
Although the gas holdup in airlift reactors may sometimes be lower than in
corresponding bubble columns this does not automatically mean that kLaLis also
lower. Enhanced kLaL has been o b s e r ~ e d ~ "In some
'~ ~ ~ ' ~ cases in the presence of
draught-tubes relative to bubble column operation. This apparent discrepancy is
explicable in terms of the different bubble size distributions which occur in bubble
columns and airlift reactors under otherwise identical conditions. Further
comparison of airlifts and bubble columns occurs in the work of Bello et aLY4
Although all fermentations involve gas, liquid and solid phases, nearly all the
bioreactor studies in simulated media have been limited to gas-liquid systems. In
most cases the liquid used was either water-like or a pseudoplastic homogeneous
AIRLIFT BIOREACTORS
fluid. In a few studies which made use of suspended slurries in airlift vessels (e.g.,
95-98) the solids employed were not a simulation of microbial, particularly fungal
mycelial, solids. In suspensions which were carefully formulated to simulate
fungal fluids we large reductions in gas holdup and mass transfer in an
airlift reactor and in a similar bubble column relative to solid-free operation.
Under otherwise identical conditions the airlift produced marginally lower
holdups and mass transfer than the bubble ~ o l u m n . ~ * . ~ ~
The properties of liquid such as surface tension, density, viscosity and ionic
strength are known to affect gas holdup and overall mass transfer coefficient. The
magnitude of various influences depends on the reactor type. Gas holdup, for
example, has been found3' to be more sensitive to liquid properties in a bubble
column than in an external loop (dispersion height =8.5 m, downcomer
diameter = 0.050 m, riser diameter = 0.1 m) reactor. The gas holdup in aqueous
2-propanol in the bubble column (riser of external loop used as bubble column)
was up to 100% higher than in pure water.3s For the same two liquids the gas
content in the airlift differed only by about 20%.3s

5.1.2 Airlift reactors with internals Ancilliary internals such as impeller~,7.~'.~~


projecting baffles35 and static mixers6' have been the subject of some study in
airlift reactors with a view to improving performance. In general these devices
enhance gas holdup and mass transfer, mixing is usually reduced except when
impellers are used. We do not foresee any widespread use of ancilliary pieces in
airlift reactors because for bacterial and yeast fermentations these reactors
already have excellent performance so the additional complexity of internals is
not desired, while for highly viscous media and mycelial broths the use of
internal-particularly static one-poses problems with mixing and stagnant
zones and sometimes with shear sensitivity.
Other unusual designs of airlifts such as the s p l i t - ~ y l i n d e r ,rectangular
~~~~~~~~~
internal loop^^'^^.^" and rectangular external loops,39 and fluidized beds with
d r a ~ ~ h t - t u b e shave
" ~ ~ received
~ only limited attention. Some of these designs are
quite promising. For example, while the overall volumetric mass transfer
coefficientz1 and the gas-liquid interfacial area49in concentric draught-tube airlift
reactors were observed to increase with power input raised to some exponent
which was less than unity, the mass transfer in a rectangular internal loop was
found3u to increase with power input raised to some value greater than 1. Thus
higher mass transfer to power input ratios may be achieved in rectangular internal
loops.

5.1.3 Modelling for mass transfer and economic modelling In the experimental
determination of the overall volumetric mass transfer coefficient, kLaL, certain
assumptions must be made about the extent of mixing in the liquid (slurry) and
gas phases in the reactor. Usually, either constant composition or plug flow for
the gas phase is assumed, and liquid is considered to be either plug flow or
backmixed. In most cases, even for quite tall airlift reactors, the assumption of
fully backmixed liquid phase is satisfactory. The contribution of very small gas
bubbles to mass transfer is small, while larger bubbles rise so rapidly that their
228 M.Y. CHISTI AND M. MOO-YOUNG
composition is nearly constant. For example, it may be shown1' that for an air
bubble of 1 mm diameter in a water-like fluid the time taken for 63% of the
possible oxygen transfer to have occurred is approximately 50 s. If a bubble rise
velocity of 0.28 ms-I is assumed then the residence time of the bubble in a 5 m
tall pool of liquid is only about 18 seconds. For larger bubbles the extent of
oxygen transfer during the residence time of the bubble in the reactor would be
even lower. For the liquid phase it has been s h ~ w n "that ~ ~ if~ the
~ criterion
k,aLtc 5 2 (t, = liquid circulation time) is satisfied then the well-mixed liquid
phase assumption is correct. A fully backmixed tank-in-series model for liquid
mixing in airlifts, utilizing a different number of tanks in the riser and
downcomer, was proposed by H o et a1.I6 Computer simulation of the dissolved
oxygen profiles based on the model showed these profiles to be quite flat with
column height for airlifts up to 6 m tall.I6 Even in columns as much as 9 m tall,
more or less uniform oxygen concentration profiles were obtained.I6 Similar
results were reported in another s i m ~ l a t i o n We
. ~ ~have experimentally confirmed
quite flat steady state axial dissolved oxygen profiles in the riser of a 6 m tall split
cylinder airlift (d, = 0.243 m, AdIA, = 0.411, L, = 4.8 m, Lb = 0.102 m) in water
as well as in highly viscous cellulose fibre suspensions even for a low riser
superficial gas velocity of 0.039ms-'. Because stage to stage variation in the
dissolved oxygen level is small, there may be some justification for utilizing the,
much simpler, well mixed liquid phase in kLaLcomputations. Observations such
as those of Lin et and Onken and weilandlo4 have shown, however, that
complete backmixing does not always exist particularly in the presence of
microorganisms when oxygen consumption is involved.
Economic modelling of airlift reactors is mostly lacking. For high mass transfer
performance tall airlifts have been r e c ~ m m e n d e d . ' ~ ,Moresi13
~' modelled the
oxygen mass transfer in a draught-tube airlift based on perfect mixing of liquid
and plug flow in gas, and for oxygen consumption appropriate to a continuous
whey fermentation by Kluveromyces fragilis. Operating costs were concluded to
decline asymptotically with increasing aspect ratio (Lld,) of the airlift. For
fermentation volumes s 100 m3, aspect ratios greater than 30 led to lower
operating costs, while for volumes ranging from 250 to 1000m3, the optimal
aspect ratio decreased to 15.13 As far as power consumption per unit of oxygen
transfer was concerned, the resvonses of the model were influenced mainlv. bv. the
equations used for estimating k,a,. Operation of the airlifts at headspace
pressures higher than atmospheric was found less expedient than atmospheric

5.2 Mixing and Liquid Circulation


The nature and the degree of mixing are important in airlift reactor design.
Mixing may be characterized in terms of a mixing time, or as an overall liquid
phase dispersion coefficient in the reactor, as liquid dispersion coefficients in
various regions-riser, downcomer and head-space--of the reaction vessel, and
by the liquid circulation velocity. One or more of these indices of mixing may be
suitable for a particular purpose. For example, mixing time for 95% mixing of a
AIRLIFT BIOREACTORS 229
tracer pulse may be more appropriate for pH control purposes; liquid phase
dispersion coefficients would determine the quantity and strength of an acid
pulse, for example, so that local pH values do not rise so high as to damage the
microorganisms, reactants or products; and for toxic and inhibiting substrates the
location of feed points in the reactor would be influenced by axial and radial
dispersion in the liquid. Velocity of liquid circulation impacts upon mixing and
liquid phase dispersion, affects gas holdup, mass transfer and the extent of shear
in the reactor.
The nature of mixing phenomena in airlift reactors is quite different from that
found in bubble columns. This is clearly revealed by the typical responses to a
pulse tracer input observed in a batch liquid bubble column and external or
internal loop airlift devices. As shown in Figure 16, the mixing in bubble columns
is purely dispersive; whereas a definite liquid recirculation is superimposed on
mixing by dispersion in airlift reactors which leads to the characteristic decaying
sinusoidal tracer response pattern shown. The calculation of mixing or blending
time for 95% difference between the final and instantaneous tracer concentrations
can be done directly from the tracer response curve in a very straightforward
manner. Similarly, the circulation time-the distance between adjacent tracer
output peaks-can be read directly and converted to circulation velocity from a
knowledge of the length of the circulation path. The liquid phase axial dispersion
coefficient ( E L ) for the circulation path may be obtained by fitting the equation

]@ER DETECTOR

TRACER
INLET

DETECTOR

FIGURE 16 Typical responses to pulse input of tracer in (a) airlifl loop reactors, and (b) batch
bubble columns.
230 M.Y. CHISTI A N D M. MOO-YOUNG
to the experimental tracer response c ~ r v e ; ~with . ' ~ Peclet
~ number (Pe) as the
fitting parameter. In practice Eq. (50) evaluated for the limits j = 0 f 2 may be
s u f f i ~ i e n t .In~ Eq. (50) C and CE are the instantaneous and final (equilibrium)
tracer concentrations, respectively, c is the dimensionless tracer concentration
(= C/CE), 6 is the dimensionless time and j is the dimensionless distance. 0 and j
are given by

and

where t and t, are the instantaneous time and the mean liquid circulation time,
respectively; and x, and x, are the total distance travelled by an element of fluid in
the direction of flow and the distance covered in one circulation, respectively.
The Peclet number in Eq. (50) is defined as

where UL is the mean liquid circulation velocity.


A combination of momentum balance with empirical gas holdup and two-phase
pressure drop correlations may be used to predict the liquid circulation velocities
in airlift reactors. At steady state the hydrostatic pressure difference driving force
for circulation resulting from gas holdup differences between the riser and the
downcomer is balanced by the total pressure drop in the circulation path due to
friction, valves, bends, flow area changes and internals. Similar approaches have
been used by several investigators.52~70.81~92~99~1061L1
Liquid circulation depends on
the superficial gas velocity and this dependence is ~ l a i m e d ~ ~ . ' ~ , to
" ~ be
, " ~of the
form:
UL= .u!8 (54)
in which cu is a function of the reactor geometry and of the properties of the
liquid, whereas /3 is determined by the flow regime as well as by reactor
geometry."' Gas sparger has little influence on liquid circulation rate39.114as long
as the entire cross-section of the riser is uniformly sparged; a contrary report1"
also exists. In two very different external loop airlift reactors' using water or
water-like fluids, independent found the exponent /3 in Eq. (54)
to be approximately 0.4. Using water or salt solution (0.15 k m ~ l m NaCI) -~ in
several external and concentric-tube internal loop reactors Bello et a1.'I3 observed
a value of 113 for this exponent; a depended"' on the reactor type and on
geometry as follows:

4 was 1.55 and 0.66 for the external and internal loop reactors, respectively, and
had the respective values of 0.74f 0.04 and 0.78f 0.08.113 Higher liquid
AIRLIFT BIOREACTORS 23 1
circulation in external loop reactors relative to internal loop systems occurred
because of the peculiar geometry of the former which allowed substantial gas
disengagement before the liquid entered the downcomer and this resulted in a
higher driving force for circulation in external loops. In almost the same external
loop vessels as used by Bello et d 1 1 3 another i n ~ e s t i g a t i o nconducted
~~ in
non-Newtonian CMC solutions (pap, of 0.015 to 0.5 Pas) showed decreasing
liquid circulation with fluid viscosity according to:

where cu was 0.052 and 0.0204 for bubble- and slug-flow regimes, respectively.
Notice that the exponents on the gas velocity and on the geometric term in Eq.
(56) were almost the same as reported earlier113for water and salt solution thus
pointing to their independence of fluid properties. Decline in liquid circulation
with increasing fluid viscosity has been observed by other researcher^^^^^^^^^^ also;
it arises probably because higher viscosities enhance internal friction losses. In
solutions containing small amounts of a drag reducing polymer such as xanthan
gum the liquid circulation was more rapid than in water.li5 Even in these fluids
increased viscosity at higher polymer concentration slowed down the circulation
of liquid.'15 For relatively large solids (d, =0.0027 m) dispersed in water in an
external loop device reducing liquid circulation with increase in solids loading has
been reported.97 We have observed similar results for mycelia-like solids in
external and internal loop reactors. For draught-tube gas sparged internal loops
weilandS7 recommended d J d , ratios between 0.8 and 0.9 for efficient mixing
(and oxygen transfer). Largest liquid circulation rates were reported at diameter
ratios of 0.5957 and approximately 0.5.1m Another report116 claimed minimal
mixing times for 0.6 < d,ld, < 1.0 (where dCi= d, corresponded to bubble column
operation) and for minimum mixing time equal riser and downcomer cross-
sectional areas (i.e. dcldci= d 2 ) were recommended116 in internal loops. Com-
paring bubble columns and concentric draught-tube internal loops with either one
or two draught-tubes, Margaritis and Sheppardz8 found that the presence of
draught-tubes lead to systematically higher mixing times relative to bubble
column operation. In addition, the double draught-tube configuration resulted in
higher mixing times than the single draught-tube geometry28 probably because of
the extended liquid circulation path in the former. Other similar observations
have been reported.l17 A problem with many existing correlations for liquid
circulation in airlift reactors is that they are not very successful for liquid velocity
prediction in anything other than the particular reactors used for obtaining them.
For low viscosity Newtonian fluids this difficulty has been overcome with a
recently proposed118 theoretical equation which related the superficial liquid
velocity in the riser (UL,) to various geometric and operating parameters as
follows:
232 M.Y. CHISTI AND M. MOO-YOUNG
KT and K B in Eq. (57) are the dimensionless frictional loss coefficients for fluid
turn around in the top and bottom riser-downcomer connecting zones, respec-
tively. Nearly all the published air-water liquid circulation data70~92~97~106~109"11~119
including that of the authors' could be correlated with Eq. (57) within f30%.
The data used included external and internal loop airlift reactors, the latter in
annulus sparged, draught-tube sparged and split cylinder geometries. Broad
ranges of reactor volume, reactor dispersion height, AJAd ratio and circulation
velocity were covered by Eq. (57) which demonstrated its usefulness as a scale-up
tool. The authors118 pointed out that Eq. (57) may be simplified for particular
reactor types and they also discussed the dependence of frictional loss coefficients
on reactor geometry.
Another empirical liquid circulation correlation specific to draught-tube inter-
nal loop reactors was reported79 recently for a limited amount of air-water data.
The mixing characteristics of the riser, downcomer and the headspace in airlift
vessels can be very different. Our tracer studies have shown that the reactor
headspace region is the best mixed whereas the downcomer has the poorest
mixing performance. Similar results have been found in a large external loop
In internal loops with concentric draught-tubes Weiland" observed that
the mixing time declined with increasing volume of liquid in the reactor
headspace above the draught-tube (Figure 17). Our observations partly confirmed
this trend. In a tall split cylinder airlift (d, = 0.243 m, A d / A ,= 0.411, L, = 4.8 m;
Lb = 0.102 m; air-water) we found improved mixing performance with increasing
height (h,) of liquid above the baffle up to a height of about 0.5 m. Further
increase in hH lead to different results: while the mixing in the zone above the

0 0
0 0.10 0.20 0.30 0.40 0.50
DISTANCE h, ( m )

FIGURE 17 Effect of the clear liquid height in the reactor headspace above the draught-tube on
mixing time (95% mixing) in air-water in an internal loop reactor (draught-tube sparged at
U,, = 0.0105 ms-', dJd, = 0.76). Data from ~ e i l a n d . ~ '
AIRLIFT BIOREACTORS 233
baffle (Figure 2a) remained good, the intermixing of the fluid in this zone with
that in the baffled height of the split cylinder slowed down markedly. There was
no contradiction between our results and those of weilands7 because the results
of the latter (Figure 17) did show a definite flattening of mixing time versus h,
curve near hH of 0.4 to 0.5m. Based on our results we recommend that the clear
liquid height above the top of the baffle (or draught-tube) should not exceed
about 0.5m. This value is independent of the total height of the airlift reactor.
Significant reductions in mixing time may also be obtained by horizontally
splitting (Figure 2c) the central draught-tube of internal loops into two or more
vertical sections.99 We are currently experimenting with a split cylinder airlift with
multiple baffles.
Because the liquid circulation is very sensitive to the gas holdup difference
between the riser and the downcomer of an airlift vessel, incorporation of various
types of gas-liquid separators in the headspace would enhance circulation and
mixing depending on the separator efficiency. This in fact is the reason for the
different circulation performances of internal and external loop airlifts. In the
former there usually is no gas-liquid separator whereas in the latter the top
connection between the riser and downcomer acts as a gas-liquid disengagement
unit. Gas-liquid separator design considerations have recently been d i s c u s ~ e d . ~ ~
For best circulation performance separators should be designed for minimum flow
resistance.
Liquid circulation effects in some actual fermentation systems have been
discussed by Schiigerl et al. lZO

5.3 Heat Transfer


Fluid-wall heat transfer is one of the least studied phenomenon in airlift reactors
probably because heat transfer limitations are only rarely encountered for a
greater proportion of industrial fermentations. Fermentations which utilize highly
reduced carbon sources such as hydrocarbons and methanol generate much larger
quantities of heat. For most productive fermentations typical metabolic heat
generation1 is of the order of 3 to 5 k W n ~ - ~Rate
. of heat evolution is dependent
on the rate of oxygen consumption. Bailey and O l ~ i s 'quote~ ~ a study which
relates heat evolution during separate growth of E. coli, Candida intermedia, B.
subtilis, and A. niger on carbohydrates in the form

where Q (kWm-3) and Ro, (kgm-3s-') are the rates of heat evolution and
oxygen consumption, respectively. Once the heat transfer rate, which equals the
heat evolution rate plus the heat generation due to agitation at steady state, is
established then the reactor wall area needed to obtain this rate is calculated
from:

where A, is the heat transfer area and AT is a mean temperature difference


driving force. U H , the overall heat transfer coefficient, is the sum of the
234 M.Y. CHISTI AND M. MOO-YOUNG
resistances to heat transfer due to the fluid films on either side of the heating or
cooling surface, fouling resistances on either side, and the resistance due to the
metal wall through which heat must pass. Hence,

where h's are the individual heat transfer coefficients and the subscripts i, f, m
and o refer to heating/cooling fluid film, fouling, metal wall, and the film between
the fermentation broth and heating/cooling surface, respectively. Empirical
correlations and methods of calculation for hi, hv and h, abound in chemical
engineering texts and works on heat transfer; her, the fouling coefficient between
the broth and reactor wall, is essentially constant and typical data given in
handbooks should apply. The calculation of h,, the film coefficient between the
reactor fluid and the heating/cooling surface, is the object of research in airlift
and other bioreactors. For airlift vessels with very low rates of fluid circulation
the heat transfer data obtained in batch bubble columns may apply. This work
has been reviewed by other^.^^.^^.'^^ For air-water system in industrial batch
bubble columns (diameters> 0.45 m) Fair el found the film coefficient h, to
increase with superficial gas velocity in the following way:

indicating that the film coefficient was independent of column diameter. A


correlation suitable for a greater range of fluids in bubble columns was provided
by ~ e c k w e and
r ~ ~it took the form:

which also predicted an increase in the film coefficient with gas velocity raised to a
power of 0.25. In Eq. (62) k and C,, are, respectively, the thermal conductivity
and the specific heat capacity of the liquid in the reactor. Compared to bubble
columns, heat transfer coefficients in airlift reactors are more than 2-fold
higher,li2 due to higher liquid velocities in the latter.
Only two correlations for film heat transfer coefficient in concentric-tube
internal loop airlift reactors are available. That of Chakravarty et a1.Iz3 which is:

in which U, is based on the riser diameter. Equation (63) was obtained using
Newtonian fluids ( p L = 0.78 - 5.27 mPas) in draught-tube sparged airlifts over
approximate U, and A,/A, ranges, respectively of 0.008 to 0.16 ms-', and 0.25 to
1.20. The actual values of h, variedIz3 from 0.6 to 2.4 kWm-zOC-l. The second
equation was proposedIz4 by us for air-water in concentric draught-tube internal
AIRLIFT BIOREACTORS
loops:

, .
Equation (64) covered a gas velocity (based on total reactor cross-section) range
of 0.01 to 0.04 ms-'. whereas A -J A ., values of 0.242 and 0.452 were used in its
determination. Equation (64) predicts significantly higher heat transfer
coefficients than does correlation (63). . . The differences could be due to different
types of gas sparging modes used: annulus sparged as used by us us. draught-tube
sparged as used by Chakravarty et ~ 1 . Additionally,
' ~ ~ the geometric configura-
tions of the bottoms of the reactors used by the two groups of investigators were
quite different and may have lead to different liquid velocities for otherwise
similar conditions. S t u d i e ~ ' ~of~ .heat
' ~ ~transfer in airlift vessels in suspensions of
Aspergillus niger and PeniciNium chrysogenum claimed that the presence of
mycelia enhanced film heat transfer coefficients. The data presented'z showed h,
values as high as 8 kWm-'K-'. A more detailed studyJZ4conducted with
cellulose fibre suspensions which simulated mycelial media showed that the
dependence of heat transfer on solid contents can be very complex. Solids may
enhance or reduce heat transfer depending on the location (riser or downcomer)
of heating/cooling surface in the reactor. Useful recommendations on the
practical problem of positioning of heat exchange devices in airlift vessels were
presented by Chisti and coworker^.'^^ Nishikawa et ~ 1claimed . ~ film~ heat transfer
coefficient to be independent of liquid velocity for UL5 0.015 ms-', however for
higher liquid velocities h, depended on the velocity of liquid as follows:

Heat transfer work done in vertical two-phase flows may be applicable to airlift
reactors provided that the fluid properties, gas holdup and relative velocities of
the two phases are identical for the airlift and vertical two-phase flow device.
In many fermentations, particularly those which have high solids content or
which produce very viscous broths, the presence of heat transfer coils and similar
projections in the reactor is undesirable because of the formation of stagnant
zones, problems with cleaning and sterilization. Consequently, jacket form of
heating/cooling is preferred. As the reactor volume increases, however, the
surface-to-volume ratio declines and a point is soon reached beyond which a
jacket alone is insufficient to handle the heat duty. In airlift reactors this problem
is largely overcome by using double walled draught-tubes which provide
additional heat transfer surface, and by other similar modifications.

6. CONCLUDING REMARKS

While the practical feasibility of airlift reactors for a variety of fermentations has
already been established, systematic studies of these devices have only just
begun. Even with very limited knowledge a coherent picture of some aspects of
236 M.Y. CHISTI AND M. MOO-YOUNG
the airlift reactors can be made as has been done ,in the foregoing sections of this
review. Specific design recommendations appear throughout the review. For low
viscosity systems it is now possible to predict liquid circulation velocity for known
gas holdup using well known principles of fluid mechanics. This ability is lacking
for viscous Newtonian and rheologically complex media. All existing heat and
mass transfer and gas holdup correlations for airlift reactors are essentially
empirical although possible underlaying hydrodynamic basis for them has been
discussed. It would be most useful to show theoretically the effects of various
geometric and operational variables on reactor performance parameters. Mechan-
ically simple reactor designs-split cylinder airlift for e x a m p l e n e e d to be
studied. Gas-liquid separator designs as well as multipoint riser/downcomer gas
injection should be examined.
The problem of quantification of local and global shear rates in airlift vessels is
of critical importance not only because of the shear sensitivity of biological
materials, but also because of the shear dependant flow behaviour of non-
Newtonian systems. This problem remains to be addressed.

ACKNOWLEDGEMENT

This work was supported by a grant from the Natural Sciences and Engineering
Research Council of Canada.

NOMENCLATURE

(M =mass, L =length, T = time, 6 =temperature, m = mole)


A Total cross-sectional area of reactor L2
Ad Cross-sectional area of downcomer LZ
Heat transfer surface area L2
Cross-sectional area of riser L2
Gas-liquid interfacial area per unit dispersion volume L-'
Gas-liquid interfacial area per unit liquid volume L-'
Instantaneous tracer concentration MLW3
Dimensionless tracer concentration (= C/CE)(-)
Equilibrium tracer concentration MLT3
Concentration of the transferring component in liquid ML-'
Specific heat capacity of liquid L'T-~O-'
Dry weight solids per volume of suspension %
Saturation concentration of transferring component in liquid ML-3
Diffusivity of transferring component in liquid L ~ T - '
Sauter mean bubble diameter L
AIRLIFT BIOREACTORS
Diameter of the ith bubble L
Reactor diameter, outer tube diameter L
Draught-tube diameter L
Sparger hole diameter L
Solid particle diameter L
Riser diameter L
Liquid phase axial dispersion coefficient LZT-'
Froude number (= U&lgd,)(-)
Fraction of free sparger area relative to riser area (-)
Gravitational acceleration L T - Z
Dispersion height L
Distance between top of draught-tube and clear liquid surface L
Inside film coefficient for heat transfer MT-3%-'
Inside fouling coefficient for heat transfer M T - ~ % - '
Metal wall heat transfer coefficient MT-3%-'
Outside film coefficient for heat transfer M T - ~ % - '
Outside fouling coefficient for heat transfer MT-3%-1
Dimensionless distance (= x,/x,) (-)
Consistency index of power law fluid ML-'T"-'
Bottom frictional loss coefficient (-)
Top frictional loss coefficient (-)
Thermal conductivity of liquid MLT-'%-I
True mass transfer coefficient LT-I
Unaerated liquid height L
Draught-tube clearance from reactor base L
length of draught-tube L
Molar mass of gas Mm-'
Morton number ( = g p ~ / p L u 3(-)
)
Flow behaviour index of power law fluid (-)
Frequency of occurrence of bubble of diameter dBi (-)
Peclet number (= ULx,IEL) (-)
Pressure at the bottom of reactor M L - ' T - ~
Power input due to gas ML2T-'
Headspace pressure ML-lT-'
Rate of heat transfer M L - 1 T - 3
Molar gas flow rate mT-'
Gas constant M L Z ~ - 2 % - ' m - 1
M.Y. CHISTI AND M. MOO-YOUNG
Oxygen consumption rate M L - 3 T - 1
Substrate consumption rate M L - ~ T - '
Substrate concentration ML-3
Maximum tolerable substrate concentration ML-3
Minimum acceptable substrate concentration M L - ~
Absolute temperature 0
Temperature difference 6
Time T
Circulation time T
Bubble rise velocity LT-I
Overall heat transfer coefficient M T - 3 6 - 1
Linear liquid velocity LT-'
Superficial liquid velocity in downcomer LT-'
Superficial liquid velocity in riser LT-'
Gas velocity in the sparger orifice LT-'
Superficial gas velocity LT-'
Mean superficial gas velocity LT-I
Mean superficial riser gas velocity LT-I
Volume of dispersion in reactor L3
Volume of liquid in reactor L~
Volume of liquid in riser L~
Axial distance L
Distance travelled per circulation L
Total distance travelled in time t L

Greek Letters
LY A constant
B A constant
Y Shear rate T-'
EG Overall fractional gas holdup (-)
E G ~ Downcomer fractional gas holdup (-)
EG~ Riser fractional gas holdup (-)
rl Efficiency (-)
0 Dimensionless time (= tlt,) (-)
pmPp Apparent viscosity of nowNewtonian liquid ML-'T-'
Gas viscosity ML-'T-'
PL Liquid viscosity ML-IT-'
AIRLIFT BIOREACTORS
PW Viscosity of water ML-IT-'
PC Density of gas ML-3
PL Density of liquid ML-3
u Surface tension MT-2
4) A constant
111 A constant

Abbreviations
CMC Carboxymethyl cellulose
SCP Single cell protein
SF Solka-Floc
VVM Volume of gas per volume of liquid per minute

REFERENCES

1. Kossen, N.W.F., Paper presented at the 3rd EFB Conference, Munich, 10-14 September, 1984.
2. Bailey, J.E., Chem. Eng. Sci., 35, 1854 (1980).
3. Maclean, G.T., Erickson, L.E., Hsu, K.H., and Fan, L.T., Biotech. Bioeng., 19, 493 (1977).
4. Viesturs, U.E., Sturmanis, I.A., Krikis, V.V., Prokopenko, D., and Erickson, L.E., Biotech.
Bioeng., 22, 799 (1980).
5. Viesturs, U.E., Sturmanis, LA., Krikis, V.V., Prokopenko, V.D., and Erickson, L.E., Biotech.
Bioeng., 23, 1171 (1981).
6. Baird. M.H.I.. and Garstane. J.H.. Chem. Enn. Sci.. 27. 823 (1972) \ -1

7. ~ e i t e l G.,
, and Onken, ~ . , % e r .Chem. ~ n ~ . , - 250
4 , (1981).
8. Russell, T.W.F., Dunn, I.J., and Blanch, H.W., Biotech. Bioeng., 16, 1261 (1974).
9. Ziegler, H., Meister, D., Dunn, I.J., Blanch, H.W., and Russell, T.W.F., Biotech. Bioeng., 19,
507 (1977).
10. Stein, W.A., and Schafer, H., Ger. Chem. Eng., 6, 91 (1983).
11. Stein, W.A., and Schafer, H., Ger. Chem. Eng., 7 , 115 (1984).
12. Zanker, A,, Chemical Engineering, 21 November (1977), pp. 207.
13. Moresi, M., Biotech. Bioeng., 23, 2537 (1981).
14. Blakebrough, N., Shepherd, P.G., and Nimmons, Biotech. Bioeng., 9, 77 (1967).
15. Westlake, R., Chem. Ing. Tech., 58, 934 (1986).
16. Ho, C.S., Erickson, L.E., and Fan, L.T., Biotech. Bioeng., 19, 1503 (1977).
17. Malfait, J.L., Wilcox, D.J., Mercer, D.G., and Barker, L.D., Biotech. Bioeng., 23, 863 (1981).
18. Erickson, L.E., Patel, S.A., Glasgow, L.A., and Lee, C.H., Process Biochem., 18(3), 16 (1983).
19. Koenig, B., Seewald, C., and Schiigerl, K., in Advances in Biotechnology, Vol. 1 (Moo-Young,
M., Robinson, C.W., and Vezina, C., eds.), Pergamon Press (Toronto), (1981). p. 573.
20. Seipenbusch, R., Birckenstaedt, J.W., Blenke, H., and Schindler, F., Abstracu: Fifrh
International Fermentation Symposium, Berlin (1976), p. 65.
21. Hatch, R.T., in Single Cell Protein 11, Tannenbaum, S.R., and Wang, D.I.C., (Editors), The
MIT Press (Cambridge), (1975), pp. 46.
22. Huang, S.Y., Yeh, M.C., and Liou, K.T., Abstracu: Fifrh lnrernational Fermentation
Symposium, Berlin (1976). pp. 68.
23. Smart, N.J., and Fowler, M.W., Journal of Experimental Botany, 35 (153), 531 (1984).
24. Gallo, T., and Sandford, D.S., Paper . presented
. - o.f A I C h E . A ~ r i l .1-5.
at 86rh National Meen'nn
1979, Houston, Texas.
25.
- ~
Hines. D.A.. Bailev. M.. Ousbv. J.C.. and Roesler., F.C.. ~ Paoer
, oresented
. , at the IChemE-
~

conference, ~ o r k land), 16-17 (1975).


26. CIL Publication (1978): Deep Shaft Effluent Treatment Process, CIL Canada, Inc.
27. Wood, L.A., and Thompson, P.W., Paper presented at the International Conference on
Bioreactor Fluid Dynamics, Cambridge (England), 15-17 April (1986), Paper 11, p. 157.
240 M.Y. CHISTI AND M. MOO-YOUNG
28. Margaritis, A., and Sheppard, J.D., Biorech. Bioeng., 23, 2117 (1981).
29. Siegel, M.H., Merchuk, J.C., and Shiigerl, K., Paper presented at the International Conference
on Bioreacror Fluid Dynamics, Cambridge (England). 15-17 April, 1986.
30. Gasner, L.L., Biorech. Bioeng., 16, 1179 (1974).
31. Piggott, A.M., MASc Thesis, University of Waterloo, Ontario (1985).
32. Chisti, M.Y., and Moo-Young, M., Biorech. Bioeng., Accepted (1987).
33. Chisti, M.Y., Fujimoto, K., and Moo-Young, M., Paper 117a presented at A I C h E Annual
Meerinn. Miami Beach, November 2-7, 1986.
34. ~ e c k w i r ,W.-D., in "Biorechnology", vol. 2, (Rehm, H.-J., and Reed, G., editors), VCH,
(Weinheim), (1985). Chapter 20, pp. 445.
35. Lin, C.H., Fang, B.S., Wu, C.S., Fang, H.Y ., Kuo, T.F., and Hu, C.Y., Biorech. Bioeng., 18,
1557 (1976).
36. ~ h a h , ' ~ . ~Kelkar,
:, B.G., Godbole, S.P., and Deckwer, W.-D., A I C h E J., 28, 353 (1982).
37. Collier, J.G., "Convecriue Boiling and Condensation,'' McGraw-Hill, New York (1972); Hewitt,
G.F.. and Roberts, D.N.. "Srudies parterns by simulraneolrs X-ray andflash
phorography", AERE-M 2159, HMSO (1969).
38. Weiland, P., and Onken, U., Ger. Chem. Eng., 4, 174 (1981).
39. Merchuk, J.C., Chem. Eng. Sci., 41, 11 (1986).
40. Moo-Young, M., and Blanch, H.W., Adu. Biochem. Eng., 19, 1 (1981).
41. Guy, C., Carreau, P.J., and Paris, J., Canad, J. Chem. Eng., 64, 521 (1986).
42. Robinson, C.W., Canad. I . Chem. Eng., 64, 523 (1986).
43. Zlokarnik, M., in "Biorechnology", vol. 2 (Rehm, H.-J., and Reed, G., eds), VCH-Verlag
Chemie, Weinheim (1985), Chapter 23, pp. 537.
44. McManamey, W.J., Wase, J.D.A., Raymahasay, S., and Thayanithy, K., J. Chem. Tech.
Biorechnol., 348, 151 (1984).
45. Chisti, M.Y., and Moo-Young, M., Chem. Eng. J., Accepted (1987).
46. Hills, J.H., Chem. Eng. I . , 12, 89 (1976).
47. Bhavaraju, S.M., Russell, T.W.F., and Blanch, H.W., A I C h E J., 24, 454 (1978).
48. Azbel, D., "Two-phase Flows in Chemical Engineering", Cambridge University Press (Cam-
bridge), (1981).
49. Dussap, G., and Gros, J.B., Chem. Eng. J . , 25, 151 (1982).
50. Godbole, S.P., Schumpe, A., Shah, Y.T., and Carr, N.L., A I C h E J . , 30, 213 (1984).
51. Glasgow, L.A., Erickson, L.E., Lee, C.H., and Patel, S.A., Chem. Eng. Commun., 29, 311
(1984).
52. Kubota, H., Hosono, Y., and Fujie, K., J. Chem. Eng. Jpn., 11(4), 319 (1978).
53. van der Lans, R.G.J.M., and Smith, J.M., Fourth European Conference on Miring, April 27-29
(1982). Leeuwenhont (The Netherlands).
54. Maslen, F.P., Ousby, J.C., and Senior, P.J., US Parenr 4, 306, 026, December 15, 1981.
55. Fields, P.R., and Slater, N.K.H., Biorech. Bioeng., 26, 719 (1984).
56. Merchuk, J.C., and Stein, Y., Biorech. Bioeng., 23, 1309 (1981).
57. Weiland, P., Ger. Chem. Eng., 7, 374 (1984).
58. Kawase, Y., and Moo-Young, M., Chem. Eng. Commun., 40, 67 (1986).
59. Chudacek, M.W., Chemical Engineering, 91 (20). 79 (1984).
60. Joshi, J.B., and Sharma, M.M., Trans. Inst. Chem. Eng., 57, 244 (1979).
61. Schiigerl, K., Adv. Biochem. Eng., 19, 71 (1981).
62. Yoshida, F., Annual Repors on Fermentation Processes, vol. 5 , (Tsao, G.T., editor), Academic
Press (London), (1982), Chapter 1, pp. 1-34.
63. Sharma, M.M., in "Recent Aduances in the Engineering Analysis of Chemically Reacting
Systems", (Doraiswamy, L.K., editor), Wiley Eastern Limited (New Delhi), (1984), pp. 291.
64. Linek, V., and Vacek, V., Chem. Eng. Sci., 36, 1747 (1981).
65. Noltenkamper, R., Steiff, A., and Weinspach, P.-M., Ger. Chem. Eng., 6, 147 (1983).
66. Cheremisinoff, N.P., Ind. Eng. Chem. Process Des. Develop., 25, 329 (1986).
67. Shah, Y.T., Stiegel, G.J., and Sharma, M.M., A I C h E I . , 24, 369 (1978).
68. Kargi, F., and Moo-Young, M., in "Comprehensive Biorechnology", vol. 2 (Moo-Young, M.,
editor). Pergamon Press (Oxford), (1985), p. 5.
69. Akita, K., and Yoshida, F., Ind. Eng. Chem. Process Des. Develop., 13, 84 (1974).
70. Bello, R.A., Ph.D. Thesis, University of Waterloo, Ontario (1981).
71. Bello, R.A., Robinson, C.W., Moo-Young, M., Chem. Eng. Sci., 40, 53 (1985).
72. Bello, R.A., Robinson, C.W., and Moo-Young, M., Biorech. Bioeng., 27, 369 (1985).
73. Popovic, M., and Robinson, C.W., Proceedings of rhe 34th Canadian Chemical Engineering
Congress. 30th September-3rd October (1984). Quebec City, p. 258.
AIRLIFT BIOREACTORS
74. Stejskal, J., and Potucek, F., Biotech. Bioeng., 27, 503 (1985).
75. Koide, K., Sato, H., and Iwamoto, S., J. Chem. Eng. Jpn., 16(5), 407 (1983).

413 (1983j.
-
76. Koide. K.. Kurematsu.. K... Iwamoto.. S... Iwata.. Y... and Horibe.. K... J. Chem. Enn. Jon..
' 16f5).
77. Andrt, G., Ph.D. Thesis, University of Waterloo, Ontario (1982).
78. El-Gabbani, D.H., MASc Thesis, University of Waterloo, Ontario (1977).
79. Kawase. Y.. and Moo-Youne. M.. J. Chem. Tech. Biotechnol.. 36. 527 (1986).
80. ~hakravarty,M., Begum, S, Singh, H.D., Baruah, J.N., and l y e n g a r , ' ~ . s ,Biotech. Bioeng.
Symp. No. 4, 363 (1973).
81. Miyahara, T., Hamaguchi, M., Sukeda, Y., and Takahashi, T., Canad. J. Chem. Eng., 64, 718
(1986).
82. ~ishikawa,M., Kato, H., and Hashimoto, K., Ind. Eng. Chem. Process Des. Develop. 16, 133
\11977).
--..,-
83. McManamey, W.J., and John Wase, D.A., Biotech. Bioeng., 28, 1446 (1986).
84. Deckwer, W.-D., Nguyen-Tien, K., Schumpe, A,, and Serpemen, Y., Biotech. Bioeng., 24, 461
(1987)
\----I'
85. Henzler, H.J., Chem.-1ng.-Tech., 53, 634 (1980).
86. Nakanoh, M., and Yoshida, F., Ind. Eng. Chem. Process Des. Develop., 19, 190 (1980).
87. Akita, K., and Yoshida, F., Ind. Eng. Chem. Process Des. Develop., U,76 (1973).
88. Kataoka, H., Takeuchi, H., Nakao, K., Yagi, H., Tadaki, T., Otake, T., Miyauchi, T.,
Washimi, K., Watanabe, K., Yoshida, F., I . Chem. Engr. Jpn., 12(2), 105 (1979).
89. Fair, J.R., Lambright, A.J., and Andersen, J.W.. Ind. Eng. Chem. Process Des. Develop., 1, 33
(1967)
\-'--,'
90. Lippert, I., Adler, I., Meyer, H.D.. Lubbert. A., and Schiigerl. K.. Biotech. Bioeng., U,437
(1983).
91. Menzel, T., Kantorek, H.J., Franz, K., Buchholz, R., and Onken, U., Chem.-lng. Tech., 57(2),
S139 (1985).
92. Merchuk, J.C., and Stein, Y., A I C h E I . , 27, 377 (1981).
93. Moo-Young, M., and Kawase, Y., Canad. I . Chem. Eng., 65, 113 (1987).
94. Bello, R.A., Robinson, C.W., and Moo-Young, M., in Advances in Biotechnology, Vol. 1,
(Moo-Young, M., Robinson, C.W., and Vezina, C., eds.). Pergamon Press (Toronto), (1981).
p. 547.
95. Fan, L.-S., Fujie, K., and Long, T.-R., AlChE Symp. Ser. 80 (241), 102 (1984).
96. Fan, L.4. Hwang, S.-J., and Matsuura, A,, Chem. Eng. Sci. 39, 1677 (1984).
97. Verlaan, P., Tramper, J., van't Riet, K., and Luyben, K.Ch.A.M., Paper presented at the
International Conference on Bioreactor Fluid Dvnamics. Cambridee (Eneland). 15-17 Aoril.
- ~ u

1986, Paper 07.


98. Herskowitz, M., and Merchuk, J.C., Canad. J. Chem. Eng., 64, 57 (1986).
99. Blenke. H.. Adv. Biochem. Ene.. 13. 121 (1979).
100. braze& M'.E., Fan, L.T., and%ckon, L.E., ~iotech.Bioeng., 21, 1579 (1979).
101. Erickson, L.E., and Deshpande, V., in Advances in Biotechnology, Vol. 1, (Moo-Young, M.,
Robinson, C.W., and Vezina, C., eds.), Pergamon Press (Toronto), 1981, p. 553.
102. Stern, A.M., and Gasner, L.L., US Patent no. 3, 764,014, October 9, 1973.
103. Andre, G., Robinson, C.W., and Moo-Young, M., Chem. Eng. Sci., 38, 1845 (1983).
104. Onken, U., and Weiland, P., in Advances in Biotechnology, Vol. 1 (Moo-Young. M., Robinson,
C.W., and Vezina, C., eds)., Pergamon Press (Toronto), 1981, p. 559.
105. Voncken, R.M., Holms, D.B., and den Hartog, H.W., Chem. Eng. Sci., 19, 209 (1964).
106. Chakravarty, M., Singh, H.D., Baruah, J.N., and lyengar, MS.. Indian Chemical Engineers,
16(3), 17 (1974).
107. Hsu, Y.C., and Dudukovic, M.P., Chem. Eng. Sci., 35, 135 (1980).
108. Weiland, P., and Onken, U., Ger. Chem. Eng. 4, 42 (1981).
109. Jones, A.G., Chem. Eng. Sci., 40, 449 (1985).
110. Lee, C.H., Glasgow, L.A., Erickson, L.E., and Patel, S.A., Paper 8d presented at the AlChE
Annual Meeting, Miami Beach. November 2-7 (1986).
111. Verlaan, P., Tramper, J., Van? Riet, K., and Luyben, K.Ch.A.M., Chem. Eng. J., 33, 843
(1986).
112. Onken, U., and Weiland, P., in "Advances in Biotechnological Processes", 1 (Mizrahi, A. and
Van Werzel, L. eds), Alan R. Liss, Inc. (New York), (1983) p. 67-95.
113. Bello, R.A., Robinson, C.W., and Moo-Young, M., Canad. J. Chem. Eng. 62, 573 (1984).
114. Onken, U., and Weiland, P., Eur. J. Appl. Microbial Biotechnol. 10, 31 (1980).
115. Fields, P.R., Mitchell, F.R.G., and Slater, N.K.H., Chem. Eng. Commun., 25, 93 (1984).
M.Y. CHISTI AND M. MOO-YOUNG
116. Rousseau, I., and Bu'lock, J.D., Biotechnol. Lett., 2, 475 (1980).
117. Pandit, A.B., and Joshi, J.B., Chem. Eng. Sci., 38, 1189 (1983).
118. Chisti, M.Y., Halard, B., and Moo-Young, M., Submitted for publication (1987).
119. Hatch, R.T., Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge (1973).
120. Schugerl, K., Burschiipers, J . , Czech, K., Frieling, M.v., Frohlich, S., Gebauer, A., Lorenz, T . ,
Lubbert. A., Ross, A., and Scheper, T . , Paper presented at the International Conference on
Bioreoctor Fluid Dynamics, Cambridge (England), 15-17 April (1986).
121. Bailey, J.E., and Ollis, D.F., "Biochemical Engineering Fundamenmls", McGraw-Hill (New
York), (1977). p. 482.
122. Deckwer, W.-D., Chem. Eng. Sci., 35, 1341 (1980).
123. Chakravarty, M.. Singh, H.D., Baruah, J.N., Iyengar, M.S., and Sinha, A.P., Paper presented
at the First National Heat and Mass Transfer Conference, Madras (India), (1971). Paper no.
HMT-53-71.
124. Ouyoung, P.K., Chisti, M.Y., and Moo-Young, M., Submitted for publication (1987).
125. Blakebrough, N., McManamay, W.J., and Walker, G., Chem. Eng. Res. Des., 61, 264 (1983).
126. Blakebrough, N., Fatile, I.A., McManamey, W.J., and Walker, G . , Chem. Eng. Res. Des., 61,
383 (1983).

You might also like