Sei sulla pagina 1di 25

Biochemical Engineering Journal 39 (2008) 405429

Review
Oxygen transfer in hydrocarbonaqueous dispersions and its
applicability to alkane bioprocesses: A review
K.G. Clarke

, L.D.C. Correia
DST-NRF Centre of Excellence in Catalysis (c* change), Department of Process Engineering, University of Stellenbosch, South Africa
Received 13 June 2007; received in revised form 31 October 2007; accepted 25 November 2007
Abstract
Accumulation of alkane by-products fromgas to liquid fuel processes presents an attractive feed stock opportunity with potential for bioconversion
to a wide variety of valuable commodity products. This review highlights the need to address the complexities of the oxygen transfer rate and
overall volumetric oxygen transfer coefcient (K
L
a) in hydrocarbon bioprocesses so that this potential can be realised.
Three markedly different K
L
a behavioural trends have been identied in hydrocarbonaqueous dispersions and characterised according to the
hydrocarbon type and concentration, operating conditions and geometric constraints. A fundamental conceptual understanding of the mechanisms
which dene the exact behaviour of K
L
a in response to changes in turbulence and uid properties is provided. Further, the behaviour is quantied
in terms of the parameters which underpin this response viz. bubble diameter, gasliquid interface rigidity, gas hold up, surface tension, viscosity
and diffusivity.
Consideration is given to existing predictive correlations for interfacial transfer area, bubble diameter, gas hold up, oxygen transfer coefcient
and K
L
a. It is envisaged that through the elucidation and quantication of the parameters which shape the behaviour of K
L
a, these correlations may
be successfully extended to predict the complex behavioural K
L
a trends in hydrocarbon-based bioprocesses.
2007 Elsevier B.V. All rights reserved.
Keywords: Alkane bioconversion; Aerobic processes; Gasliquid transfer; Oxygen transfer; Agitation; Fluid properties
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
1.1. Alkanes as an attractive feedstock opportunity for bioconversion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
1.2. Adequacy of oxygen supply to alkane-based bioprocesses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
2. Oxygen transfer rate and overall volumetric oxygen transfer coefcient in hydrocarbonaqueous dispersions . . . . . . . . . . . . . . . . . . . . . 408
2.1. Oxygen solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408
2.2. Trends in overall volumetric oxygen transfer coefcient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
3. Factors inuencing overall volumetric oxygen transfer coefcient in hydrocarbonaqueous dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . 412
3.1. Inuence of turbulence on interfacial area and oxygen transfer coefcient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412
3.1.1. Inuence of turbulence on interfacial area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412
3.1.2. Inuence of turbulence on oxygen transfer coefcient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
3.2. Inuence of uid properties on interfacial area and oxygen transfer coefcient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
3.2.1. Inuence of uid properties on interfacial area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
3.2.2. Inuence of uid properties on oxygen transfer coefcient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
4. Prediction of the overall volumetric oxygen transfer coefcient and oxygen transfer rate in hydrocarbonaqueous dispersions . . . . . . 421
5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425

Corresponding author at: Department of Process Engineering, University of Stellenbosch, Private Bag X1, Stellenbosch 7602, South Africa.
Tel.: +27 21 8084421; fax: +27 21 8082059.
E-mail address: kclarke@sun.ac.za (K.G. Clarke).
1369-703X/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.bej.2007.11.020
406 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
Nomenclature
a gas/liquid interfacial area per unit volume
(m
2
/m
3
)
A function of ow index {280(n +1)/
39(3/2)
2
}
1/(n+1)
A
s
area of sieve plate (m
2
)
C constant
C
a
additive concentration (mol/L)
C
G
concentration of the solute in the gas phase
(mol/m
3
)
C
N
carbon number
C
o
concentration of dissolved oxygen in bulk liquid
(mol/m
3
)
C
o,sat
saturation concentration of dissolved oxygen
(mol/m
3
)
Crk
2
/ parameter of bubble coalescence proposed by
Marrucci [145]
C
s
average solid concentration gas-free slurry
(
S

S
/(
L
+
S
)) (kg/m
3
)
C
x
biomass concentration (g/L)
d
b
gas bubble diameter (m)
D
32
Sauter mean diameter (m)
D
d
inner diameter of draught tube (m)
D
G
gas phase diffusivity (m
2
/s)
D
i
impeller diameter (m)
D
L
liquid phase diffusivity (m
2
/s)
D
S
orice diameter (m)
D
t
reactor/column diameter (m)
E
t
total power input (W)
g gravitational constant (m
2
/s)
H Henrys law constant (Pa m
3
/mol)
H
t
dispersion height (m)
H
WO
partition coefcient of oxygen in wateroil sys-
tem
L length parameter (m)
k consistency index in a power law model (Pa s)
k
e
local mass transfer coefcient (m/s)
k
1
(d/dC
a
)(3V
sl
d
b
/4V
B
)/(d
b
RT/2)
K
L
liquid phase oxygen transfer coefcient
(m/s)
(K
L
)
GO
K
L
through the hydrocarbon lm surrounding the
gas bubbles (m/s)
(K
L
)
GW
K
L
through the aqueous lm surrounding the gas
bubbles (m/s)
(K
L
)
OW
K
L
through the aqueous lm surrounding the
hydrocarbon droplets (m/s)
K
L
a overall volumetric oxygen transfer rate (s
1
)
K
L
a
GW
K
L
a through the aqueous lmsurrounding the gas
bubbles (s
1
)
K
L
a
OW
K
L
a through the aqueous lm surrounding the
hydrocarbon droplets (s
1
)
K
L
a
max
maximumoverall volumetric oxygen transfer rate
(s
1
)
M {(
GT,o
/
GH,o
) 1}at U
GC
M
v
molecular volume (m
3
/mol)
n ow index in a power law model
N impeller speed (rpm)
N
G
number of holes in gas distributor
N
s
number of holes on sieve plate
N
VD
Van Dierendonck speed (rpm)
OTR oxygen transfer rate (mg/Ls)
p
o
partial pressure of oxygen (Pa)
P power dissipated by impeller (W)
P/V power per unit volume of liquid phase (W/m
3
)
P
e
/V effective power per unit volume of liquid phase
(W/m
3
)
q height of the agitator blades (m)
Q volume air per volume liquid per minute (vvm)
Q
G
gassed volumetric ow rate (m
3
/s)
R universal gas constant (Pa m
3
/mol K)
t time (s)
T absolute temperature (

C)
T
+
dimensionless time period
U
G
gas velocity based on cross section of column and
based on average static pressure in column (m/s)
U
GC
gas velocity where maximum value of
{(
GT,o
/
GH,o
) 1} is observed (m/s)
V volume of liquid (m
3
)
V
B
bubble rise velocity (m/s)
V
D
drift ux (m/s)
V
gel
volume of gels (m
3
)
V
L
supercial liquid velocity (m/s)
V
s
supercial gas velocity (m/s)

V
s
mean supercial gas velocity at H
t
/2 (m/s)
V
sl
slip velocity (m/s)
V
t
terminal gas bubble velocity in free rise (m/s)
X
n
aqueous phase xanthan concentration (kg/m
3
)
Greek letters
, & exponent constants

r
apparent yield stress to shear stress ratio

N
nozzle diameter (m)
power rate per unit mass (W/kg)

G
gas holdup; fraction of total volume of aerated
liquid

GH,o
gas holdup for heterogeneous ow regime

GT,o
gas holdup for transition regime

L
liquid holdup

S
solid holdup

S/L
V
gel
/(V
gel
+V) =partical concentration
characteristic material time (s)

a
apparent viscosity (Pa s)

c
continuous phase viscosity (Pa s)

d
dispersed phase viscosity (Pa s)

G
gas viscosity (Pa s)

L
liquid viscosity (Pa s)

L
kinematic viscosity of liquid phase (m
2
/s)

surface pressure, C
a
(d/dC
a
)

a
density of air at operating conditions (kg/m
3
)

d
density of dispersed phase (kg/m
3
)
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 407

g
gas density (kg/m
3
)

L
liquid density (kg/m
3
)

S
solid density (kg/m
3
)
volume fraction of dispersed phase

W
volume fraction of aqueous phase

WO
water to oil phase volume ratio
surface tension (mN/m)
1. Introduction
1.1. Alkanes as an attractive feedstock opportunity for
bioconversion
Gas-to-liquid and solid-to-liquid fuels processes are expand-
ing globally, resulting in an accumulation of associated alkane
by-products. Currently, alkanes are marketed at their fuel value,
but increasingly have been viewed as an exciting feed stock
opportunity for the production of a broad range of commod-
ity products. In particular, attention has been focused on the
potential of alkane feed stocks for commercial biotechnology
ventures. The large range of bacteria and fungi which can
efciently biofunctionalise these hydrocarbons under moderate
temperatures and pressures, and the considerable variety of ne
chemicals and versatile chemical intermediates which can be
produced, make the bioconversion route exceptionally attrac-
tive [13]. Chemicals which have been successfully produced
biologically fromalkane substrates include amino acids, organic
acids, carbohydrates, lipids, nucleic acids, vitamins, enzymes,
co-enzymes and antibiotics [2], polyhydroxyalkanoates [46],
biosurfactants [7] and dioic acids [8,9]. Developments in genet-
ically modied organisms which carry out these bioconversions
[10,11] further promote the biological route as an option with
extensive commercial potential.
Inadditiontoits wide rangingsuitabilityas a substrate for bio-
products in general, the alkane moiety confers specic process
advantages. Alkanes promote the accumulation of hydropho-
bic products (e.g. biosurfactants), products which require an
enhanced pool of acetyl-CoA as precursor (e.g. co-enzyme
Q) and products which are converted directly from the alkane
molecular structure and cannot be produced from a carbohy-
drate substrate alone (e.g. dioic acids, polyhydroxyalkanoates)
[2]. Moreover, the immiscible nature of the alkane substrate
provides prospects for in situ extraction of products that are
preferentially soluble in the organic phase. This is particularly
advantageous for processes in which the microbial growth is
inhibited by the product formed. For instance, in situ extraction
was used to decrease the alcohol concentration around the cell
in the aqueous phase [1214].
These considerations, together with the numerous microor-
ganisms well capable of converting alkanes to valuable products,
ranks biofunctionalisation of alkanes as a leading route to alkane
conversion, and further, emphasises the need to address engi-
neering complexities associated with process operation, so that
its potential can be optimised.
1.2. Adequacy of oxygen supply to alkane-based
bioprocesses
Current interest in the development of alkane-based biopro-
cesses has raised the question of an adequate supply of oxygen in
these media. Sufcient oxygen to satisfy the organisms demand
is crucial in any aerobic bioprocess if successful process oper-
ation is to be realised. Further, the role of the oxygen transfer
rate in ensuring optimal yields and productivities becomes pro-
portionately more important as the scale of operation increases
and a change in the limiting regime from kinetic to transport
control is likely [15]. Accordingly, the oxygen transfer rate has
traditionally been cited as the key criterion in the design and
scale up of aerobic bioprocesses [16].
In alkane-based bioprocesses, the difculty in supplying
adequate oxygen is exacerbated by the deciency of oxy-
gen in the molecular structure of the substrate. This means
that the oxygen requirement has to be met solely by trans-
fer to the culture, contrary to carbohydrate-based bioprocesses
where the substrate supplies approximately 66% of the oxy-
gen demand [1,17]. The proportionately higher requirement for
oxygen transfer, relative to carbohydrate processes, remains a
common feature of alkane metabolism and is likely to inu-
ence all alkane bioprocesses, irrespective of the chain length of
the alkane. This has been known since early studies showed
that the oxygen requirement for equivalent yeast growth on
hydrocarbon was almost triple that on carbohydrate [1719].
A 2.5-fold higher oxygen requirement per unit of biomass
of Candida petrophilum for growth on n-hexadecane relative
to growth on glucose provides further support [20]. Simi-
larly, a 2.5-fold higher oxygen requirement was evident for
Pseudomonas oleovorans growing on octane compared with
Escherichia coli growing on glucose at the same specic growth
rate [21].
The importance of an adequate oxygen transfer rate in
hydrocarbon-based bioprocesses for the enhancement of the
organisms potential for biofunctionalisation has been widely
recognised. The oxygen transfer rate, and especially the overall
volumetric oxygen transfer coefcient (K
L
a), have been exten-
sively documented in hydrocarbon-based systems. Nonetheless,
the response of K
L
a to changes in uid dynamics and uid
properties in hydrocarbonaqueous dispersions is still not well
understood.
Further, the prediction of K
L
a in hydrocarbonaqueous dis-
persions is complex. This highlights the importance of the
development of a fundamental understanding of the behaviour
of K
L
a in these systems, and the identication and quantication
of the parameters which underpin this behaviour so that exist-
ing predictive models may be successfully extended to provide
reliable estimates of K
L
a for hydrocarbon bioprocesses.
This review focuses on the oxygen transfer rate and K
L
a
in hydrocarbonaqueous dispersions and provides a basis for
the development of a knowledge platform for oxygen transfer
in alkane-based bioprocesses. Three types of K
L
a behavioural
trends in hydrocarbonaqueous dispersions have been quantied
over a wide range of operating conditions, reactor geometries
and hydrocarbon types and concentrations.
408 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
The key factors which underpin these behavioural trends have
been identied and their relative impact on oxygen transport in
these systems has been assessed. Particular attention has been
given to the inuence of hydrocarbon on bubble diameter, gas
hold up, gasliquid interfacial transfer area and liquid phase
oxygen transfer coefcient. Finally, consideration is given to
existing models and how these could be modied or extrapo-
lated to predict K
L
a for the design, operation and scale up of
hydrocarbon-based bioprocesses.
2. Oxygen transfer rate and overall volumetric oxygen
transfer coefcient in hydrocarbonaqueous dispersions
The two-lmmodel dening oxygen transfer in aqueous sys-
tems (Eq. (1)) is well documented [22,23]
OTR =
dC
o
dt
= K
L
a(C
o,sat
C
o
) (1)
Here the oxygen transfer rate is quantied in terms of the con-
centration driving force, or oxygen solubility, and K
L
a. K
L
a, a
composite parameter comprising the liquid phase oxygen trans-
fer coefcient (K
L
) and the gasliquid interfacial area per unit
volume (a) is frequently highlighted as the key criterion in
design, operation and scale up of aerobic bioprocesses [24]. In
aqueous systems, K
L
a is predicted according to Eqs. (2) or (3)
for stirred tank reactors (STRs) and bubble columns, or airlift
reactors, respectively [25] where the empirical constants ,
and vary with physicochemical properties [2628] and scale
[24].
K
L
a =
_
P
V
_

s
(2)
K
L
a = V

s
(3)
2.1. Oxygen solubility
Typically, oxygen solubilities in aqueous solutions are usu-
ally less than 10 mg/L [29]. On the other hand, oxygen
solubilities in hydrocarbonaqueous dispersions are signi-
cantly higher, due to the increased saturation level of oxygen
in pure hydrocarbon. There is, however, some lack of consensus
in the literature with respect to the actual values of the oxygen
solubility in hydrocarbons (Fig. 1). For instance, oxygen solu-
bility in n-C
12
alkane is generally reported as 260 mg/L [30],
324 mg/L [31] and 305 mg/L [32] at similar temperatures. Oxy-
gen solubility in the longer chain n-C
16
alkane was reported
as 322 mg/L [33], 190 mg/L [30], 340 mg/L [29], 255 mg/L
[32] and 271 mg/L [31] over a temperature range of 2230

C.
Nevertheless, when considering alkanes of chain lengths char-
acteristically employed in bioprocesses, the oxygen solubility
in the pure alkanes is more than 20 times that in water. At
chain lengths of n-C
8
, possibly the level below which substrate
inhibition would make any bioprocess impracticable, a 40-fold
increase in oxygen solubility over that in water is reported.
It should be noted, however, that the oxygen solubility in
n-dodecane does not follow the general trend. While oxygen
Fig. 1. Solubility of oxygen in pure alkanes at 1 atm and 2230

C. () Ju and
Ho [29]; () Ho et al. [33]; () Makranczy et al. [30]; () Blanc and Batiste
[32]; () Wilcock et al. [34]; () Thomsen and Gjaldbaek [42]; () Hesse et al.
[31]; () Rols et al. [35], Jia et al. [36] and Galaction et al. [38].
solubility in this chain length had been reported to lie between
260 mg/L and 324 mg/L, at temperatures ranging from 10

C to
40

C [3032,34], more recent studies have reported the solu-


bility of a commercially purchased n-dodecane as 54.9 mg/L at
35

C [3541]. It is not known why the latter should differ so


widely from the solubility data reported in the earlier literature
studies.
In hydrocarbonaqueous dispersions, the effect of increased
oxygen solubility in hydrocarbons on the oxygen transfer rate
may be considerable. Assuming applicability of the volumetric
relationship for oxygen solubility in hydrocarbonaqueous dis-
persions proposed by Ju and Ho [29], the increase in oxygen
solubility of a 10% n-hexadecaneaqueous suspension may be
three to fourfold that in an aqueous solution. This phenomenon
has prompted the modication of aqueous carbohydrate-based
media to carry more oxygen by hydrocarbon addition, thus ele-
vating the maximum oxygen transfer attainable. Ho et al. [33]
were one of the rst groups to propose the use of hydrocarbon
droplets as anenhancer of oxygentransfer incarbohydrate-based
bioprocesses. Oxygen transfer has been successfully enhanced
in Aspergillus niger cultures through addition of n-dodecane,
n-hexadecane and soybean oil during glucose oxidase produc-
tion [43] and through addition of n-dodecane during citric acid
production [44]. Oxygen transfer in Saccharomyces cerevisiae
cultures has similarly beneted from addition of n-C
12-16
alka-
nes, n-dodecane and peruorocarbon [36,45].
Although the strategy of using hydrocarbons to function as
oxygen vectors has wide appeal, and has been used to good
effect in bioprocesses utilising carbohydrate substrates, this does
not imply that bioprocesses which utilise hydrocarbon sub-
strates are invariably unlikely to be oxygen limited. Firstly,
the hydrocarbon-based process is compromised by the de-
ciency of oxygen in the molecular structure which leads to
a higher requirement for oxygen transfer, relative to that in
carbohydrate-based bioprocesses in which oxygen is also con-
tributed directly by the substrate. Secondly, the hydrocarbon
substrate may depress K
L
a. Regimes of depressed (less than
that in water) and enhanced (greater than that in water) K
L
a have
been dened according to agitation rate and alkane concentra-
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 409
tion [46]. In regions of depressed K
L
a, the resultant variation in
oxygen transfer rate will depend directly on the relative magni-
tudes of the K
L
a and the transfer driving force (i.e. the oxygen
solubility), and indirectly on the process conditions which dene
these magnitudes. Under conditions of depressed K
L
a, the ele-
vated oxygen solubility does not always compensate for the
increased requirement for oxygen transfer. Clarke et al. [46]
have reported a plateau in the oxygen transfer rate from 10%
n-C
12-13
alkane, when depressed K
L
a was observed, despite an
increase in the oxygen solubility with increasing alkane con-
centration. This study also showed that at lower agitation rates,
K
L
a depression was more likely, suggesting that the increased
solubility was less likely to offset the depression of K
L
a at the
lower agitation.
On the other hand, in regions of enhanced K
L
a, the higher
K
L
a coupled with the increased oxygen solubility in the hydro-
carbon doubly benets the oxygen transfer rate. Ho and Ju [47]
report both an increase in the K
L
a and the oxygen solubility
in n-hexadecane relative to water. Under these conditions, the
increased oxygen transfer rate is likely to compensate for the
greater requirement for oxygen transfer.
Consequently, despite the superior oxygen solubility in
hydrocarbons, and the effective use of hydrocarbons as oxy-
gen vectors in carbohydrate-based bioprocess, oxygen transfer
may still be suboptimal in hydrocarbon-based bioprocess. These
considerations underline the critical importance of the quanti-
cation of K
L
a behaviour in hydrocarbonaqueous dispersions,
and the development of a model to predict optimal transport rates
for the successful operation and scale up in these systems.
2.2. Trends in overall volumetric oxygen transfer
coefcient
In aqueous systems, where the behaviour of K
L
a is well
understood, the quantication and prediction of the oxygen
transfer rate has been established. The behaviour of K
L
a in
hydrocarbon containing systems, on the other hand, is consider-
ably less well understood, notwithstanding the numerous studies
conducted in hydrocarbonaqueous dispersions in systems with
differing reactor and impeller geometries, over a wide range of
operating conditions, hydrocarbon moieties and concentration
ranges.
These studies have invariably shown that the hydrocarbon
fraction impacts markedly on K
L
a, although paradoxically, not
always in the same manner (Tables 1a1c). The manner in which
the hydrocarbon affects K
L
a differed widely, and depended on,
inter alia, the reactor type and the uid properties of the aqueous
and hydrocarbon phases. Three distinct trends in K
L
a behaviour
in hydrocarbonaqueous dispersions can be dened. Firstly, a
K
L
a trend with an increase on hydrocarbon addition to a max-
imum value, and subsequently, a decrease on further addition
of hydrocarbon, has been rmly established (type 1 behaviour).
Secondly, a K
L
a increase on hydrocarbon addition but with no
subsequent decline has also been frequently reported (type 2
behaviour). Thirdly, no increase in K
L
a has with been observed
with hydrocarbon addition, with K
L
a either being maintained
constant or decreasing (type 3 behaviour).
Those studies which exhibit type 1 behaviour, report a char-
acteristic K
L
a peak associated with a specic system-dependent
hydrocarbon concentration (Table 1a). In STRs, where the aque-
ous phase comprised either pure water or a salt solution, peak
K
L
a was observed at 5% n-C
1213
alkane [46], 10% n-C
1013
alkane [48], 10% n-hexadecane [49], 1% silicone oil and 2%
soybean oil [50].
Similar trends were observed in this reactor type when the
aqueous phase contained cells. K
L
a peaks were reported in S.
cerevisiae cultures at 1% or 2% lard oil, depending on the agi-
tation rate, and 1% olive oil [51]. In E. coli cultures, peak K
L
a
occurred at 30% PFC-40 [52]. Aerobacter aerogenes cultures
exhibited peak K
L
a at 23% n-dodecane [35] and 15% soybean
oil [53].
Type 1 behaviour was also observed in airlift and bub-
ble column reactors in both cell-culture and cell-free systems.
K
L
a peaks were reported in S. cerevisiae cultures at 3% n-
dodecane [36], 2% PFC-40 [36] and 4% n-C
1216
[45], and
in A. niger cultures, at 5% n-dodecane [37]. In cell-free bub-
ble columns K
L
a peaks were reported at 0.5% hexanol [54]
and 6% silicone oil [55]. A K
L
a peak was even reported in
soybean oilwater dispersion at 2% in a shake ask culture
[56].
Although a hydrocarbon concentration associated with peak
K
L
a can be identied, this concentration varies considerably
between studies. For instance, in STRs, the hydrocarbon con-
centration associated with the K
L
a peak was lower (12%) in
silicone oil [50], soybean oil [50], lard oil [51] and olive oil [51]
than that (530%) in n-C
1213
[46], n-C
1013
[48], n-hexadecane
[49], PFC-40 [52] and n-dodecane [35]. In the studies where the
peak occurred at a lower hydrocarbon concentration, the viscos-
ity of the hydrocarbon phase was relatively high at 510 cP for
silicone oil [55,64], 41 cP for soybean oil [53], 53 cP for lard oil
[51] and 58 cP for olive oil [51]. In the studies where the peak
occurred at a comparatively higher hydrocarbon concentration,
the viscosity of the hydrocarbon phase was about a order of
magnitude lower, viz. 3 cP for n-hexadecane [55,64,65], 34 cP
for PFC-40 [52,65] and 1.31.5 cP for n-dodecane [55,59,65],
with the viscosities of n-C
1213
and n-C
1013
likely to be in a
similar range. This suggests that the hydrocarbon concentration
at which the K
L
a will peak depends on the dispersion viscosity
such that an increased dispersion viscosity will tend to decrease
the K
L
a, resulting in a decline in the K
L
a at a lower concentration
of hydrocarbon.
It is unlikely, however, that a variation in uid properties
brought about by hydrocarbon addition is the sole contributing
factor accounting for the discrepancy in hydrocarbon concentra-
tions at peak K
L
a. Clearly, system geometry also plays a major
role. In STRs, K
L
a was observed to peak at both 2% soybean oil
and 15% soybean oil [53], notwithstanding the same viscosity
of 41 cP. Further, K
L
a peaked at signicantly lower hydrocar-
bon concentrations in column reactors than in STRs with the
same hydrocarbon type. For example, when comparing cell cul-
tures in different reactor congurations, K
L
a peaked at 35%
n-dodecane [36,37] and 2% PFC-40 [36] in column reactors,
compared with 23% n-dodecane [35] and 30% PFC-40 [52] in
STRs.
410 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
Type 2 behaviour was also demonstrated in numerous studies
(Table 1b). Here one of two trends was evident. Either K
L
a
continued to increase up to the last hydrocarbon concentration
examined, or increased to a maximum value at which a plateau
was reached. In both of these trends, no K
L
a peak was observed.
The absence of a peak K
L
a may be explained in part by the lack
of K
L
a data at hydrocarbon concentrations where the peak value
could conceivably occur.
In the rst case, where K
L
a continued to increase to that at the
maximum hydrocarbon concentration examined, it is possible
that the K
L
a peak may yet have occurred at a higher hydro-
carbon concentration where data had not been acquired. These
studies include STR congurations where the maximum K
L
a
was observed at 1% castor oil [51], 20% toluene [57,58] as well
as bubble column congurations where the maximum K
L
a was
observed at 10%n-decane, 10%n-heptane and 10%n-dodecane
Table 1a
Behaviour of K
L
a in aerated hydrocarbonaqueous dispersions: type 1 behaviour (K
L
a peak)
Reactor type Aqueous phase Hydrocarbon phase Hydrocarbon
concentration (v/v)
Hydrocarbon
concentration (v/v)
at maximum K
L
a
Reference
STR Distilled water n-C
1213
020% 5% Clarke et al. [46]
STR Distilled water n-C
1013
020% 10% Correia et al. [48]
STR Distilled water n-Hexadecane 033% 10% Nielsen et al. [49]
STR 0.2 M Na
2
SO
4
Silicone oil 110% 1% Mor ao et al. [50]
STR 0.2 M Na
2
SO
4
Soybean oil 110% 2% Mor ao et al. [50]
STR Saccharomyces cerevisiae Lard oil 04% 12% Liu et al. [51]
STR S. cerevisiae Olive oil 03% 1% Liu et al. [51]
STR Escherichia coli K12 wild type PFC-40 040% 30% McMillan and Wang [52]
STR Aerobacter aerogenes NRRL B199 n-Dodecane 033% 23% Rols et al. [35]
STR A. aerogenes NRRL B199 Soybean oil 030% 15% Rols and Goma [53]
Airlift reactor S. cerevisiae AY-12 n-Dodecane 015% 3% Jia et al. [36]
Airlift reactor S. cerevisiae AY-12 PFC-40 07% 2% Jia et al. [36]
Airlift reactor Aspergillus niger n-Dodecane 014% 5% Jianlong [37]
Tower reactor S. cerevisiae AY-12 n-C
1216
08% 4% Jia et al. [45]
Bubble column Distilled water Hexanol 01.5% 0.5% Koide et al. [54]
Bubble reactor Distilled water Silicone oil 47V5 010% 6% Dumont et al. [55]
500 mL ask Distilled water Soybean oil 04% 2% Jia et al. [56]
Table 1b
Behaviour of K
L
a in aerated hydrocarbonaqueous dispersions: type 2 behaviour (K
L
a increase)
Reactor type Aqueous phase Hydrocarbon phase Hydrocarbon
concentration (v/v)
Hydrocarbon
concentration (v/v)
at maximum K
L
a
Reference
STR S. cerevisiae Castor oil 01% 1% Liu et al. [51]
STR Carboxymethyl cellulose/salt solutions n-Dodecane 020% 20% Cascaval et al. [39]
STR Carboxymethyl cellulose/salt solutions n-Dodecane 020% 20% Galaction et al. [38]
STR Distilled water Toluene 020% 20% Cents [57]
STR Distilled water Toluene 120% 20% Yoshida et al. [58]
Bubble column Distilled water n-Decane 010% 10% Kundu et al. [59]
Bubble column Distilled water n-Dodecane 010% 10% Kundu et al. [59]
Bubble column Distilled water n-Heptane 010% 10% Kundu et al. [59]
Bubble reactor Distilled water n-Hexadecane 010% 10% Dumont et al. [55]
Bubble reactor Distilled water PFC-40 04% 4% Dumont et al. [55]
Bubble reactor Distilled water Silicone oil 010% 10% Dumont et al. [55]
STR Carboxymethyl cellulose/salt solutions n-Dodecane 020% 10%
a
Cascaval et al. [39]
STR Carboxymethyl cellulose/salt solutions n-Dodecane 020% 10%
a
Galaction et al. [38]
STR Distilled water Ethanol 00.8% 1%
a
Bi et al. [60]
STR Tap water n-Dodecane 010% 1%
a
da Silva et al. [40]
STR Distilled water/salt solutions n-Dodecane 010% 6%
a
Hassan and Robinson [61]
STR Distilled water/salt solutions n-Hexadecane 010% 6%
a
Hassan and Robinson [61]
STR Distilled water n-Dodecane 05% 5%
a
Wong and Shiuan [24]
STR Distilled water Oleic Acid 0.0122% 8%
a
Yoshida et al. [58]
Bubble column Distilled water Toluene 010% 10%
a
Kundu et al. [59]
Bubble column Distilled water 2-Ethyl-1-hexanol 010% 10%
a
Kundu et al. [59]
Bubble reactor Distilled water n-Dodecane 010% 6%
a
Dumont et al. [55]
a
Plateau reached.
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 411
Table 1c
Behaviour of K
L
a in aerated hydrocarbonaqueous dispersions: type 3 behaviour (K
L
a decrease or K
L
a constant)
Reactor type Aqueous phase Hydrocarbon phase Hydrocarbon
concentration (v/v)
K
L
a constant
or decreased
Reference
STR Distilled water Ethanol 00.8% Constant Bi et al. [60]
STR Distilled water PFC-40 010% Decreased Ces ario et al. [62]
STR Distilled water n-C
12-13
020% Decreased Clarke et al. [46]
STR Distilled water n-C
10-13
020% Decreased Correia et al. [48]
STR S. cerevisiae Olive oil 03% Constant Liu et al. [51]
STR S. cerevisiae Castor oil 01% Constant Liu et al. [51]
STR Distilled water n-Hexadecane 033% Decreased Nielsen et al. [63]
STR Distilled water n-Hexadecane 033% Decreased Nielsen et al. [49]
STR Distilled water Parafn 121% Decreased Yoshida et al. [58]
Bubble column Distilled water Heptanol 01% Decreased Koide et al. [54]
Bubble column Distilled water Octanol 00.5% Decreased Koide et al. [54]
[59] and 10% n-hexadecane, 4% PFC-40 and 10% silicone oil
[55].
In the second case, where a plateau has been dened, a K
L
a
peak may not have been detected if there were large intervals
between data points. For example, da Silva et al. [40] report
a maximum K
L
a at 1% n-dodecane, and essentially the same
value at 10% n-dodecane. However, the lack of data between
1% and 10% n-dodecane means it cannot be said with certainty
that a peak K
L
a would not have been obtained at an intermediate
hydrocarbon concentration.
Nevertheless, the lack of data at critical hydrocarbon con-
centrations does not necessarily account for the absence of the
characteristic K
L
a peak in all cases. In several studies, a K
L
a
plateau was rigorously demonstrated, with no evidence of a
subsequent K
L
a decrease. Moreover, even in studies where a
plateau was not detected, the absence of a peak cannot invari-
ably be attributed simply to insufcient data, as evidenced by the
K
L
a maximumin a dispersion of n-dodecane and carboxymethyl
cellulose (CMC)/salt solution [38,39]. Here, maximal K
L
a was
observed at 20% n-dodecane, the highest concentration exam-
ined, when the dispersion viscosity was 100330 cP. However,
at a reduced viscosity of 1040 cP, a K
L
a plateau was reached
at 10% n-dodecane.
In addition, K
L
a plateaux where also recognised in water at
6% n-dodecane [61], 6% n-hexadecane [61], 5% n-dodecane
[24] and 8% oleic acid [58] in STRs. In bubble columns, K
L
a
plateaux were similarly documented at 10% toluene and 10%
2-ethyl-1-hexanol [59] and at 6% n-dodecane [55]. As with the
studies in which K
L
a peak was reported, these studies show a
dependence of the hydrocarbon concentration at which maximal
K
L
a was attained and on the type of hydrocarbon and reactor
conguration used. A maximal K
L
a at 6% n-dodecane both in
a STR [61] and a bubble column [55], however, suggests a less
marked inuence of the reactor conguration and operation in
these studies.
In yet other studies, type 3 behaviour occurred, where no
increase in K
L
a was observed, and a decrease often reported
(Table 1c). In STRs this trend may be partly ascribed to a low
agitation rate. Bi et al. [60] reported a constant K
L
a in an aque-
ous ethanol solution at 135 rpm, while at 300600 rpm, K
L
a
increased to reach a plateau at 1% ethanol. Similarly, Clarke
et al. [46] and Correia et al. [48] report a decrease in K
L
a in
n-C
1213
alkane and n-C
1013
alkane, respectively, at 600 rpm,
although K
L
a peaks at 510% were dened at agitation rates
between 800 and 1200 rpm, over the same hydrocarbon concen-
tration range. Similarly, Liu et al. [51] document an essentially
K
L
a constant at 200 rpm, while K
L
a increased to 1% olive oil
and castor oil at 400 rpm.
Nielsen et al. [49], on the contrary, showed a decrease in
K
L
a with increasing n-hexadecane to 33% at agitation rates of
600 and 800 rpm, and a K
L
a peak at 10% n-hexadecane at the
comparatively lower agitation of 400 rpm. Earlier studies by this
group [63] reported a decreased K
L
a at 400 rpmas well as at 600
and 800 rpm. This suggests that, in addition to lower agitation
rates, K
L
a will be negatively inuenced by other parameters.
Certainly, K
L
a behaviour has been inuenced by hydrocarbon
type. K
L
a was observed to decrease with parafn addition at an
agitation rate of 500 rpm, but increase with toluene and oleic
acid addition at the relatively lower agitation rate of 350 rpm
[58]. Further support for the inuence of hydrocarbon type is
provided by the increased K
L
a with n-hexanol addition, while
n-heptanol and n-octanol addition both resulted in a decreased
K
L
a under the same conditions [54].
Notwithstanding the importance of dening the trend of K
L
a
behaviour with hydrocarbon addition, consideration should also
be given to the absolute value of K
L
a in these systems, relative
to that in the absence of hydrocarbon, i.e. whether hydrocar-
bon addition results in K
L
a depression or enhancement relative
to that in water. Generally, where K
L
a decreases with hydro-
carbon addition, depression is observed and correspondingly,
where K
L
a increases with hydrocarbon addition, enhancement
occurs.
The latter, however, does not always hold, and depres-
sion may occur even when K
L
a increases with hydrocarbon
addition. It has been found, in both stirred tank and bubble
reactors, that the addition of hydrocarbon at low concentra-
tions (less than 2.5%) sometimes yields an initial decrease in
K
L
a to below the level attained in the absence of hydrocar-
bon (K
L
a*) (Figs. 2 and 3). A subsequent increase in K
L
a on
further hydrocarbon addition may compensate for the initial
decrease, leading to an enhanced K
L
a at higher hydrocarbon
concentrations (Fig. 2). However, this increase is not invariably
412 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
Fig. 2. Depression and subsequent enhancement of overall volumetric oxygen
transfer coefcient by hydrocarbon. () Cents [57], toluene; () Yoshida et al.
[58], toluene; () Dumont et al. [55], n-dodecane; () Kundu et al. [59], toluene;
(+) Kundu et al. [59], 2-ethyl-1-hexanol.
sufcient to offset the initial decrease, when a depressed K
L
a
results (Fig. 3). Depression appears to be inuenced by both
the agitation and the hydrocarbon moiety; depression occurred
with silicone oil but not n-dodecane [55] and with agitation at
800 rpm but not 1000 rpm or 1200 rpm [48].
The behaviour of K
L
a in these studies suggests that the
uid properties (as determined by the type and concentra-
tion of the hydrocarbon) as well as the reactor geometry
and operating parameters (especially agitation) strongly inu-
ence K
L
a behaviour in hydrocarbonaqueous dispersions. It
is likely that these parameters do not act independently, but
that multiple causative factors relate to K
L
a behaviour and fur-
ther, that the predominance of each of these factors varies,
depending on the prevailing process conditions. If a predictive
model to dene locales of optimum K
L
a is to be developed in
hydrocarbonaqueous dispersions, it is essential to acquire a
fundamental understanding of the factors which underpin the
behaviour of K
L
a, and the complex interaction of the factors, in
these systems.
Fig. 3. Depression of overall volumetric oxygen transfer coefcient by hydro-
carbon. () Correia et al. [48], n-C
1013
; () Dumont et al. [55], silicone oil
47V5; () Dumont et al. [55], silicone oil 47V10; () Wong and Shiuan [24],
n-dodecane; (+) Yoshida et al. [58], oleic acid.
3. Factors inuencing overall volumetric oxygen
transfer coefcient in hydrocarbonaqueous dispersions
In hydrocarbonaqueous dispersions, K
L
a is signicantly
inuenced by turbulence and uid properties, the extent of this
inuence being ultimately dened by the operating conditions
and hydrocarbon type and concentration. Here the inuences of
turbulence and uid properties on interfacial area and K
L
are
discussed separately; nevertheless, cognisance is taken of their
interactive inuence on both these parameters.
3.1. Inuence of turbulence on interfacial area and oxygen
transfer coefcient
Turbulence is essential in promoting oxygen transfer and with
its positive effect both on the interfacial area and K
L
, is the
key to supplying adequate oxygen. Traditionally, turbulence has
been promoted through increased energy input, as dened by
power input per unit volume (P/V), increased impeller speed
(N) or increased aeration rate (V
s
). Turbulence is often further
enhanced by the modication of vessel and impeller geometries.
3.1.1. Inuence of turbulence on interfacial area
The inuence of turbulence on interfacial area has been doc-
umented in a wide variety of uids and over a broad range
of operating conditions in systems with differing reactor and
impeller geometries. Empirical models dening the contribu-
tion made by turbulence, through the parameters P/V, N and V
s
,
to the interfacial area are listed in Table 2.
These correlations exhibit a range of dependencies of inter-
facial area on P/V, N and V
s
. The exponents of P/V varied from
a low of 0.25 in water [70] to highs of 0.9 and 1.06 in elec-
trolyte solutions [69,71,72] and electrolyte/polymer solutions
[82]. In uids containing hydrocarbons, the exponent was rela-
tively small at 0.4 [26,75]. Similarly, a wide range of exponents
of N and V
s
is reported, with exponents of N from 0.7 in an
electrolyte solution [66] to 2.55 in a polymer solution [79] and
exponents of V
s
from 0.06 in a polymer solution [79] to 0.9 in a
hydrocarbon solution [27].
The inuences of P/V, N and V
s
on interfacial area are, how-
ever, not independent. When comparing the interaction of P/V
and V
s
for instance, the inuence of V
s
is generally less at high
P/V[71,72,76,78], than at comparatively lower P/V[70]. Conse-
quently, whenassessingthe inuence of turbulence oninterfacial
area, the interaction between P/V, N and V
s
, as well as their
individual inuences, needs to be considered.
The positive exponents on P/V, N and V
s
in these correla-
tions nevertheless conrm the benecial impact of turbulence
on the interfacial area in all uids and systems examined. Turbu-
lence increases the interfacial area available for transfer through
increased shear which modies the coalescence and break up of
the gas bubbles, thereby reducing the bubble size. The reduced
bubble size results in an increased residence time of the gas in
the liquid, thus concomitantly increasing the gas hold up. The
effect of bubble size and bubble hold up on interfacial area is
dened by Eq. (4) which relates interfacial area per unit volume
to the volume-surface mean diameter, or Sauter mean diameter
K
.
G
.
C
l
a
r
k
e
,
L
.
D
.
C
.
C
o
r
r
e
i
a
/
B
i
o
c
h
e
m
i
c
a
l
E
n
g
i
n
e
e
r
i
n
g
J
o
u
r
n
a
l
3
9
(
2
0
0
8
)
4
0
5

4
2
9
4
1
3
Table 2
Empirical correlations for the prediction of interfacial area
System Operating conditions Liquid phase(s) Equation Reference
STR N=2001200 rpm; V
s
=0.3050.61 cm/s Aqueous: ethanol, methanol, i-propanol,
n-butanol, ethylene glycol, CCl
4
; ethyl
acetate, nitrobenzene; toluene
a = 1.44
_
(P/V)
0.4

0.2
L

0.6
_
_
Vs
Vt
_
0.5
Calderbank [26]
Sieve plates V
s
=6.160.1 cm/s Aqueous: aliphatic alcohols, glycol a = 0.38
_
Vs
Vt
_
0.775
_
Vs
L
(Ns/As)Ds
L
_
0.125
_

L
g
Ds
_
1/3
Calderbank [27]
STR N=60400 rpm;
V
s
=0.16930.762 cm/s
Aqueous:
NaOH
a (ND
t
)
1.1
V
0.75
s
Yoshida and Miura
[66] STR a (ND
t
)
0.70.9
STR N=255, 325 & 455 rpm; Q
G
=0.68 m
3
/h Aqueous: NaCl, sodium dodecyl
sulphate, dow corning antifoam C
a N
1.75
Benedek and Heideger [67]
Bubble column V
s
=0.9254.28 cm/s Aqueous: glycol, methanol, CCl
4
a =
1
3Dt
_
gD
2
t

L

_
0.5
_
gD
3
t
v
2
L
_
0.1

1.13
G
Akita and Yoshida [68]
STR N=8002000 rpm; V
s
=0.457 cm/s Aqueous: Na
2
SO
3
, KOH, K
2
CO
3
a (P/V)
0.891.06
Robinson and Wilke [69]
STR N=250500 rpm; V
s
=0.6341.27 cm/s Aqueous a = 593(P/V)
0.25
V
0.75
s
de Figueiredo and Calderbank [70]
STR N=7321900 rpm; V
s
=0.371.11 cm/s Aqueous: Na
2
SO
4
; KOH; KCl; K
2
SO
4
a (P/V)
0.850.9
V
0.39
s
Hassan and Robinson [71]
STR N=7321900 rpm; V
s
=0.371.11 cm/s Aqueous: Na
2
SO
4
, KOH, KCl, K
2
SO
4
a (P/V)
0.850.9
V
0.390.42
s
Hassan and Robinson [72]
Review a = 3.00
_

L
g

1/2
_
Q
G
NV

5/12
_
N
2
D
4
t
gqV
2/3
_
0.658
_

L
d
b
N
2
D
4
t
V
2/3
_
3/16
Hughmark [73]
Bubble column V
s
=020 cm/s Aqueous: CMC a = 4.65 10
2
V
0.51
s

0.51
L
Schumpe and Deckwer [74]
STR N=4801800 rpm; V
s
=15 cm/s Cyclohexane a = 1.44
_
(P/V)
0.4

0.2
L

0.6
_
_
Vs
Vt
_
0.5
_
Et
P
__
g
a
_
0.16
Sridhar and Potter [75]
STR N=400850 rpm; Q
G
=0.120.48 m
3
/h Aqueous: NaCl a (P/V)
0.62
V
0.45
s
Ho et al. [76]
Bubble column V
s
=27 cm/s Aqueous: CMC 7H4,
carboxypolymethylene;
a = 6C
_

0.4

0.6
L

0.6
_

1
G
_

d
c
_
0.25
Kawase et al. [77]
Stirred slurry reactor P/V=30010000 W/m
3
; V
s
=0.344.6 cm/s Aqueous: CMC, Na
2
SO
3
-glass beads a = 2.87(P/V)
0.76
V
0.34
s

0.18
L
Schmitz et al. [78]
a = 3.17(P/V)
0.39
V
0.48
s

1.11
L
STR N=7501500 rpm; Q
G
=0.127.26 m
3
/h Aqueous a = 0.14N
0.77
Q
0.28
G
Al Taweel and Cheng [79]
Aqueous: PGME a = 10
23.9
N
2.55
Q
0.06
G

15.4
STR N=180 rpm; Q
G
=5.90424.7 m
3
/h Aqueous: NaCl a N
0.85
(P/V)
0.3
V
0.40.5
s
Barigou and Greaves [80]
Prediction models a N
1.6
(P/V)
0.5
V
0.65
s

0.3
L
Garcia-Ochoa and Gomez [81]
STR N=01130 rpm; V
s
=1854 cm/s Aqueous: Na
2
SO
3
a (P/V)
0.863
Linek et al. [82]
Aqueous: Na
2
SO
3
, Sokrat a (P/V)
0.913
Aqueous: Na
2
SO
3
, CMC TS.5 a (P/V)
0.632
Aqueous: Na
2
SO
3
, CMC TS.20 a (P/V)
0.689
Aqueous: Na
2
SO
3
, ocenol a (P/V)
0.492
Aqueous: Na
2
SO
3
, PEG 100 a (P/V)
0.903
414 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
(D
32
), through the fractional gas hold up (
G
).
a =
6
G
D
32
(4)
Empirical equations relating turbulence to D
32
(Table 3) and

G
(Table 4), similar to those relating turbulence to interfacial
area, have been proposed.
Turbulence, induced through P/V and N, both resulted
in decreased d
b
or D
32
. d
b
or D
32
(P/V)
0.4
has been
found to be similar in systems containing hydrocarbons
[26,83,84,85,89,91,92] and electrolytes [84]. Alves et al. [95]
found the effect of P/V on D
32
to differ in the turbine and bulk
regions of the reactor. In electrolyte solutions (or noncoalescing
systems) D
32
correlated with (P/V)
0.52
in the turbine region
and (P/V)
0.37
in the bulk region, which corresponds well with
the hydrocarbon and electrolyte systems in which a correlation
with (P/V)
0.4
was observed. On the other hand, D
32
was pro-
portional to (P/V)
0.24
in the turbine region and (P/V)
0.14
in
the bulk region suggesting a relatively greater inuence of tur-
bulence on D
32
in a hydrocarbon system relative to an aqueous
system.
The effect of agitation on D
32
varied, with exponents of N
of 0.76 to 1.4 [96] in hydrocarbon containing systems com-
pared with 0.1 to 0.3 in electrolyte solutions [66]. Overall,
increased agitation shifted the bubble size distribution towards
a smaller D
32
, and further, improved the uniformity of bubble
size distribution, particularly at the lower gas ow rates [114].
The mechanism by which N inuences D
32
is attributed to the
increase inentrainedgas bubbles whichreachthe impeller blades
[100].
Hu et al. [96] used various dispersed phases to bring about
different owregimes in a stirred reactor under constant aeration
and agitation. Addition of decanol, diethylene glycol and cyclo-
hexanol as solutes resulted in D
32
proportional to N
1.4
D
2.8
i
.
Use of butanol and propanol caused turbulent ow resulting in
D
32
proportional to N
0.76
D
1.52
i
, i.e. with a comparatively
lesser dependence of bubble size on impeller speed.
The inuence of agitation on D
32
, however, reaches a maxi-
mum after which D
32
remains constant with further increases in
agitation. Calderbank [26] and Westerterp et al. [115] showed
that, in their system, D
32
reached a constant minimum value
between 1200 rpm and 2300 rpm. The value of the minimum
D
32
has been found to be dependent on the solute. Bi et al.
[60] observed a minimum D
32
around 1.8 mm with pure water,
whereas, the addition of more than 1 g/L ethanol lowered this
value to 0.8 mm. An analogous inuence of agitation on inter-
facial area has been observed where N corresponding to the
maximum interfacial area correlated well with that correspond-
ing to the minimum D
32
[28,75,116].
The effect of V
s
on D
32
, on the contrary, is generally opposite
to that of P/V and N. D
32
V
0.10.4
s
holds in stirred agitators
[90], otation cells [88] and bubble columns [87]. Only Akita
and Yoshida [68], working in three different bubble columns,
found the bubble size to decrease with increasing V
s
. The
increase in D
32
with increased V
s
is a probable consequence of
the reduction in turbulence as aeration reduces the mechanical
power dissipation in the liquid through the growth of gas cavities
behind the impeller blades, and through the dampening of veloc-
ityuctuations inthe uid[114]. Inaddition, a higher probability
of bubble collision and coalescence because of greater gas hold
up is likely to lead to an increase in D
32
[114,117,118].
In contrast, gas hold up is positively inuenced by increased
V
s
(Table 4), which probably accounts for the increase in inter-
facial area with increased V
s
, despite the frequent decrease
in D
32
. Hold up is enhanced in STRs with both increasing
P/V and V
s
[26,70,76,85,100,109,110,111]. In these reactors,
similar dependencies on P/V and V
s
were reported in water
[70,100,111], electrolyte solutions [76,110], cyclohexane [85]
and in hydrocarbonaqueous dispersions [26,109]. The expo-
nents of P/V and V
s
varied from 0.25 to 0.85 and 0.2 to
1 respectively, over widely differing impeller geometries and
scale, suggesting a lesser role of system geometry in dening
hold up. The dependence on P/V (exponent 0.8) in an air-lift
reactor is by comparison, considerably larger [112]. Increased
hold up similarly results from increased agitation [66,98,103]
and exponents of Nranging from0.6 to 1.04 have been reported.
The improved gas hold up, through increased P/V and N,
results from an increased dispersion of small bubbles within
the system, with slower bubble rise velocities [79]. The longer
residence time of smaller bubbles allows more time for the oxy-
gen to transfer to the liquid phase [116]. On the other hand, a
larger V
s
increases the bubble collision frequency leading to a
higher coalescence [117] with correspondingly larger bubbles.
The preferential release of larger bubbles may result in greater
proportion of smaller bubbles in the uid and consequently, a
larger gas hold up.
In bubble columns, the uid dynamic characterization has a
signicant effect on the operation and performance. Here the
exponents depend strictly on the ow regime prevailing in the
column. Three types of ow regimes are commonly observed
and are classied according to the prevailing V
s
, viz. the homo-
geneous (bubbly ow), the heterogeneous (churn-turbulent) and
the slug ow regimes. The bubbly ow regime and churn turbu-
lent regimes are observed at low and high V
s
, respectively. The
slug ow regime is of lesser importance, being limited solely to
small diameter columns at high V
s
. The transition from bubbly
to churn-turbulent ow depended, inter alia, on the liquid phase
and reactor type. Transitional phases were reported at V
s
values
of 0.10.2 m/s in a slurry bubble column with an parafn oil-
silica suspension [119], 0.20.7 m/s in water and cyclohexane
[120] and 0.080.2 m/s in water, tetradecane, parafn and tellus
oil [121].
The bubbly ow regime is characterised by relatively small
uniform bubbles sizes and slow rise velocities. There is neg-
ligible bubble coalescence or break up, thus bubble size is
essentially dictated by the sparger design and system proper-
ties. Generally it is found that hold up in this regime increases
with increasing V
s
with the exponent varying from 0.7 to
1 in both electrolyte solutions and hydrocarbon dispersions
[99,105,107,113]. Kelkar et al. [104] reported an exponent of
0.58 in alcohol dispersions but show that hold up increased with
increasing carbon chain length.
The churn-turbulent regime exhibits unsteady ow patterns
with enhanced turbulent motion of the bubbles and high liq-
K
.
G
.
C
l
a
r
k
e
,
L
.
D
.
C
.
C
o
r
r
e
i
a
/
B
i
o
c
h
e
m
i
c
a
l
E
n
g
i
n
e
e
r
i
n
g
J
o
u
r
n
a
l
3
9
(
2
0
0
8
)
4
0
5

4
2
9
4
1
5
Table 3
Empirical correlations for the prediction of bubble diameter
System Operating conditions Liquid phase(s) Equation Reference
Theoretical models D
32
= 2

0.6
(P/V)
0.4

0.2
L
Hinze [83]
STR N=2001200 rpm; V
s
=0.3050.61 cm/s Aqueous: ethanol, methanol, i-propanol,
n-butanol, ethylene glycol, CCl
4
, ethyl
acetate, nitrobenzene, toluene
D
32
= 4.15
_

0.6
(P/V)
0.4

0.2
L
_

0.5
G
+0.09 Calderbank [26]
STR N=60400 rpm;
V
s
=0.16930.762 cm/s
Aqueous:
NaOH
D
32
N
0.3
Yoshida and Miura
[66] STR D
32
N
0.1
STR N=30600 rpm; V
s
=0.360.91 cm/s Aqueous: NaCl, Na
2
SO
4
d
b
= 4.25

0.6
[(P/V)(1/(1
G
))]
0.4

0.2
L

0.5
G
Lee and Meyrick [84]
Bubble column V
s
=0.9254.28 cm/s Aqueous: glycol, methanol, CCl
4
D
32
Dt
= 26
_
gD
2
t

L

_
1/2
_
gD
3
t
v
2
L
_
0.12
_
Vs

gDt
_
0.12
Akita and Yoshida [68]
STR N=4801800 rpm; V
s
=15 cm/s Cyclohexane D
32
(P/V)
0.4
Sridhar and Potter [85]
Bubble column V
s
=0.266.7 cm/s Aqueous: glycerol-glass beads D
32
= 0.59
(V
D
/
G
)
2
g
Fukuma et al. [86]
Bubble column V
s
=8cm/s Aqueous: sucrose, ethanol-Ca alginate D
32
= 0.87
_
gD
2
t

L

_
0.44
_
gD
3
t
v
2
L
_
0.1
_
Vs

gDt
_
0.39
exp(2.0
S/L
) Sun and Furusaki [87]
Flotation cell Q
G
=0.096 m
3
/h Aqueous D
32
V
0.4
s
OConnor et al. [88]
STR N=125800 rpm; V
s
=0.0250.125 cm/s Aqueous: 4-methyl 2-pentanol D
32
= 2.0

0.6
(P/V)
0.4

0.2
L
Parthasarathy et al. [89]
STR Aqueous D
32
N
0.5
V
0.1(region of the impeller)
s
Takahashi et al. [90]
STR N=1251000 rpm; V
s
=0.0250.125 cm/s Aqueous: 4-methyl 2-pentanol D
32
(P/V)
0.4
Parthasarathy and Ahmed [91]
STR N=770 rpm; Q=1.1 vvm Aqueous: Na
2
SO
4
, ethanol, n-butanol D
32
N
1.2
(P/V)
0.4
Machon et al. [92]
STR D
32
( )
0.4
()
0.6
(
L
)
0.6
Pacek et al. [93]
STR N=300450 rpm; Q=0.25 and 5 vvm Aqueous: Na
2
SO
4
, PEG D
32
= C
_

0.6
(P/V)
0.2

0.6
L
_
Alves et al. [94]
STR N=300450 rpm; Q=0.25 and 5 vvm Aqueous: Na
2
SO
4
, PEG D
32
(P/V)
0.52 (turbine) and 0.37 (bulk)
Alves et al. [95]
Tap water D
32
(P/V)
0.24 (turbine) and 0.14 (bulk)
Prediction models d
b


0.6
(P/V)
0.4

0.2
L
Garcia-Ochoa and Gomez [81]
STR = 17 47 W/kg Aqueous: decanol, diethylene glycol,
cyclohexanol
D
32
= 2.72D
i
_

L
N
2
D
3
i

_
0.70
(transitional ow) Hu et al. [96]
Aqueous: propanol, butanol D
32
= 0.113D
i
_

L
N
2
D
3
i

_
0.38
(turbulent ow)
4
1
6
K
.
G
.
C
l
a
r
k
e
,
L
.
D
.
C
.
C
o
r
r
e
i
a
/
B
i
o
c
h
e
m
i
c
a
l
E
n
g
i
n
e
e
r
i
n
g
J
o
u
r
n
a
l
3
9
(
2
0
0
8
)
4
0
5

4
2
9
Table 4
Empirical correlations for the prediction of gas holdup
System Operating conditions Liquid phase(s) Equation Reference
STR N=2001200 rpm; V
s
=0.3050.61 cm/s Aqueous: ethanol, methanol, i-propanol,
n-butanol, ethylene glycol, CCl
4
; ethyl
acetate, nitrobenzene, toluene

G
=
_
Vs G
Vt
_
0.5
+0.0216
_
(P/V)
0.4

0.2
L

0.6
_
_
Vs
Vt
_
0.5
Calderbank [26]
STR N=60400 rpm;
V
s
=0.16930.762 cm/s
Aqueous:
NaOH

G
N
0.8
V
0.75
s
Yoshida and Miura
[66] STR
G
N
0.60.8
Bubble column V
s
=0.445 cm/s Aqueous: kerosene, glycerol, light oil,
Na
2
SO
3
, ZnCl
2

G
= V
s
__
62.4
L
__
72

_
1/3
Hughmark [97]
STR N=2440 rpm; V
s
=0.85 cm/s Pure liquids (water, cyclohexane,
n-hexane)

G
= 0.31
_
LVs

_
2/3
_
L
3
g
4
L
_
1/6
+0.45
(NNVD)Di

gDt
Di
Dt
Van Dierendonck et al. [98]
Bubble column V
s
=0.340 cm/s Aqueous: glycol, methanol, CCl
4
, NaCl,
Na
2
SO
3
G
(1G)
4
= C
_
gD
2
t
L

_
1/8
_
gD
3
t
v
2
L
_
1/12
_
Vs

gDt
_
Akita and Yoshida [99]
Bubble column V
s
<27.78 cm/s Aqueous: glycol, methanol, CCl
4
G
(1G)
1/4
= 0.2
_
gD
2
t
L

_
1/8
_
gD
3
t
v
2
L
_
1/12
_
Vs

gDt
_
Akita and Yoshida [68]
STR N=25420 rpm; V
s
=0.7615.74 cm/s Aqueous
G
=
_
Vs G
Vs +Vt
_
0.5
+2.16 10
4
_
(Pe/V)
0.4

0.2
L

0.6
_
_
Vs
Vs +Vt
_
0.5
Miller [100]
Aqueous: sucrose, methanol, n-butanol
G
= 0.505V
0.47
s
_
0.072

_
2/3
_
0.001
L
_
0.05
Hikita and Kikukawa [101]
STR N=315900; V
s
=0.722.16 cm/s Aqueous: glycerine
G
= 10
3
_
LVs Di
L

0.2
_
LN
2
D
3
i

_
0.4
Sterbacek and Sachova [102]
STR N=502100 rpm; Q
G
=0.0631.47 m
3
/s Aqueous: propionic acid, methyl acetate,
ethylene glycol, glycerol, Na
4
SO
3

G
N
1.04
V
0.57
s
Hassan and Robinson [103]

G
= C
_
QGN
2

STR N=250500 rpm; V


s
=0.6341.27 cm/s Aqueous
G
= 0.34(P/V)
0.25
V
0.75
s
de Figueiredo and Calderbank [70]
STR N=4801800 rpm; V
s
=15 cm/s Cyclohexane
G
=
_
Vt G
Vs
_
0.5
+0.000216
_
(P/V)
0.4

0.2
L

0.6
_
_
Vt
Vs
_
0.5
_
Et
P
__
g
a
_
Sridhar and Potter [85]
Bubble column V
s
=130 cm/s Aqueous: methanol, ethanol, n-propanol,
i-propanol, n-butanol

G
=
0.96V
0.58
s
C
0.26
N
(1+2.6VL)
Kelkar et al. [104]
Bubble column V
s
=0.9815.6 cm/s Aqueous: glycerol, glycol, BaCl
2
,
Na
2
SO
4
G
(1G)
4
=
0.16(UGL/)
0.964
(L
3
/g
4
L
)
0.283
(Dd/Dt )
0.222
(N/Dt )
0.0237 exp(0.00185(Crk
2
/))
11.61[1exp(0.00565(Crk
2
/))]
Koide et al. [105]
Bubble column V
s
=315 cm/s Aqueous: glycerol, glycol, BaCl
2
,
Na
2
SO
4
G
(1G)
4
=
0.277(UGL/)
0.918
(g
4
L
/L
3
)
0.252
1+4.35(Cs /S)
0.748
((SL)/L)
0.881
(Dt UGL/L)
0.168
Koide et al. [106]
Bubble column V
s
=315 cm/s Aqueous: glycerol, glycol, BaCl
2
,
Na
2
SO
4

G
=
GH,o[1+M exp (ln(UG/UGC))
2
]
1+49.1(Cs /S)
0.619
((SL)/L)
1.14
(Dt UGL/L)
0.496
STR N=400850 rpm; Q
G
=0.120.48 m
3
/h Aqueous: NaCl
G
(P/V)
0.45
V
0.61
s
Ho et al. [76]
Bubble column V
s
=27 cm/s Aqueous: CMC, Carbopol
G
V
0.7
s
Kawase et al. [77]
Bubble column V
s
=8 cm/s Aqueous: Sucrose, Ethanol-Ca alginate
G
= 2.1 10
3
_
Vs

gDt
_
0.82
_
gD
2
t
L

_
0.8
_
gD
3
t
v
2
L
_
0.1
Sun et al. [107]
Prediction model
G
1G
= 2
(3n+5)/(n+1)
n
(n+2)/2(n+1)
_
k
L
_
1/2(n+1)
g
n/2(n+1)
V
(n+2)/2(n+1)
s
Kawase et al. [108]
STR V
s
=0.10.7 vvm Aqueous: Glycerol
G
(P/V)
0.375
V
0.62
s
Nocentini et al. [109]
STR N=400750 rpm; Q=0.290.975 vvm Aqueous: CMC, sucrose, sodium
sulphate, sodium citrate, YM broth,
silicon antifoam

G
(P/V)
0.4
V
0.5
s
Arjunwadkar et al. [110]
STR N=250850 rpm; V
s
=0.2120.848 cm/s Aqueous
G
(P/V)
0.3091
V
0.7347
s
Linek et al. [111]
Aqueous: Na
2
SO
4

G
(P/V)
0.4903
V
0.5788
s
Airlift reactor N=0300 rpm; V
s
=010 cm/s Aqueous
G
(P/V)
0.85
Rostami et al. [112]
Airlift reactor V
s
=06 cm/s Aqueous: diesel, kerosine
G
= 1.320V
0.927
s
(1 +
WO
)
0.673
Mehrnia et al. [113]
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 417
uid recirculation. High gas throughputs encourage coalescence
leading to large bubbles with short residence times. Awide bub-
ble size distribution is attained. The effect of V
s
in this regime
on hold up is less pronounced than in the bubbly ow regime,
with the exponent varying between 0.2 and 0.7. Koide et al.
[105] noted that moving fromthe transition regime to the churn-
turbulent regime by increasing the supercial velocity increases
the overall gas hold up.
3.1.2. Inuence of turbulence on oxygen transfer
coefcient
In addition to its effect on the interfacial transfer area, inten-
sied uid turbulence impacts on K
L
by decreasing the width of
the stagnant boundary layer, according to the two-phase theory.
This results in a decreased resistance to molecular diffusion and
a consequent increase in K
L
and increased overall rate of oxygen
transfer. This effect has been noted in a bubble column where
an enhancement of K
L
a on increased aeration was concluded to
result fromincreased disturbance of the concentration boundary
layer by neighbouring bubbles [122]. Thus an increase in turbu-
lence will increase K
L
a macroscopically through the increases in
the interfacial transfer area available for oxygen transfer, as well
and microscopically by increasing the effectiveness of transfer
across these interfaces.
Enhanced K
L
with increased turbulence has been empirically
demonstrated through the effect of P/V and N on K
L
and a num-
ber of correlations are available (Table 5). K
L
proportional to
(P/V)
0.090.531
has been observed in a variety of systems con-
taining hydrocarbons, similar to systems containing electrolytes
or polymers [28,66,70,82,123,124,125,126,127]. K
L
is similarly
enhanced with increasing turbulence through increased N as
observed in other several publications [66,125,128].
K
L
has been found to be only slightly dependent on V
s
over
a range of tank geometries [100] and for both aqueous and
aqueous-solid suspensions [134]. These results have been con-
rmed by Prasher and Wills [126], Schumpe and Deckwer [74]
and Sun and Furusaki [87]. Schumpe and Deckwer [74] found
K
L
V
0.8
s
with K
L
constant at V
s
over 8 cm/s. Hassan and
Robinson [71,72] and Ho et al. [76] found K
L
to be indepen-
dent of V
s
for all practical purposes. Similarly, K
L
was found to
be independent of V
s
in several other studies [28,70,81,123,130].
The lesser effect of V
s
on K
L
suggests that the inuence of V
s
on coalescence is diluted by its effect on K
L
through increased
turbulence.
In addition to enhancing K
L
through reduced resistance to
oxygen transfer, turbulence also alters K
L
through its inu-
ence on bubble size. K
L
associated with large (>2.5 mm)
bubbles was greater than K
L
associated with small (<2.5 mm)
bubbles by as much as vefold [132]. A transition zone
between the large and small bubble regimes has been observed
where K
L
decreased uniformly with decreasing bubble size
[28,67,100,124,131,135,136,137,138,139]. In this transition
zone, K
L
has been found to be proportional to D
1.04
32
[28] and
D
0.813
32
[69] in non-viscous electrolyte solutions.
The dependence of K
L
on bubble size has been attributed
to the different characteristics in the hydrodynamic regimes
of the rigid spherical surfaces of the small bubbles and the
deformable, mobile surfaces of the large bubbles [27,69,70].
Bubbles with mobile oscillating surfaces have an internal gas
circulation that promotes oxygen transfer by regenerating the
gasliquid surface at the interface. Bubbles with rigid surfaces,
on the other hand, have no internal circulation, but behave
as rigid spheres, resulting in relatively lower oxygen trans-
fer. Further, oscillating bubbles are able to wobble and move
in spirals during free rise, movements which generate con-
vective motions that greatly enhance the oxygen transfer rate
[139], with rates from oscillating bubbles 770 times larger
than those from rigid bubbles [140]. The enhanced transfer
rates are suggested to result from a marked effect of surface
rigidity on K
L
for which a value of 4.56 10
2
cm/s is gener-
ally accepted for oscillating bubbles [28] while K
L
values for
rigid bubbles are 3.6-fold lower at 1.27 10
2
cm/s [28,135].
Robinson and Wilke [69] and Akita and Yoshida [68] also
observe a decrease in K
L
with decreasing bubble size corre-
sponding to the transitional zone between oscillating and rigid
bubbles.
Calderbank [27], Calderbank and Moo-Young [28], Midoux
and Charpentier [141] and Schmitz et al. [78] have shown that
while K
L
decreased in the transition regime from oscillating
to rigid bubbles, K
L
remained constant with changing bubble
diameter for either large oscillating bubbles or small rigid bub-
bles. This suggests that in discrete bubble regimes of rigid and
oscillating bubbles, turbulence has no effect on K
L
. Hassan
and Robinson [71,72] found K
L
to be independent of P/V for
P/V<2000 W/m
3
corresponding to the rigid regime. However,
for P/V>2000 W/m
3
, K
L
decreased with increasing P/V in the
transitional zone, being proportional to (P/V)
0.56
. Ho et al.
[76] likewise found no dependence of K
L
on P/V in the rigid
regime.
These authors developed semi-theoretical equations for rigid
and oscillating bubbles based on turbulence having no effect on
K
L
in these regimes. For bubbles with a rigid interface, K
L
is
obtained by the following equation [132,137]:
K
L

V
sl
d
b
D
2/3
G
v
1/6
L
(5)
For bubbles with a mobile interface, K
L
is obtained by the
following equation [132]:
K
L

V
sl
d
b
D
1/3
G
(6)
A constant K
L
in each of the rigid and oscillating regimes is
not, however, supported by Linek and Mayrhoferov a [138] and
Linek et al. [124], who found that K
L
decreased with increasing
bubble size outside the transition zone.
3.2. Inuence of uid properties on interfacial area and
oxygen transfer coefcient
The uid properties inuence the characteristics of K
L
a
through their inuence on the interfacial transfer area or on K
L
.
Hydrocarbons may enhance K
L
a by increasing the gasliquid
4
1
8
K
.
G
.
C
l
a
r
k
e
,
L
.
D
.
C
.
C
o
r
r
e
i
a
/
B
i
o
c
h
e
m
i
c
a
l
E
n
g
i
n
e
e
r
i
n
g
J
o
u
r
n
a
l
3
9
(
2
0
0
8
)
4
0
5

4
2
9
Table 5
Empirical correlations for the prediction mass transfer coefcient
System Operating conditions Liquid phase(s) Equation Reference
K
L
= 1.13
_
V
sl
d
b
D
0.5
G
Bird, Stewart & Lightfoot [129]
Sieve plates V
s
=6.160.1 cm/s Aqueous: aliphatic alcohols, glycol
K
L
D
32
D
L
=
_

L

L
D
L
_
0.5
_
D
32
V
sl

L
_
0.766
Calderbank [27]
Review K
L
_
v
L
D
L
_
1/2
= 0.42
_
(
G

L
)
L
g

2
L
_
1/3
Calderbank and Moo-Young [28]
K
L
_
v
L
D
L
_
2/3
= 0.31
_
(
G

L
)
L
g

2
L
_
1/3
K
L
_
v
L
D
L
_
2/3
= 0.13
_
(P/V)
L

2
L
_
1/4
STR N=60400 rpm;
V
s
=0.16930.762 cm/s
Aqueous: NaOH K
L
N
0.6
(P/V)
0.2
D
0.4
32
Yoshida and Miura [66]
Bubble columns V
s
=445 cm/s Aqueous: kerosene, glycerol, light
oil, Na
2
SO
3
, ZnCl
2
K
L
D
32
D
L
= 2 +0.0187
_
_
d
b
V
sl

L
_
0.484
_
v
L
D
L
_
0.339
_
d
b
g
1/3
D
L
__
1.61
Hughmark [97]
Prediction models K
L
= 0.4
_
(P/V)v
L

L
_
1/4
_
D
L
v
L
_
1/2
Lamont and Scott [123]
STR N=250550 rpm Aqueous: Na
2
SO
3
, KI, Na
2
SO
4
K
L
(P/V)
0.19
Linek et al. [124]
STR N=255, 325 and 455 rpm;
Q
G
=0.68 m
3
/h
Aqueous: NaCl, sodium dodecyl
sulphate, Dow Corning antifoam
K
L
D
0.5
32
Benedek and Heideger [67]
STR N=180780 Aqueous: NaCl K
L
(P/V)
0.3
Koetsier and Thoenes [125]
STR N=180780 Aqueous: NaCl K
L
= 3.25 10
4
_
N
3
D
5
i
D
3
t
_
0.3
_
D
i
Dt
_
0.7
_
1
Dt
_
0.35
Koetsier and Thoenes [125]
STR N=150350 rpm;
V
s
=0.291.2 cm/s
Aqueous: NaOH K
L
= 0.592
_
(P/V)v
L

L
_
1/4
_
D
L
v
L
_
1/2
Prasher and Willis [126]
Bubble column Aqueous: glycol, methanol, CCl
4
K
L
D
32
D
L
= 0.5
_
v
L
D
L
_
0.5
_
gD
3
32
v
2
L
_
0.25
_
gD
2
32

_
0.375
Akita and Yoshida [68]
STR N=100500 rpm;
V
s
=0.1620.466 cm/s
Aqueous: carbopol K
L
= 5.11 10
3
_
11
n1
_
3n+1
4n
_
n

0.426
D
0.5
(k/
L
)
0.426
N
1.3520.426n
D
0.852
i
Perez and Sandall [128]
STR N=250500 rpm;
V
s
=0.6341.27 cm/s
Aqueous K
L
(P/V)
0.33
de Figueiredo and Calderbank [70]
STR N=7321900 rpm;
V
s
=0.371.11 cm/s
Aqueous: Na
2
SO
4
, KOH, KCl,
K
2
SO
4
K
L
(P/V)
0.56
Hassan and Robinson [71]
STR N=7321900 rpm;
V
s
=0.371.11 cm/s
Aqueous: Na
2
SO
4
, KOH, KCl,
K
2
SO
4
K
L
(P/V)
0.56
Hassan and Robinson [72]
Review
K
L

L
=
_
LV
L
v
L
_

_
v
L
D
L
_

Moo-Young and Blanch [25]


Bubble column V
s
=020 cm/s Aqueous: CMC K
L
= 4.5 10
3
V
0.08
s

0.32
L
Schumpe and Deckwer [74]
Bubble column V
s
=27 cm/s Aqueous: CMC, carbopol K
L
=
2

D
L
_
Vs
L
g
k
_
1/2(1+n)
Kawase et al. [77]
Bubble column V
s
=8 cm/s Aqueous: sucrose, ethanol-Ca
alginate
K
L
D
32
D
L
= 0.13
_

L
D
L

L
_
1/2
_
gD
2
32

_
0.1
_
gD
3
32

2
L

2
L
_
0.43
_
dp
Dt
_
0.2
Sun and Furusaki [87]
Bubble column V
s
=0.87 cm/s Aqueous: CMC-C. cellulolyticum K
L
= 0.301
_
v
L

L
_
1/4
_
D
L
v
L
_
1/2
Kawase et al. [130]
K
L
=
_
4

_
2
1/n
_
1
T
+
_
D
1/2
L
_
k

L
_
1/2(1+n)

1/2(1+n)
Theoretical models K
L
=
0.42
_
9(n+1)
2(2n+1)
__
30A(n+1)
(2n+1)
_
1/3
D
2/3
L
_
k

L
_
1/(n+2)
g
(4n)/3(n+2)

2(n1)
2
/9(n+1)(n+2)
G
Kawase and Moo-Young [131]
K
L
= 0.42
_
2

_
D
1/2
L
_
k

L
_
1/2(n+2)
g
1/(n+2)

(n1)/3(n+2)
G
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 419
T
a
b
l
e
5
(
C
o
n
t
i
n
u
e
d
)
S
y
s
t
e
m
O
p
e
r
a
t
i
n
g
c
o
n
d
i
t
i
o
n
s
L
i
q
u
i
d
p
h
a
s
e
(
s
)
E
q
u
a
t
i
o
n
R
e
f
e
r
e
n
c
e
S
T
R
N
=
2
0
0

6
0
0
r
p
m
A
q
u
e
o
u
s
:
i
-
p
r
o
p
a
n
o
l
,
a
m
y
l
a
l
c
o
h
o
l
K
L
=
0
.
1
3
2
_
g

L
_
1
/
3
_
D
G

L
_
1
/
2
_
P V
_
0
.
1
5
P
a
n
j
a
a
n
d
P
h
a
n
e
s
w
a
r
a
R
a
o
[
1
2
7
]
S
T
R
N
=
2
5
0

6
0
0
r
p
m
;
Q
G
=
0
.
6
a
n
d
1
.
2
m
3
/
h
A
q
u
e
o
u
s
:
N
a
2
S
O
4
,
P
E
G
K
L
=
0
.
6
_
V
s
l
d
b
D
2
/
3
G
v
1
/
6
L
A
l
v
e
s
e
t
a
l
.
[
1
3
2
]
S
T
R
N
=
2
5
0

6
0
0
r
p
m
;
Q
G
=
0
.
6
a
n
d
1
.
2
m
3
/
h
A
q
u
e
o
u
s
:
N
a
2
S
O
4
,
P
E
G
K
L
=
1
.
1
3
_
V
s
l
d
b
D
0
.
5
G
A
l
v
e
s
e
t
a
l
.
[
1
3
2
]
S
T
R
N
=
2
5
0

8
5
0
r
p
m
;
V
s
=
0
.
2
1
2

0
.
8
4
8
c
m
/
s
A
q
u
e
o
u
s
:
N
a
2
S
O
4
K
L
=
0
.
5
2
3
_
(
P
/
V
)
v
L

L
_
1
/
4
_
D
L
v
L
_
1
/
2
L
i
n
e
k
e
t
a
l
.
[
1
1
1
]
T
h
e
o
r
e
t
i
c
a
l
m
o
d
e
l
s
A
q
u
e
o
u
s
:
g
l
u
c
o
s
e
,
x
a
n
t
h
a
n
g
u
m
,
s
i
l
i
c
o
n
e
a
n
t
i
f
o
a
m
;
b
i
o
s
u
r
f
a
c
t
a
n
t
-
X
a
n
t
h
o
m
o
n
a
s
c
a
m
p
e
s
t
r
i
s
N
R
R
L
B
-
1
4
5
9
,
C
a
n
d
i
d
a
b
o
m
b
i
c
o
l
a
N
R
R
L
Y
-
1
7
0
6
9
K
L
=
2

D
L
_

L
k
_
1
/
2
(
1
+
n
)
G
a
r
c
i
a
-
O
c
h
o
a
a
n
d
G
o
m
e
z
[
1
3
3
]
K
L
=
2

D
L
_

L
(
1

r
)
2

L
_
0
.
2
5
S
T
R
N
=
0

1
1
3
0
r
p
m
;
V
s
=
1
8

5
4
c
m
/
s
A
q
u
e
o
u
s
:
N
a
2
S
O
3
K
L
=
0
.
4
4
8
_
(
P
/
V
)
v
L

L
_
0
.
2
5
_
D
L
v
L
_
0
.
5
L
i
n
e
k
e
t
a
l
.
[
8
2
]
A
q
u
e
o
u
s
:
N
a
2
S
O
3
K
L

(
P
/
V
)
0
.
2
3
8
A
q
u
e
o
u
s
:
N
a
2
S
O
3
,
S
o
k
r
a
t
K
L

(
P
/
V
)
0
.
2
4
3
A
q
u
e
o
u
s
:
N
a
2
S
O
3
,
C
M
C
T
S
.
5
K
L

(
P
/
V
)
0
.
2
7
6
A
q
u
e
o
u
s
:
N
a
2
S
O
3
,
C
M
C
T
S
.
2
0
K
L

(
P
/
V
)
0
.
0
9
A
q
u
e
o
u
s
:
N
a
2
S
O
3
,
o
c
e
n
o
l
K
L

(
P
/
V
)
0
.
1
8
8
A
q
u
e
o
u
s
:
N
a
2
S
O
3
,
P
E
G
1
0
0
K
L

(
P
/
V
)
0
.
2
4
6
interfacial area. Alternatively, hydrocarbon may decrease K
L
a
by increasing the resistance to oxygen transfer.
3.2.1. Inuence of uid properties on interfacial area
The interfacial area is highly dependent on the liquid phase
physicochemical properties, such as surface tension and vis-
cosity, which dene bubble size through their inuences on
bubble break up and coalescence. Hydrocarbons change these
uid properties, and hence the bubble size, gas hold up and inter-
facial area available for transfer. The type of hydrocarbon plays
a major role in the degree to which it affects the uid properties
of the dispersion [142].
Hydrocarbons have been demonstrated to decrease the
surface tension. Bi et al. [60] reported surface tensions of
71.4 mN/m, 70.2 mN/m and 67.7 mN/m with ethanol con-
centrations of 1 g/L, 3 g/L and 8 g/L, respectively. Likewise,
surface tension was lowered from 73 mN/m to 66 mN/m when
polypropylene glycol methyl ether (PGME) concentration was
increased from 0 ppm to 100 ppm [79] and by almost two thirds
on addition of cyclohexane to an airaqueous system [117].
The surface active effect of the hydrocarbon is a con-
sequence of the coexistence of a hydrophilic functional
group (e.g. OH; CO; COOH) and a hydrophobic struc-
tural group. The hydrocarbon molecules orientate at the
gasliquid interface with the hydrophobic group directed
towards the gas phase and the hydrophilic group towards
the liquid phase. The positioning of the hydrophilic group
around the surface of the gas bubbles results in surface
polarization and repulsive forces which inhibit coalescence
when the bubbles come into close contact [127,143]. Sur-
face charges have similarly been implicated in coalescence
inhibition by hydrocarbons [92,143]. Other mechanisms of coa-
lescence inhibition proposed include spatial blocking by the
oil drops between the bubbles, restricting bubble movement
and bubblebubble contact [36,45,61,144] and a decrease in
the force available to rupture the lm for coalescence to occur
[117,145].
Increased hydrocarbon concentration has been shown exper-
imentally to result in enhancement of interfacial area. Dodecane
addition up to 25%(v/v) increased the interfacial area by approx-
imately 15% [35]. Zieminski et al. [146] found increases in
interfacial area of 1100% for alcohols, 1100% for monocar-
boxylic acids and 500% for dicarboxylic acids upon addition
of additives of 90 ppm, 500 ppm and 50 ppm, respectively. Al
Taweel and Cheng [79] obtained a 45-fold increase in interfa-
cial area when increasing the additive concentrations of PGME
(>20 ppm).
The interfacial area relates inversely to the bubble size
through the gas hold up, according to Eq. (4). The bub-
ble size is determined by the balance of surface tension
and viscous forces [29] and, as hydrocarbon concentration
increases, bubbles become smaller with decreasing surface ten-
sion [50,68,79,117,127,146,147]. Thus the surface tension has
been related to the interfacial area through its effect on bubble
size.
The link between surface tension and bubble size has been
established in a study using diethylene glycol and decanol
420 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
whose uid properties differ signicantly only in their surface
tension [96]. Here D
32
was considerably smaller in decanol
which had a correspondingly lower surface tension. The cor-
relation of decreased surface tension and reduced bubble size
is further supported by an observed decrease in D
32
due to
addition of ethanol [60,117,148], propanol [96], i-propanol
[127], n-hexanol [54], 2-ethyl hexanol [142], n-heptanol [54],
n-octanol [54], amyl alcohol [127], n-diols [149,150], alkanes
[35], ketones [149,150], castor oil [144], silicone oil [50,55,64],
toluene [142] and monocarboxylic acids [146,149,150].
The magnitude of the effect of reduced surface tension on
bubble size is marked. Keitel and Onken [149,150] reported that
addition of a variety of hydrocarbons, including n-alcohols (n-
C
18
), n-diols (n-C
25
), ketones (C
34
, C
7
) and monocarboxylic
acids (C
13
), decreased the average diameter of 4.12.5 mm.
Zieminski et al. [146] and Zieminski and Whittemore [151]
added mono- and dicarboxylic acids and observed a decrease
in average bubble diameter of 300% with addition of 100 ppm
organic solutes. Galindo et al. [144] found that in a simulated
fermentation broth containing castor oil, bubble size decreased
from 2 mm to 0.8 mm as the oil content was increased to 15%
(v/v). Calderbank and Moo-Young [28] found the mean bubble
size reduced with an increase in oil concentration up to 15%in a
salt rich aqueous solution. Rols et al. [35], using a photographic
method, reported a 15% reduction in D
32
when n-dodecane was
added up to 25% (v/v).
The reduction of surface tension and inhibition of bub-
ble coalescence by hydrocarbons promotes gas dispersions
with slower rising velocities and, thus, an increased gas
hold up [89,152]. Increased gas hold up was reported with
ethanol addition [60,107,148], butanol [153] and oleic acid
concentrations [58]. Addition of alcohols methanol, ethanol,
butanol, hexanol and octanol to water also increased the
gas hold up by 70100% [154]. i-Propanol at concentra-
tions ranging from 0.01 to 1% (v/v) increased the gas hold
up with increasing alcohol concentration [155]. Al Taweel
and Cheng [79] found gas hold up to increase up to ve-
fold with increasing concentration of PGME. Chaumat et al.
[120] observed 3075% greater gas hold up in cyclohexane
compared with water. The degree to which the hydrocarbon
affected gas hold up was dependent on its chain length. Long
chain alcohols affected gas hold up considerably more than
short chain alcohols. The gas hold up increased with chain
length in the following order: water >methanol >ethanol >i-
propanol >n-propanol >n-butanol [104].
In addition to their inuence on surface tension, hydrocar-
bons also signicantly inuence the viscosity of the dispersion,
but in an opposite manner. As the hydrocarbon concentration is
increased, so is the viscosity, with a correspondingly decreased
interfacial area. Viscosity is proposed to decrease interfacial area
according to one or more of several mechanisms inuencing
break up, coalescence or formation of the bubbles. Firstly, an
increased viscosity decreases the turbulence in the liquid. This
results in a reduction in the energy of the eddies, with concomi-
tant damped bubble break up [117]. The damped bubble break
up is likely to be exacerbated by an increased boundary layer
thickness around the bubbles which would require eddies with
high energy for penetration and break up to occur [156]. Sec-
ondly, increased viscosity enhances bubble coalescence [155].
Thirdly, increased viscosity results in a slower formation of the
liquid lm at the point of bubble formation thereby trapping a
larger amount of air in each bubble, and thus larger but fewer
bubbles [88].
The negative inuence of viscosity on interfacial area has
been demonstrated experimentally through its effect on bub-
ble size. In dispersions containing propanol and decanol, where
decanol is more viscous than propanol but has a comparable sur-
face tension, D
32
was signicantly larger in the decanol than in
the propanol dispersion [96]. Similar results were demonstrated
by these authors for viscous alcohols where the bubbles in the
more viscous dispersions were relatively larger. Several other
studies have conrmed an increase in average bubble size with
increasing viscosity [88,113,115,117,144,157,158].
The interfacial area is further decreased due to a decreased
gas hold up in viscous uids, as a result of the formation of larger
bubbles under these conditions. The larger bubbles pass more
quickly through the reactor with high bubble rise velocities and
low residence times [113,156,159,160,161]. As much as a 30%
decrease in gas hold up in viscous liquids has been attributed to
these large bubbles.
Since larger bubbles have a shorter residence time relative to
smaller bubbles, increased viscosity may result in an increased
proportion of smaller bubbles in the population. Apositive effect
of viscosity on gas hold up due to an increased proportion of
smaller bubbles in the population [162] is suggested to balance
the negative effect of viscosity on gas hold up as a result of
the generation of larger bubbles [113]. The behaviour of rising
bubbles inevitably becomes restricted with increasing viscos-
ity, however, due to the increase of drag force on the bubbles
[155,163], even in the case of large bubbles.
The resultant effect of increased hydrocarbon fraction on
the interfacial area and gas hold up of the dispersion will ulti-
mately depend on the relative predominance of the inuences of
decreased surface tension and increased viscosity in the system
under consideration. Studies which report a minimum constant
bubble size, irrespective of an increase of hydrocarbon concen-
tration, suggest that under the conditions of these studies, the
negative effect of increased viscosity on interfacial area with
hydrocarbon addition was compensated by a coincident posi-
tive effect of reduced surface tension. A minimum bubble size
was observed at 34% octane in an octaneaqueous suspension
[164] and at ethanol concentrations above 1 g/L [60]. Keitel and
Onken [149,150] related specic critical concentrations to a con-
stant bubble size for each of a wide variety of additives. Other
authors who have found similar observations of critical con-
centrations include Albal et al. [116] and Machon et al. [92].
Yamamoto et al. [165] found the bubble size to be almost inde-
pendent of volume fraction of PFC-43 less that 60% (v/v). This
trend is also evident in the study of PFC-40 addition to 15%
(v/v), in which no effect on the interfacial area was observed
[65].
A cancellation of these two effects is supported by the obser-
vation of a constant maximum gas hold up with addition of
glycerol [160], addition of PGME [79], peruorotributylamine
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 421
[165], 2-ethyl hexanol [142] and toluene [142]. A maximum
gas hold up was similarly observed at a liquid viscosity of 4 cP
[155]. However, Das et al. [142] also noted that at concentra-
tions above 10%in both 2-ethyl hexanol and toluene, increasing
hydrocarbon concentration resulted in bubble size increase. This
suggests that over a critical hydrocarbonconcentration, the effect
of increased viscosity may outweigh that of decreased surface
tension.
The empirical models which dene the inuence of P/V,
N and V
s
on interfacial area (Table 2) also incorporate the
inuence of viscosity and surface tension on this parameter.
Calderbank [26] rst described the inuence of the surface
tension on the interfacial area by assessing the effect of 10
different aqueoushydrocarbon emulsions at the 5 L and 100 L
scale. This model and its extensions by Sridhar and Potter
[75] and Kawase et al. [77] predict an increase in interfacial
area with a decrease in surface tension according to a
0.6
.
Similar dependencies of interfacial area on surface tension,
namely a
0.5
and a
0.33
, were formulated by Akita
and Yoshida [68] and Calderbank [27], respectively. Al Taweel
and Cheng [79] observed a substantially higher dependency of
interfacial area on surface tension with PGME, according to
a
15.4
.
The inuence of viscosity on interfacial area was incorpo-
rated by Kawase et al. [77]. Examination of CMC solutions
showed the interfacial area to be inversely proportional to con-
tinuous phase viscosity (
c
) according to (
d
/
c
)
0.25
. Negative
inuence of viscosity on the interfacial area was also noted in by
Calderbank [27], Schmitz et al. [78], Garcia-Ochoa and Gomez
[81] and Schumpe and Deckwer [74] who reported a
0.125
L
,
a
0.18
L
, a
0.3
L
and a
0.51
L
, respectively. Akita and
Yoshida [68] found a similar effect on kinematic viscosity on
interfacial area in a bubble column, namely a v
0.2
L
.
The inuence of surface tension and viscosity on bubble size
and gas hold up is similarly incorporated into these equations
(Tables 3 and 4) such that a decreased surface tension is related
to a smaller bubble size and increased gas hold up. Increased
viscosity may be related to increased or decreased bubble size
and generally a decreased gas hold up, depending on the system.
3.2.2. Inuence of uid properties on oxygen transfer
coefcient
The inuence of hydrocarbon on the oxygen transfer coef-
cient (K
L
) has generally been found to be negative. Further, the
degree to which the hydrocarbon decreases K
L
depends on the
hydrocarbon type. Koide et al. [54] reported a 44%, 50% and
62.5% decrease in K
L
upon addition of n-hexanol, n-heptanol
and n-octanol, respectively, suggesting a more marked effect of
alcohols with a long chain length when compared with those
of a comparatively shorter chain length. Similar results have
been found by Raymond and Zieminski [135] who report a
K
L
decrease of 53.8% for n-hexanol and 700% for n-octanol,
attributing the larger drop in K
L
to the increased chain length of
the hydrocarbon.
A major mechanism through which the hydrocarbon inu-
ences K
L
is through its effect on the bubble mobility. Calderbank
and Moo-Young [28] ascribed a steady decrease in K
L
with
increasing oil concentration from 14 to 69.5% (v/v) to the
adsorption of the hydrocarbon at the gasliquid interface and
reduced surface tension. A reduction in surface tension results
in a decrease in the bubble surface mobility [135] and the bub-
ble behaves as a rigid sphere. Consequently internal motion is
retarded and K
L
is reduced [26,166].
The link between reduced surface tension and decreased K
L
is supported experimentally. Linek et al. [82] added substances
that inhibit surface tension (ocenol, polyethylene glycol (PEG),
CMC and TS.20) and observed a reduction in K
L
of up to 80%.
Yagi and Yoshida [167] also observed a decreased K
L
on the
addition of 0.1%of surfactant. Further, a decrease in K
L
has been
observed on the addition of antifoam agents such as silicone oil
and soybean oil [50].
In addition to the surfactant effect of the hydrocarbon, the
increased viscosity also impacts negatively on K
L
. Several stud-
ies suggest that the comparatively more viscous nature of the
hydrocarbon increases the liquid phases resistance to transfer
[82,133,156,168,169,170]. This may be due to the decrease in
internal recirculation of bubbles or depression of the diffusivity
of dissolved gas in the liquid.
The oxygen transfer coefcient has been shown to be highly
dependent on the liquid-phase diffusion coefcient. This rela-
tionship has been found for a wide variety of process conditions
and additives for both rigid (Eq. (7)) and oscillating bubbles (Eq.
(8)).
K
L
= D
2/3
L
(7)
K
L
= D
1/2
L
(8)
Eq. (7) was found applicable for rigid bubbles in
both single and mixed electrolyte solutions [76,132], and
hydrocarbonaqueous [28] and CMC-aqueous suspensions
[131]. Similarly, Eq. (8) for oscillating bubbles has been
found true for hydrocarbons [28,68,87,127,133], electrolyte
[27,82,111,132] and aqueous solutions [77,123,126,130,131].
4. Prediction of the overall volumetric oxygen transfer
coefcient and oxygen transfer rate in
hydrocarbonaqueous dispersions
Prediction of K
L
a in hydrocarbonaqueous dispersions is
complicated. In addition to the uid turbulence, the physic-
ochemical properties of the hydrocarbonaqueous dispersion,
dened mostly by the hydrocarbon concentration and its molec-
ular structure and chain length, strongly inuence the behaviour
of K
L
a. The difculty of dening a predictive model for K
L
a
is exacerbated by the complex interactions between the hydro-
carbon and the operating variables, particularly agitation rate,
which may result in different K
L
a trends over similar hydrocar-
bon concentration ranges and comparable hydrocarbon types.
K
L
a prediction in hydrocarbonaqueous dispersions has
been restricted almost exclusively to empirical correlations
(Tables 6 and 7). K
L
a in these systems, in common with K
L
a
behaviour in aqueous systems, exhibits a dependence on the tur-
bulence generated by agitation and aeration, the uid properties
422 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
Table 6
Empirical correlations for the prediction of overall volumetric oxygen transfer coefcient
System Liquid phase(s) Equation Reference
STR; N=240570; V
s
=0.20.8 cm/s Aqueous n-C
1021
solution (K
L
a)
GW
= (1.75 10
2
e
0.115
0.8
10
3
e
46.9
)NV
1/3
s
Matsumura et al.
[171]
STR; N=240570; V
s
=0.20.8 cm/s Aqueous n-C
1021
solution (K
L
a)
OW
=
1
1+H
WO
(1/(1))
(K
L
a)
GW
Matsumura et al.
[171]
STR; two 6-blade turbines; D
t
=17.6 cm;
V
L
=4 L; N=600 rpm; Q=0.5 vvm
Xanthan solution/vegetable oil (K
L
a)
GO
H (P/V)
0.36
V
0.59
s

W
Xn
0.26
Zhao et al. [172]
STR; two 6-blade Rushton turbines; four
bafes; D
t
=11 cm; V=1 L;
N=200800 rpm; Q=0.52 vvm
n-Hexadecane/nutrient medium K
L
a = 650(P/V)
0.31
V
0.7
s
(1 )
1.7
Nielsen et al. [63]
STR; two 4-blade turbines; three bafes;
D
t
=17.5 cm; V=4 L; N=0700 rpm;
V
s
=00.5 cm/s
n-Dodecane/simulated broths
(CMC/sodium salt solution)
K
L
a = 5.6 10
3
_
(P/V)
1.54
V
2.30
s

3.78
a
_

Galaction et al. [38],


Cascaval et al. [39]
STR; two 4-blade turbines; three bafes;
D
t
=17.5 cm; V=4 L; N=0700 rpm
V
s
=00.5 cm/s
n-Dodecane/Propionibacterium
shermanii broths
K
L
a = 0.714
_
V
0.97
s
(P/V)
0.40
C
1.22
x
_

Cascaval et al. [39]


n-Dodecane/S. cerevisiae broths K
L
a = 6.72 10
2
_
(P/V)
2.74
V
0.75
s
C
0.03
x
_

N=3251200 rpm CO
2
and propene in
toluene/water; acetylene in
sulphur(s)/water
(K
L
a)

(K
L
a)
=0
=
_
(D
L
)

(D
L
)
=0
_
1/2
_
(
L
)
=0
(
L
)

( )

(
L
)

(
L
)
=0
( )
=0
_
1/4
Zhang et al. [173]
of the aqueous phase and the geometric constraints. Accord-
ingly, the conventional generalised empirical correlations for
non-viscous aqueous solutions (Eqs. (2) and (3)), which relate
K
L
a to supercial gas velocity and power per unit volume (or
agitation rate), have frequently been used as a basis for the devel-
opment of the more complex empirical correlations proposed
for hydrocarbonaqueous dispersions by additionally account-
ing for the inuences of the physical properties of a second
immiscible phase. These adaptations usually comprise either
incorporation of a term or terms relating to the hydrocarbon
fraction or the physical properties of the dispersion (Table 6).
An empirical correlation to predict the observed behaviour
of K
L
a at discrete n-hexadecane concentrations up to 33% in a
hydrocarbonaqueous dispersion incorporated a term contain-
ing the alkane fraction [63]. In this study, K
L
a was maximal
at the lowest alkane concentration and decreased consecutively
at each subsequent concentration examined. Consequently,
this correlation is limited to use where the inuence of the
hydrocarbon is negative over the entire range of hydrocarbon
concentrations.
Zhao et al. [172] similarly developed an empirical correlation
which described the behaviour of K
L
a in a vegetable oilaqueous
dispersion where the properties of the aqueous phase was var-
ied, here by altering the concentration of xanthan gum in the
aqueous phase. In this study, though, both the inuence of the
change in the properties of the aqueous phase as well as the
change in oil fraction was assessed. The predicted decrease in
K
L
a with increased xanthan gum concentration and increased
oil fraction, described their data. This correlation will, as with
that proposed by Nielsen et al. [63], be limited to use where the
inuence of the hydrocarbon is negative over the entire range of
hydrocarbon concentrations. Further, since this correlation was
developed from data obtained in a system where oil comprised
the continuous phase, its validity in hydrocarbonaqueous dis-
persions, where the hydrocarbon is the dispersed phase, remains
questionable.
Empirical correlations which predict an increased K
L
a on
hydrocarbon addition have also been developed in terms of the
ratio of K
L
a in the presence of hydrocarbon relative to that
without hydrocarbon [173]. Although these authors validated
their model with data from CO
2
and propene absorption in
a toluenewater dispersion, and absorption of acetylene in a
sulphurwater suspension, Dumont et al. [64] have suggested it
is applicability to enhanced oxygen absorption in the presence
of hydrocarbon droplets.
These correlations have all assumed that oxygen transfer in
a hydrocarbonaqueous dispersion can be represented by a sin-
gle rate-limiting K
L
a. In an aqueous solution, where the oxygen
transport takes place directly from the gas to the aqueous phase
and the major resistance to oxygen transfer is considered to
reside in the stagnant aqueous layer at the gasliquid interface,
this is generally held to be reasonable. However, in a system
containing gas bubbles and hydrocarbon droplets, oxygen trans-
port from the gas to the aqueous phase may take place by one of
several mechanisms, and the assumption of a single rate-limiting
K
L
a may not be valid.
An important consideration in ascertaining the most likely
oxygen transport mechanism is the adsorption of hydrocarbon
droplets to the bubble surface, and whether the hydrocarbon
adsorbs as beads or forms a continuous lm around the surface.
This will depend on the relative gas bubble and hydrocarbon
droplet sizes as well as the intensity of turbulence. Generally
the average bubble diameters are approximately two orders of
magnitude larger than that of the hydrocarbon droplets, suggest-
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 423
T
a
b
l
e
7
E
m
p
i
r
i
c
a
l
c
o
r
r
e
l
a
t
i
o
n
s
f
o
r
t
h
e
p
r
e
d
i
c
t
i
o
n
o
f
o
v
e
r
a
l
l
v
o
l
u
m
e
t
r
i
c
o
x
y
g
e
n
t
r
a
n
s
f
e
r
c
o
e
f

c
i
e
n
t
(
d
e
r
i
v
e
d
f
r
o
m
d
i
m
e
n
s
i
o
n
l
e
s
s
a
n
a
l
y
s
i
s
)
S
y
s
t
e
m
L
i
q
u
i
d
p
h
a
s
e
(
s
)
E
q
u
a
t
i
o
n
R
e
f
e
r
e
n
c
e
B
u
b
b
l
e
c
o
l
u
m
n
;
D
t
=
1
5
.
2

6
0
c
m
;
V
s
=
3

4
0
c
m
/
s
A
q
u
e
o
u
s
:
g
l
y
c
o
l
,
m
e
t
h
a
n
o
l
,
C
C
l
4
,
N
a
C
l
,
N
a
2
S
O
3
_
K
L
a
D
2 t
D
L
_
2
=
0
.
6
_

L
D
L
_
0
.
5
_
g
D
2 t

_
0
.
6
2
_
g
D
3 t

2 L
_
0
.
3
1

1
.
1
G
A
k
i
t
a
a
n
d
Y
o
s
h
i
d
a
[
9
9
]
S
T
R
;
o
n
e
6
-
b
l
a
d
e
t
u
r
b
i
n
e
;
f
o
u
r
b
a
f

e
s
;
D
t
=
2
5
c
m
;
N
=
3
0
0

6
0
0
r
p
m
;
V
s
=
0
.
3
8
1
c
m
/
s
A
q
u
e
o
u
s
:
g
l
y
c
e
r
o
l
,
m
i
l
l
e
t
-
j
e
l
l
y
,
p
o
l
y
a
c
r
y
l
a
t
e
(
P
A
N
a
)
,
C
M
C
K
L
a
D
2 i
D
L
=
0
.
0
6
_
D
2 i
N

L
_
1
.
5
_
D
i
N
2
g
_
0
.
1
9
_

L
D
L
_
0
.
5
_

L
V
s

_
0
.
6
_
N
D
i
V
s
_
0
.
3
2
[
1
+
2
(

N
)
0
.
5
]

0
.
0
7
Y
a
g
i
a
n
d
Y
o
s
h
i
d
a
[
1
6
7
]
B
u
b
b
l
e
c
o
l
u
m
n
;
D
t
=
1
4
.
5
5
c
m
;
V
s
=
2
.
7
8
c
m
/
s
A
q
u
e
o
u
s
:
s
u
c
r
o
s
e
,
C
M
C
,
s
o
d
i
u
m
p
o
l
y
a
c
r
y
l
a
t
e
K
L
a
D
t
D
L
=
0
.
0
9
_
v
L
D
L
_
0
.
5
_
g
D
3 t
v
2 L
_
0
.
3
9
_
g
D
2 t

_
0
.
7
5
_
V
s

g
D
t
_
1
.
0
_
1
+

_
V
B

D
3
2
_

1
N
a
k
a
n
o
h
a
n
d
Y
o
s
h
i
d
a
[
1
7
4
]
B
u
b
b
l
e
c
o
l
u
m
n
;
D
t
=
1
0
a
n
d
1
9
c
m
;
V
s
=
4
.
2

3
8
c
m
/
s
A
q
u
e
o
u
s
:
s
u
c
r
o
s
e
,
m
e
t
h
a
n
o
l
,
n
-
b
u
t
a
n
o
l
K
L
a
V
s
g
=
1
4
.
9
_
V
s

_
1
.
7
6
_

4 L
g

3
_

0
.
2
4
8
_

L
_
0
.
2
4
3
_

L
D
L
_

0
.
6
0
4
H
i
k
i
t
a
e
t
a
l
.
[
1
7
5
]
S
T
R
;
f
o
u
r
b
a
f

e
s
;
V
=
2
L
;
N
=
1
0
0

1
3
0
0
r
p
m
A
q
u
e
o
u
s
:
g
l
y
c
e
r
i
n
e
,
C
M
C
,
t
r
i
t
o
n
C
F
-
3
2
-
g
l
a
s
s
b
e
a
d
s
K
L
a
D
2 i
D
L
=
1
.
4
1

1
0

3
_

L
D
L
_
0
.
5
_
D
2 i
N

L
_
0
.
6
7
_

L
N
2
D
3 i

_
1
.
2
9
A
l
b
a
l
e
t
a
l
.
[
1
1
6
]
B
u
b
b
l
e
c
o
l
u
m
n
;
D
t
=
1
0
,
1
4
,
2
1
.
8
a
n
d
3
0
c
m
;
V
s
=
0
.
9
8

1
5
.
6
c
m
/
s
A
q
u
e
o
u
s
:
g
l
y
c
e
r
o
l
,
g
l
y
c
o
l
,
B
a
C
l
2
,
N
a
2
S
O
4
K
L
a

D
L
g

L
=
2
.
2
5
_

L
D
L
_
0
.
5
_

3
g

4 L
_
0
.
1
3
6
_

N
D
t
_

0
.
0
9
0
5

1
.
2
6
G
K
o
i
d
e
e
t
a
l
.
[
1
0
5
]
B
u
b
b
l
e
c
o
l
u
m
n
;
D
t
=
2
3
a
n
d
7
6
c
m
;
V
=
4
0
a
n
d
1
0
0
0
L
A
q
u
e
o
u
s
:
C
M
C
;
c
a
r
b
o
p
o
l
K
L
a
D
2 t
D
L
=
1
2
C
1

1
.
0
7
n
1
/
3
_
k
/

L
D
1

n
t
D
V
1

n
s
_
1
/
2
_
D
n t
V
2

n
s
k
/

L
_
(
2
+
n
)
/
1
(
1
+
n
)
_
V
s

g
D
t
_
2
(
1
1
n

4
)
/
3
9
(
1
+
n
)
_
g
D
2 t

_
3
/
5
K
a
w
a
s
e
e
t
a
l
.
[
7
7
]
B
u
b
b
l
e
c
o
l
u
m
n
;
D
t
=
1
4
,
2
1
.
8
a
n
d
3
0
c
m
A
q
u
e
o
u
s
:
m
e
t
h
a
n
o
l
,
e
t
h
a
n
o
l
,
b
u
t
a
n
o
l
,
h
e
x
a
n
o
l
,
o
c
t
a
n
o
l
-
c
a
l
c
i
u
m
a
l
g
i
n
a
t
e
g
e
l
p
a
r
t
i
c
l
e
s
,
p
o
l
y
s
t
y
r
e
n
e
p
a
r
t
i
c
l
e
s
K
L
a

L
D
L
g
=
1
2
.
9
(
v
L
/
D
L
)
0
.
5
(
g

4 L
/

3
)

0
.
1
5
9
(
g
D
2 t

L
/

0
.
1
8
4

1
.
3
G
[
0
.
4
7
+
0
.
5
3
e
x
p
(

4
1
.
4

k
1
/

L
V
s
l
(
d
b
V
s
l

L
/

L
)

1
/
2
)
]
1
+
0
.
6
2

S
/
L
S
a
l
v
a
c
i
o
n
e
t
a
l
.
[
1
5
4
] ing that in the case of a continuous hydrocarbon lm, a large
number of hydrocarbon droplets coalesce and spread around
the gas bubble [35,53]. Hydrodynamic instabilities can lead to
the disruption or removal of this supercial hydrocarbon lm
[35,38,39]. Oil has also been observed to trap air bubbles under
conditions of particularly low agitation rates of 300700 rpm
[144].
For beading hydrocarbons, the pathway from gas to water
may occur in series with no direct gas to hydrocarbon trans-
fer taking place [61,171,176]. For spreading hydrocarbons, on
the other hand, direct gas to hydrocarbon contact is likely to
occur. Linek and Benes [176] and Das et al. [142] suggest,
however, that transfer takes place from the gas bubble into the
aqueous and hydrocarbon phases in both spreading and beading
hydrocarbons.
Matsumura et al. [171] developed correlations to predict
interphasic K
L
a in a dispersion containing discrete gas bubbles
and hydrocarbon droplets for transport through the aqueous lm
surrounding the gas bubbles (K
L
a
GW
) and the aqueous lm sur-
rounding the hydrocarbon droplets (K
L
a
OW
). These correlations
predicted increased K
L
a values with n-C
1418
alkane addition up
to 10%, the maximum concentration examined, to within 15%
of their experimental data.
Interphasic K
L
a was also considered by Rols and Goma
[177] and Rols et al. [35] who proposed several possible
alternate mechanisms for oxygen transport. They present the
most probable scenario as that in which the hydrocarbon is
adsorbed onto the bubble surface and oxygen transport takes
place from the gas to the aqueous phase via the hydrocarbon,
either directly to an organism adsorbed on the hydrocarbon
droplet, or indirectly through the aqueous phase. Roles et al.
[35] measured the resistances to oxygen transfer between the
gas, hydrocarbon and water phases for n-dodecane and forane
F66E and report transfer coefcients as (K
L
)
GO
=1.29 (K
L
)
GW
;
(K
L
)
OW
=0.97 (K
L
)
GW
for n-dodecane and (K
L
)
GO
=2.02
(K
L
)
GW
; (K
L
)
OW
=0.95 (K
L
)
GW
for forane F66E. These results
suggest that the main resistance to oxygen transfer is located
in the aqueous boundary layer at the hydrocarbonaqueous
interface.
Empirical correlations for K
L
a prediction have also been
obtained by dimensional analysis (Table 7). Here the physi-
cal properties of the gas and liquid, which could conceivably
inuence K
L
a, namely surface tension, viscosity, diffusivity
and density, were considered in the development of the cor-
relations. These have found application mainly in viscous
aqueous solutions or dispersions of polymers or short chain
hydrocarbons such as carboxymethyl cellulose [77,116,174],
Na-polyacrylate [174], Ca-alginate glycol [154], glycerine
[116], glycol [99,105], glycerol [105]; carbopol [77], methanol
[99,154,175], butanol [154,175] and hexanol and octanol [154].
These correlations invariably predict an improvement in K
L
a
with decreased surface tension and viscosity. Dimensional anal-
ysis has been shown to be a feasible option for the incorporation
of the effect of the physical properties conferred the hydro-
carbon and has potential for extension into K
L
a prediction
in hydrocarbonaqueous dispersions of longer chain hydrocar-
bons.
424 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
In addition to empirical correlations, several mechanisms
have been modelled for systems containing two immisci-
ble liquid phases. These models have been classied into
homogeneous [178,179,180] and heterogeneous [52,180184]
models. Homogeneous models ignore bubbledroplet geome-
try. In addition, the fraction of the dispersed phase is assumed
to be equally uniform at the gasliquid interface and in
the bulk liquid. Heterogeneous models consider the geome-
try such as the bubbledroplet distance and dropletdroplet
distance. Heterogeneous models are generally preferred from
a physical perspective. However, homogeneous models are
numerically simpler requiring less detailed parameter values,
and with their shorter computation times are preferred for sim-
ple cases where they describe the experimental data equally
well.
However, to date, both homogeneous and heterogeneous
models are valid only for the process conditions relevant to the
study under consideration. Further, these models involve the use
of a number of physical parameters which can neither be accu-
rately predicted nor measured and for which insufcient data
have been reported in literature [180]. Consequently, the use of
these models for the quantication of K
L
a behaviour is limited
and further research is required for the development of a uni-
ed model to explain oxygen transfer in hydrocarbonaqueous
systems.
5. Concluding remarks
In hydrocarbonaqueous dispersions, the hydrocarbon
impacts markedly on K
L
a, but in widely differing manners
depending on the hydrocarbon type and concentration, the
process conditions and the geometric constraints under consid-
eration. This gives rise to three distinct K
L
a behavioural trends
which have been classied as type 1 behaviour (K
L
a increase to
a maximum, with subsequent decease), type 2 behaviour (K
L
a
increase to a maximum with no decrease) and type 3 behaviour
(no increase in K
L
a or K
L
a decrease). Under conditions where
K
L
a is depressed, the enhanced oxygen solubility in hydrocar-
bons may not be able to compensate for the greater oxygen
transfer rate required in hydrocarbon-based bioprocesses due
to the absence of oxygen in the molecular structure. In view
of the importance of an adequate oxygen transfer rate for the
optimisation of the organisms performance and its potential
for hydrocarbon bioconversion, the need for predictive models
to provide reliable estimates of K
L
a in hydrocarbonaqueous
dispersions has been realised.
Currently, models for K
L
a prediction in the presence of
hydrocarbons are empirically derived and based either on a cor-
relation of K
L
a with power input and gas velocity in non-viscous
aqueous solutions, or on dimensional analyses applied to vis-
cous aqueous solutions and dispersions. The correlation of K
L
a
to power input and gas velocity is made applicable to a range
of hydrocarbon dispersions, including the potentially impor-
tant bioconversion substrates n-hexadecane and n-dodecane, by
the incorporation of the hydrocarbon concentration or its uid
properties into the equation. However, these correlations pre-
dict solely a K
L
a increase or a K
L
a decrease, depending on
the behavioural trend in the K
L
a data used to formulate the
correlation. Further, correlations based on dimensional analy-
sis procedures are limited to only a few hydrocarbons, none of
which is generally considered suitable for bioconversion pur-
poses.
The K
L
a trends in hydrocarbonaqueous dispersions are
shaped by the pressures imposed by uid turbulence and physic-
ochemical properties of the dispersion, through alteration of a
number of parameters such as Sauter mean diameter, gasliquid
interface rigidity, gas hold up, surface tension, viscosity and
diffusivity. The K
L
a behaviour is further complicated by the
interactions between the pressures imposed by the turbulence
and uid properties, which may result in different K
L
a trends
under similar conditions of agitation and system geometry, or
over similar hydrocarbon concentration ranges and comparable
hydrocarbon types.
Enhanced turbulence has a positive inuence on K
L
a in all
hydrocarbonaqueous dispersions by increasing the interfacial
area available for transfer anddecreasingthe resistance tomolec-
ular diffusion. The interfacial area available increases directly
through a shear-induced decrease in the Sauter mean bubble
diameter and indirectly through an increased gas hold up as a
result of the smaller bubbles. The resistance to molecular diffu-
sion decreases due to a reduction in the thickness of the stagnant
uid layer surrounding the bubble.
The uid properties, surface tension and viscosity, are strong
inuences in K
L
a behaviour through their effect on bubble
break up and coalescence, albeit with opposite consequences.
Hydrocarbons tend to decrease surface tension, resulting in a
smaller Sauter mean bubble diameter with a lower rise veloc-
ity and, therefore, an increase in gas hold up, both of which
serve to amplify K
L
a through increased available transfer area.
On the other hand, hydrocarbons also increase the viscosity
which damps turbulence, thereby reducing K
L
a macroscopi-
cally through attenuated bubble break up, and microscopically
through increased resistance to molecular diffusion. Viscos-
ity also enhances bubble coalescence and the entrapment of
a greater amount of air per bubble during formation, which
together with attenuated bubble break up, decreases K
L
a through
a decrease in interfacial area. The effect on gas hold up is not
always negative, however, since the larger bubbles leave the dis-
persion more rapidly and a shift in size distribution to smaller
bubbles may occur.
The positive response of K
L
a to enhanced turbulence and
lower surface tension is limited by the effect of these parameters
on the gasliquid interface rigidity. As the Sauter mean diameter
decreases, so does the bubble surface mobility, until a threshold
minimum diameter is reached at which the bubble behaves as a
rigid sphere with no internal circulation. So although the inter-
facial area is maximal, K
L
a may be decreased through a decline
in effective transfer across the interface. This is exacerbated
by an increased resistance to molecular diffusion imposed by
the concomitant increase in viscosity with a decrease in surface
tension.
This review provides a fundamental conceptual understand-
ing of the mechanisms which dene the exact behaviour of
K
L
a in response to changes in turbulence and uid properties,
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 425
and quanties this behaviour in terms of the parameters which
underpin the response. It is envisaged that through the eluci-
dation and quantication of the parameters which are directly
responsible for the behaviour of K
L
a in hydrocarbonaqueous
dispersions, current knowledge of existing models may be suc-
cessfully extended in order to predict the criteria for optimal
oxygen transfer in hydrocarbon-based bioprocesses.
Acknowledgements
The authors gratefully acknowledge the DST-NRF Centre
of Excellence in Catalysis (c*change) and the University of
Stellenbosch for funding of this research. LDC Correia also
acknowledges bursary funding from and DST-NRF Centre of
Excellence in Catalysis (c*change).
References
[1] J.L. Shennan, J.D. Levi, The growth of yeasts on hydrocarbons, in: D.J.D.
Hockenhull (Ed.), Progress in Industrial Microbiology, Churchill Living-
stone, Edinburgh, 1974, pp. 157.
[2] S. Fukui, A. Tanaka, Production of useful compounds from alkane media
in Japan, in: Advances in Biochemical Engineering, Springer, Berlin,
1980, pp. 135.
[3] M.E. Singer, W.R. Finnerty, Microbial metabolism of straight-chain and
branched alkanes, in: R.M. Atlas (Ed.), PetroleumMicrobiology, Macmil-
lan, New York, 1984, pp. 159.
[4] H. Preusting, R. van Houten, A. Hoefs, E.K. van Langenberghe, O.
Favre-Bulle, B. Witholt, High cell density cultivation of Pseudomonas
oleovorans: growth and production of poly(3-hydroxyalkanoates) in two-
liquid phase batch and fed-batch systems, Biotechnol. Bioeng. 41 (1993)
550556.
[5] K. Jung, W. Hazenberg, M. Prieto, B. Witholt, Two-stage continuous
process development for the production of medium-chain-length poly(3-
hydroxyalkanoates), Biotechnol. Bioeng. 72 (2001) 1924.
[6] B. Kessler, B. Witholt, Poly(3-hydroxyalkanoates), in: M.C. Flickinger,
S.W. Drew (Eds.), Encyclopaedia of Bioprocess Technology: Fermen-
tation, Biocatalysis and Bioseparation, John Wiley and Sons, 1999, p.
2756.
[7] N. Kosaric, Biosurfactants, in: M. Roehr (Ed.), Biotechnology: Products
of Primary Metabolism, 2nd ed., VCH, Weinheim, 1996.
[8] E.-C. Chan, C.-S. Cheng, Y.-H. Hsu, Continuous production of dicar-
boxylic acid by immobilized Pseudomonas aeruginosa cells, J. Ferment.
Bioeng. 83 (1997) 157160.
[9] E.-C. Chan, J. Kuo, Biotransformation of dicarboxylic acid by immo-
bilized Cryptococcus cells, Enzyme Microb. Technol. 20 (1997) 585
589.
[10] S. Mauersberger, M. Ohkuma, W.-H. Schunck, M. Takagi, Candida mal-
tosa, in: K. Wolf (Ed.), Nonconventional Yeasts in Biotechnology: A
Handbook, Springer-Verlag, Berlin, 1996, pp. 411580.
[11] T. Juretzek, H.-J. Wang, J.-M. Nicaud, S. Mauersberger, G. Barth,
Comparison of promoter suitable for regulated overexpression of
-galactosidase in the alkane-utilizing yeast Yarrowia lipolytica, Biotech-
nol. Bioprocess. Eng. 5 (2000) 320326.
[12] R.G. Mathys, O.M. Kut, B. Witholt, Alkanol removal from the apolar
phase of a two-liquid phase bioconversion system. Part 1. Comparison
of a less volatile and a more volatile in-situ extraction solvent for the
separation of 1-octanol by distillation, J. Chem. Technol. Biotechnol. 71
(1998) 315325.
[13] R.G. Mathys, A. Schmid, O.M. Kut, B. Witholt, Alkanol removal fromthe
apolar phase of a two-liquid phase bioconversion system. Part 2. Effect of
fermentation mediumon batch distillation, J. Chem. Technol. Biotechnol.
71 (1998) 326334.
[14] R.G. Mathys, A. Schmid, B. Witholt, Integrated two-liquid phase bio-
conversion and product-recovery processes for the oxidation of alkanes:
process design and economic evaluation, Biotechnol. Bioeng. 64 (1999)
459477.
[15] M.L. Shuler, F. Kargi, Bioprocess engineering: basic concepts, in: N.R.
Amundson (Ed.), Physical and Chemical Engineering Sciences, 2nd ed.,
Prentice-Hall Inc., Upper Saddle River, 2002, pp. 1553.
[16] B. Bandyopadhyay, A.E. Humphrey, H. Taguchi, Dynamic measurement
of the volumetric oxygen transfer coefcient in fermentation systems,
Biotechnol. Bioeng. 9 (1967) 533544.
[17] M. Moo-Young, Microbial reactor design for synthetic protein produc-
tion, Can. J. Chem. Eng. 53 (1975) 113118.
[18] W.A. Darlington, Aerobic hydrocarbon fermentationa practical evalu-
ation, Biotechnol. Bioeng. 6 (1964) 241242.
[19] A.E. Humphrey, Acritical reviewof hydrocarbon fermentations and their
industrial utilization, Biotechnol. Bioeng. 9 (1967) 324.
[20] A. Mimura, M. Sugeno, T. Ooka, I. Takeda, Biochemical engineering
analysis of hydrocarbon fermentation, J. Ferment. Technol. 49 (1971)
245254.
[21] H. Preusting, J. Kingma, G. Huisman, A. Steinb uchel, B. Witholt, For-
mation of polyester blends by a recombinant strain of Pseudomonas
oleovorans: different poly(3-hydroxyalkanoates) are stored in separate
granules, J. Polym. Environ. 1 (1993) 1121.
[22] B. Atkinson, F. Mavituna, Gasliquid mass transfer and mixing, in:
Biochemical Engineering and Biotechnology Handbook, McMillian Pub-
lishers, New York, 1991, pp. 697751.
[23] P.M. Doran, Bioprocess Engineering Principles, 1st ed., Academic Press,
New York, 1997, 439 pp.
[24] C.W. Wong, J.H. Shiuan, Effect of additives on mass transfer in an aerated
mixing vessel, Chem. Eng. Commun. 43 (1986) 133145.
[25] M. Moo-Young, H.W. Blanch, Design of biochemical reactors: mass
transfer criteria for simple and complex systems, in: Reactors and Reac-
tions, Springer, Berlin, 1981, pp. 169.
[26] P.H. Calderbank, Physical rate processes in industrial fermentation. Part
I. The interfacial area in gasliquid contacting with mechanical agitation,
Trans. I. Chem. Eng. 36 (1958) 443459.
[27] P.H. Calderbank, Physical rate processes in industrial fermentation. Part
II. Mass transfer coefcients in gasliquid contacting with and without
mechanical agitation, Trans. I. Chem. Eng. 37 (1959) 173185.
[28] P.H. Calderbank, M.B. Moo-Young, The continuous phase heat and
mass transfer properties of dispersions, Chem. Eng. Sci. 16 (1961) 39
54.
[29] L.-K. Ju, C.S. Ho, Oxygen diffusion coefcient and solubility in n-
hexadecane, Biotechnol. Bioeng. 34 (1989) 12211224.
[30] J. Makranczy, K. Megyery-Balog, L. Rusz, L. Patyi, Solubility of gases
in normal-alkanes, Hung. J. Ind. Chem. 4 (1976) 269280.
[31] P.J. Hesse, R. Battino, P. Scharlin, E. Wilhelm, Solubility of gases in
liquids. 20. Solubility of He, Ne, Ar, Kr, N
2
, O
2
, CH
4
, CF
4
and SF
6
in n-
alkanes n-C
l
H
2l+2
(6 <l <16) at 298.15 K, J. Chem. Eng. Data 41 (1996)
195201.
[32] C. Blanc, M. Batiste, Solubility coefcient of oxygen in parrafns, Bull.
Centre Recherches de Pau 4 (1970) 235241.
[33] C.S. Ho, L.-K. Ju, R.F. Baddour, Enhancing penicillin fermentations
by increased oxygen solubility through the addition of n-hexadecane,
Biotechnol. Bioeng. 36 (1990) 11101118.
[34] R.J. Wilcock, R. Battino, W.F. Danforth, E. Wilhelm, Solubilities of gases
in liquids. II. The solubilities of He, Ne, Ar, Kr, O
2
, N
2
, CO, CO
2
, CH
4
,
CF
4
, and SF
6
in n-octane 1-octanol, n-decane, and 1-decanol, J. Chem.
Thermodyn. 10 (1978) 817822.
[35] J.L. Rols, J.S. Condoret, C. Fonade, G. Goma, Mechanism of enhanced
oxygen transfer in fermentation using emulsied oxygen-vectors,
Biotechnol. Bioeng. 35 (1990) 427435.
[36] S. Jia, M. Wang, P. Kahar, Y. Park, M. Okabe, Enhancement of yeast fer-
mentation by addition of oxygen vectors in air-lift bioreactor, J. Ferment.
Bioeng. 84 (1997) 176178.
[37] W. Jianlong, Enhancement of citric acid production by Aspergillus niger
using n-dodecane as an oxygen-vector, Process Biochem. 35 (2000)
10791083.
426 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
[38] A.-I. Galaction, D. Cascaval, C. Oniscu, M. Turnea, Enhancement of
oxygen mass transfer in stirred bioreactors using oxygen-vectors. 1.
Simulated fermentation broths, Bioproc. Biosyst. Eng. 26 (2004) 231
238.
[39] D. Cascaval, A.-I. Galaction, E. Folescu, M. Turnea, Comparative study
on the effects of n-dodecane addition on oxygen transfer in stirred biore-
actors for simulated, bacterial and yeasts broths, Biochem. Eng. J. 31
(2006) 5666.
[40] T.L. da Silva, V. Calado, N. Silva, R.L. Mendes, S.S. Alves, J.M.T. Vas-
concelos, A. Reis, Effects of hydrocarbon additions on gasliquid mass
transfer coefcients in biphasic bioreactors, Biotechnol. Bioprocess. Eng.
11 (2006) 245250.
[41] T.L. da Silva, A. Mendes, R.L. Mendes, V. Calado, S.S. Alves, J.M.T.
Vasconcelos, A. Reis, Effect of n-dodecane on Crypthecodinium cohnii
fermentations and DHA production, J. Ind. Microbiol. Biotechnol. 33
(2006) 408416.
[42] E.S. Thomsen, J.C. Gjaldbaek, The solubility of hydrogen, nitrogen, oxy-
gen and ethane in normal hydrocarbons, Acta Chem. Scand. 17 (1963)
127133.
[43] T.-H. Li, T.-L. Chen, Enhancement of glucose oxidase fermentation by
addition of hydrocarbons, J. Ferment. Bioeng. 78 (1994) 298303.
[44] W. Jianlong, Production of citric acid by immobilized Aspergillus niger
using a rotating biological contactor (RBC), Bioresour. Technol. 75
(2000) 245247.
[45] S. Jia, P. Li, Y.S. Park, M. Okabe, Enhanced oxygen transfer in tower
bioreactor on addition of liquid hydrocarbons, J. Ferment. Bioeng. 82
(1996) 191193.
[46] K.G. Clarke, P.C. Williams, M.S. Smit, S.T.L. Harrison, Enhancement and
repression of the volumetric oxygen transfer coefcient through hydro-
carbon addition and its inuence on oxygen transfer rate in stirred tank
bioreactors, Biochem. Eng. J. 28 (2006) 237242.
[47] C.S. Ho, L.-K. Ju, Effects of microorganisms on effective oxygen dif-
fusion coefcients and solubilities in fermentation media, Biotechnol.
Bioeng. 32 (1988) 313325.
[48] L.D.C. Correia, K.G. Clarke, C. Aldrich, S.T.L. Harrison, The inu-
ence of gasliquid interfacial area on the oxygen transfer coefcient
in alkane-aqueous suspensions, in: South African Chemical Engineer-
ing Congress, International Convention Centre, Durban, South Africa,
September 2022, 2006.
[49] D.R. Nielsen, A.J. Daugulis, P.J. McLellan, Arestructured framework for
modelling oxygen transfer in two-phase partitioning bioreactors, Biotech-
nol. Bioeng. 91 (2005) 773777.
[50] A. Mor ao, C.I. Maia, M.M.R. Fonseca, J.M.T. Vasconcelos, S.S. Alves,
Effect of antifoam addition on gasliquid mass transfer in stirred fermen-
tors, Bioproc. Biosyst. Eng. 20 (1999) 165172.
[51] H.-S. Liu, W.-C. Chiung, Y.-C. Wang, Effect of lard oil, olive oil and
castor oil on oxygen transfer in an agitated fermentor, Biotechnol. Tech.
8 (1994) 1720.
[52] J.D. McMillian, D.J.C. Wang, Enhanced oxygen transfer using oil-in-
water dispersions, Ann. N.Y. Acad. Sci. 506 (1987) 569582.
[53] J.L. Rols, G. Goma, Enhanced oxygen transfer rates in fermentation using
soybean oil-in-water dispersions, Biotechnol. Lett. 13 (1991) 712.
[54] K. Koide, S. Yamazoe, S. Harada, Effects of surface-active substances on
gas holdup and gasliquid mass transfer in bubble column, J. Chem. Eng.
Jpn. 18 (1985) 287292.
[55] E. Dumont, Y. Andr` es, P. Le Cloirec, Effect of organic solvents on oxy-
gen mass transfer in multiphase systems: application to bioreactors in
environmental protection, Biochem. Eng. J. 30 (2006) 245252.
[56] S. Jia, G. Chen, P. Kahar, D.B. Choi, M. Okabe, Effect of soybean oil on
oxygen transfer in the production of tetracycline with an airlift bioreactor,
J. Biosci. Bioeng. 87 (1999) 825827.
[57] A.H.G. Cents, Mass Transfer and Hydrodynamics in Stirred
GasLiquidLiquid Contactors, Ph.D. Thesis, Universiteit of Twente,
Twente, 2003, pp. 1264.
[58] F. Yoshida, T. Yamane, Y. Miyamoto, Oxygen absorption into oil-in-water
emulsions. A study on hydrocarbon fermentors, Ind. Eng. Chem. Proc.
Des. Dev. 9 (1970) 570577.
[59] A. Kundu, E. Dumont, A.-M. Duquenne, H. Delmas, Mass transfer char-
acteristics in gasliquidliquid system, Can. J. Chem. Eng. 81 (2003)
640646.
[60] Y. Bi, G.A. Hill, R.J. Sumner, Enhancement of the overall volumetric
oxygen transfer coefcient in a stirred tank bioreactor using ethanol, Can.
J. Chem. Eng. 79 (2001) 463467.
[61] I.T.M. Hassan, C.W. Robinson, Oxygen transfer in mechanically agitated
aqueous systems containing dispersed hydrocarbon, Biotechnol. Bioeng.
19 (1977) 661682.
[62] M.T. Ces ario, W.A. Beverloo, J. Tramper, H.H. Beeftink, Enhancement of
gasliquid mass transfer rate of apolar pollutants in the biological waste
gas treatment by a dispersed organic solvent, Enzyme Microb. Technol.
21 (1997) 578588.
[63] D.R. Nielsen, A.J. Daugulis, P.J. McLellan, Anovel method of simulating
oxygen mass transfer in two-phase partitioning bioreactors, Biotechnol.
Bioeng. 83 (2003) 735742.
[64] E. Dumont, Y. Andr` es, P. Le Cloirec, Mass transfer coefcients of styrene
and oxygen into silicone oil emulsions in a bubble reactor, Chem. Eng.
Sci. 61 (2006) 56125619.
[65] J.D. McMillian, D.I.C. Wang, Mechanisms of oxygen transfer enhance-
ment during submerged cultivation in peruorochemical-in-water
dispersions, Ann. N.Y. Acad. Sci. 589 (1990) 283300.
[66] F. Yoshida, Y. Miura, Gas absorption in agitated gasliquid contactors,
Ind. Eng. Chem. Proc. Des. Dev. 2 (1963) 263268.
[67] A. Benedek, W.J. Heideger, Effect of additives on mass transfer in turbine
aeration, Biotechnol. Bioeng. 13 (1971) 663684.
[68] K. Akita, F. Yoshida, Bubble size, interfacial area, and liquid-phase mass
transfer coefcient in bubble columns, Ind. Eng. Chem. Proc. Des. Dev.
13 (1974) 8491.
[69] C.W. Robinson, C.R. Wilke, Simultaneous measurement of interfacial
area and mass transfer coefcients for a well-mixed gas dispersion in
aqueous electrolyte solutions, AIChE J. 20 (1974) 285294.
[70] M.M.L. de Figueiredo, P.H. Calderbank, The scale-up of aerated mixing
vessels for specied oxygen dissolution rates, Chem. Eng. Sci. 34 (1979)
13331338.
[71] I.T.M. Hassan, C.W. Robinson, Mass transfer coefcient in mechani-
cally agitated gas-aqueous electrolyte dispersions, Can. J. Chem. Eng. 58
(1980) 198205.
[72] I.T.M. Hassan, C.W. Robinson, Mass-transfer-effective bubble coales-
cence frequency and specic interfacial area in a mechanically agitated
gasliquid contactor, Chem. Eng. Sci. 35 (1980) 12771289.
[73] G.A. Hughmark, Power requirements and interfacial area in gasliquid
turbine agitated systems, Ind. Eng. Chem. Proc. Des. Dev. 19 (1980)
638641.
[74] A. Schumpe, W.-D. Deckwer, Gas holdups, specic interfacial areas, and
mass transfer coefcients of aerated carboxymethyl cellulose solutions
in a bubble column, Ind. Eng. Chem. Proc. Des. Dev. 21 (1982) 706
711.
[75] T. Sridhar, O.E. Potter, Interfacial areas in gasliquid stirred vessels,
Chem. Eng. Sci. 35 (1980) 683695.
[76] C.S. Ho, M.J. Stalker, R.F. Baddour, The oxygen transfer coefcient in
aerated stirred reactors and its correlation with oxygen diffusion coef-
cients, in: C.S. Ho, J.Y. Oldshue (Eds.), Biotechnology Processes:
Scale-up and Mixing, American Institute of Chemical Engineers, New
York, 1987, pp. 8595.
[77] Y. Kawase, B. Halard, M. Moo-Young, Theoretical prediction of volu-
metric mass transfer coefcients in bubble columns for Newtonian and
non-Newtonian uids, Chem. Eng. Sci. 42 (1987) 16091617.
[78] M. Schmitz, A. Steiff, P.-M. Weinspach, Gas/liquid interfacial area per
unit volume and volumetric mass transfer coefcient in stirred slurry
reactors, Chem. Eng. Technol. 10 (1987) 204215.
[79] A.M. Al Taweel, Y.H. Cheng, Effect of surface tension on gas/liquid
contacting in a mechanically-agitated tank with stator, Trans. I. Chem.
Eng. 73 (1995) 654660.
[80] M. Barigou, M. Greaves, Gas holdup and interfacial area distributions
in a mechanically agitated gasliquid contactor, Trans. I. Chem. Eng. 74
(1996) 397405.
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 427
[81] F. Garcia-Ochoa, E. Gomez, Theoretical prediction of gasliquid mass
transfer coefcient, specic area and hold-up in sparged stirred tanks,
Chem. Eng. Sci. 59 (2004) 24892501.
[82] V. Linek, M. Kordac, T. Moucha, Mechanism of mass transfer from bub-
bles in dispersions. Part II. Mass transfer coefcients in stirred gasliquid
reactor and bubble column, Chem. Eng. Process. 44 (2005) 121130.
[83] J.O. Hinze, Fundamentals of the hydrodynamic mechanism of splitting
in dispersion processes, AIChE J. 1 (1955) 289295.
[84] J.C. Lee, D.L. Meyrick, gasliquid interfacial area in salt solutions in an
agitated tank, Trans. I. Chem. Eng. 48 (1970) T37T45.
[85] T. Sridhar, O.E. Potter, Gas hold up and bubble diameters in pressurized
gasliquid stirred vessels, Ind. Eng. Chem. Fundam. 19 (1980) 2126.
[86] M. Fukuma, K. Muroyama, A. Yasunishi, Properties of bubble swarm in
a slurry bubble column, J. Chem. Eng. Jpn. 20 (1987) 2833.
[87] Y. Sun, S. Furusaki, Mean bubble diameter and oxygen transfer coefcient
in a three-phase uidized bed bioreactor, J. Chem. Eng. Jpn. 21 (1988) 20
24.
[88] C.T. OConnor, E.W. Randall, C.M. Goodall, Measurement of the effects
of physical and chemical variables on bubble size, Int. J. Miner. Process.
28 (1990) 139149.
[89] R. Parthasarathy, G.J. Jameson, N. Ahmed, Bubble breakup in stirred
vessels: predicting the Sauter mean diameter, Trans. I. Chem. Eng. 69
(1991) 295301.
[90] K. Takahashi, W.J. McManamey, A.W. Nienow, Bubble size distributions
in impeller region in a gas-sparged vessel agitated by a Rushton turbine,
J. Chem. Eng. Jpn. 25 (1992) 427432.
[91] R. Parthasarathy, N. Ahmed, Sauter mean and maximumbubble diameters
in aerated stirred vessels, Trans. I. Chem. Eng. 72 (1994) 565572.
[92] V. Machon, A.W. Pacek, A.W. Nienow, Bubble size in electrolyte and
alcohol solutions in a turbulent stirred vessel, Trans. I. Chem. Eng. 75
(1997) 339348.
[93] A.W. Pacek, C.C. Man, A.W. Nienow, Coalescence rates in water
(aqueous)-in-oil and oil-in-water (aqueous) dispersions, in: Proceedings
of the 9th European Conference on Mixing, Paris, Groupe Francois de
Geie des Procedes, March 1821, 1997.
[94] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, Experimental and modelling
study of gas dispersion in a double turbine stirred tank, Chem. Eng. Sci.
57 (2002) 487496.
[95] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, A.J. Serralheiro, Bubble size
in aerated stirred tanks, Chem. Eng. J. 89 (2002) 109117.
[96] B. Hu, A.W. Pacek, E.H. Stitt, A.W. Nienow, Bubble sizes in agitated
airalcohol systems with and without particles: turbulent and transitional
ow, Chem. Eng. Sci. 60 (2005) 63716377.
[97] G.A. Hughmark, Holdup and mass transfer in bubble columns, Ind. Eng.
Chem. Proc. Des. Dev. 6 (1967) 218220.
[98] L.L. van Dierendonck, J.M.F. Fortuin, D. Vanderbos, The specic con-
tact area in gasliquid reactors, in: Proceedings of the 4th European
Symposium on Chemical Reactor Engineering, Brussels, September 9,
1968.
[99] K. Akita, F. Yoshida, Gas holdup and volumetric mass transfer coefcient
in bubble columns. Effects of liquid properties, Ind. Eng. Chem. Proc.
Des. Dev. 12 (1973) 7680.
[100] D.N. Miller, Scale-up of agitated vessels gasliquid mass transfer, AIChE
J. 20 (1974) 445453.
[101] H. Hikita, H. Kikukawa, Liquid-phase mixing in bubble columns: effect
of liquid properties, Chem. Eng. J. 8 (1974) 191197.
[102] Z. Sterbacek, M. Sachova, Retention of gas and pressure drops in vertical
agitated and nonagitated liquidgas systems, Int. Chem. Eng. 16 (1976)
104109.
[103] I.T.M. Hassan, C.W. Robinson, Stirred-tank mechanical power require-
ment and gas holdup in aerated aqueous phases, AIChE J. 23 (1977)
4856.
[104] B.G. Kelkar, S.P. Godbole, M.F. Honath, Y.T. Shah, N.L. Carr, W.-D.
Deckwer, Effect of addition of alcohols on gas holdup and backmixing
in bubble columns, AIChE J. 29 (1983) 361369.
[105] K. Koide, H. Sato, S. Iwamoto, Gas holdup and volumetric liquid-phase
mass transfer coefcient in bubble column with draught tube and with
gas dispersion into annulus, J. Chem. Eng. Jpn. 16 (1983) 407413.
[106] K. Koide, A. Takazawa, M. Komuru, H. Matsunaga, Gas holdup and vol-
umetric liquid-phase mass transfer coefcient in solid-suspended bubble
columns, J. Chem. Eng. Jpn. 17 (1984) 459466.
[107] Y. Sun, T. Nozawa, S. Furusaki, Gas holdup and volumetric oxygen trans-
fer coefcient in a three-phase uidized bed bioreactor, J. Chem. Eng. Jpn.
21 (1988) 1520.
[108] Y. Kawase, S. Umeno, T. Kumagai, The prediction of gas hold-up in
bubble column reactors: Newtonian and non-Newtonian uids, Chem.
Eng. J. 50 (1992) 17.
[109] M. Nocentini, D. Fajner, G. Pasquali, F. Magelli, gasliquid mass transfer
and holdup in vessels stirred with multiple Rushton turbines: water and
waterglycerol solutions, Ind. Eng. Chem. Res. 32 (1993) 1926.
[110] S.J. Arjunwadkar, K. Saravanan, P.R. Kulkarni, A.B. Pandit, Gasliquid
mass transfer in dual impeller bioreactor, Biochem. Eng. J. 1 (1998)
99106.
[111] V. Linek, M. Kordac, M. Fujasov a, T. Moucha, Gasliquid mass transfer
coefcient in stirred tanks interpreted through models of idealized eddy
structure of turbulence in the bubble vicinity, Chem. Eng. Process. 43
(2004) 15111517.
[112] K. Rostami, W. Fu, M. Moo-Young, Mass transfer studies in stirred airlift
reactor, Chem. Eng. Commun. 192 (2005) 108124.
[113] M.R. Mehrnia, J. Towghi, B. Bonakdarpour, M.M. Akbarnejad, Gas
hold-up and oxygen transfer in a draft-tube airlift bioreactor with
petroleum-based liquids, Biochem. Eng. J. 22 (2005) 105110.
[114] M. Barigou, M. Greaves, Bubble-size distributions in a mechani-
cally agitated gasliquid contactor, Chem. Eng. Sci. 47 (1992) 2009
2025.
[115] K.R. Westerterp, L.L. van Dierendonck, J.A. de Kraa, Interfacial areas in
agitated gasliquid contactors, Chem. Eng. Sci. 18 (1963) 157176.
[116] R.S. Albal, Y.T. Shah, A. Schumpe, N.L. Carr, Mass transfer in multiphase
agitated contactors, Chem. Eng. J. 27 (1983) 6180.
[117] R. Sch afer, C. Merten, G. Eigenberger, Bubble size distributions in a
bubble column reactor under industrial conditions, Exp. Therm. Fluid
Sci. 26 (2002) 595604.
[118] R. Lemoine, B.I. Morsi, Hydrodynamic and mass transfer parameters
in agitated reactors. Part II. Gas-holdup, Sauter mean bubble diameters,
volumetric mass transfer coefcients, gasliquid interfacial areas, and
liquid-side mass transfer coefcients, Int. J. Chem. React. Eng. 3 (2005)
2061.
[119] R. Krishna, J.W.A. De Swart, J. Ellenberger, G.B. Martina, C. Maretto,
Gas holdup in slurry bubble columns: effect of column diameter and slurry
concentrations, AIChE J. 43 (1997) 311316.
[120] H. Chaumat, A.M. Billet-Duquenne, F. Augier, C. Mathieu, H. Delmas,
Mass transfer inbubble columnfor industrial conditions-effects of organic
medium, gas and liquid ow rates and column design, Chem. Eng. Sci.
60 (2005) 59305936.
[121] C.O. Vandu, R. Krishna, Volumetric mass transfer coefcients in slurry
bubble columns operating in the churn-turbulent owregime, Chem. Eng.
Process. 43 (2004) 987995.
[122] T. Paaschen, A. Lubbert, New experimental results on the mass transfer
from single oxygen bubbles into water as the liquid phase, in: Proceed-
ings of the 4th International Symposium in Bioreactor and Bioprocess
Fluid Dynamics, Edinburgh, BHR Group Conference Series, July 13,
1997.
[123] J.C. Lamont, D.S. Scott, An eddy cell model of mass transfer into the
surface of a turbulent liquid, AIChE J. 16 (1970) 513519.
[124] V. Linek, J. Mayrhoferov a, J. Monerov a, The inuence of diffusivity on
liquid phase mass transfer in solutions of electrolytes, Chem. Eng. Sci.
25 (1970) 10331045.
[125] W.T. Koetsier, D. Thoenes, Mass transfer in a closed stirred gas/liquid
contactor. Part 2. The liquid phase mass transfer coefcient k
l
, Chem.
Eng. J. 5 (1973) 7175.
[126] B.D. Prasher, G.B. Wills, Mass transfer in an agitated vessel, Ind. Eng.
Chem. Proc. Des. Dev. 12 (1973) 351354.
[127] N.C. Panja, D. Phaneswara Rao, Measurement of gasliquid parameters
in a mechanically agitated contactor, Chem. Eng. J. 52 (1993) 121129.
[128] J.F. Perez, O.C. Sandall, Gas absorption by non-Newtonian uids in
agitated vessels, AIChE J. 20 (1974) 770775.
428 K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429
[129] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, 1st ed.,
John Wiley and Sons, New York, 1960, 780 pp.
[130] Y. Kawase, B. Halard, M. Moo-Young, Liquid-phase mass transfer coef-
cients in bioreactors, Biotechnol. Bioeng. 39 (1992) 11331140.
[131] Y. Kawase, M. Moo-Young, Correlations for liquid-phase mass transfer
coefcients in bubble column reactors with Newtonian and non-
Newtonian uids, Can. J. Chem. Eng. 70 (1992) 4854.
[132] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, Gasliquid mass transfer coef-
cient in stirred tanks interpreted through bubble contamination kinetics,
Chem. Eng. Process. 43 (2004) 823830.
[133] F. Garcia-Ochoa, E. Gomez, Prediction of gasliquid mass transfer coef-
cient in sparged stirred tank bioreactors, Biotechnol. Bioeng. 92 (2005)
761772.
[134] P.C. Mena, M.N. Pons, J.A. Teixeira, F.A. Rocha, Using image analysis
in the study of multiphase gas absorption, Chem. Eng. Sci. 60 (2005)
51445150.
[135] D.R. Raymond, S.A. Zieminski, Mass transfer and drag coefcients of
bubbles rising in dilute aqueous solutions, AIChE J. 17 (1971) 57
65.
[136] D. Hammerton, F.H. Garner, Gas absorption from single bubbles, Trans.
I. Chem. Eng. 32 (1954) S18S24.
[137] R.M. Grifth, Mass transfer from drops and bubbles, Chem. Eng. Sci. 12
(1960) 198213.
[138] V. Linek, J. Mayrhoferov a, The chemical method for the determination
of the interfacial area: the inuence of absorption rate on the hold-up and
on the interfacial area in a heterogeneous gasliquid system, Chem. Eng.
Sci. 24 (1969) 481496.
[139] F.J. Montes, M.A. Galan, R.L. Cerro, Mass transfer from oscillating
bubbles in bioreactors, Chem. Eng. Sci. 54 (1999) 31273136.
[140] P.H. Calderbank, Mass transfer, in: V.W. Uhl, J.B. Gray (Eds.), Mix-
ing: Theory and Practice, Academic Press, New York, 1967, pp. 1
114.
[141] N. Midoux, J.-C. Charpentier, Mechanically agitated gasliquid reactors.
Part II. Interfacial area, Int. Chem. Eng. 24 (1984) 452481.
[142] T.R. Das, A. Bandopadhyay, R. Parthasarathy, R. Kumar, Gasliquid
interfacial area in stirred vessels: the effect of an immiscible liquid phase,
Chem. Eng. Sci. 40 (1985) 209214.
[143] N. Pulido-Mayoral, E. Galindo, Phases dispersion and oxygen transfer
in a simulated fermentation broth containing castor oil and proteins,
Biotechnol. Prog. 20 (2004) 16081613.
[144] E. Galindo, A.W. Pacek, A.W. Nienow, Study of drop and bubble sizes in
a simulated mycelial fermentation broth of up to four phases, Biotechnol.
Bioeng. 69 (2000) 213221.
[145] G. Marrucci, A theory of coalescence, Chem. Eng. Sci. 24 (1969)
975985.
[146] S.A. Zieminski, M.M. Caron, R.B. Blackmore, Behaviour of air bubbles
in dilute aqueous solutions, Ind. Eng. Chem. Fundam. 6 (1967) 233
242.
[147] B.H. Junker, Measurement of bubble and pellet size distributions: past
and current image analysis technology, Bioproc. Biosyst. Eng. 29 (2006)
185206.
[148] G. Marrucci, L. Nicodemo, Coalescence of gas bubbles in aqueous solu-
tions of inorganic electrolytes, Chem. Eng. Sci. 22 (1967) 12571265.
[149] G. Keitel, U. Onken, The effect of solutes on bubble size in airwater
dispersions, Chem. Eng. Commun. 17 (1982) 8598.
[150] G. Keitel, U. Onken, Inhibition of bubble coalescence by solutes in
air/water dispersions, Chem. Eng. Sci. 37 (1982) 16351638.
[151] S.A. Zieminski, R.C. Whittemore, Behaviour of gas bubbles in aqueous
electrolyte solutions, Chem. Eng. Sci. 26 (1971) 509520.
[152] Y. Kawase, M. Moo-Young, Mass transfer at a free surface in stirred tank
bioreactors, Trans. I. Chem. Eng. 68 (1990) 189194.
[153] U. Parasu Veera, K.L. Kataria, J.B. Joshi, Effect of supercial gas velocity
on gas hold-up proles in foaming liquids in bubble column reactors,
Chem. Eng. J. 99 (2004) 5358.
[154] J.L. Salvacion, M. Murayama, K. Ohtaguchi, K. Koide, Effects of alco-
hols on gas holdup and volumetric liquid-phase mass transfer coefcient
in gel-particle-suspended bubble column, J. Chem. Eng. Jpn. 28 (1995)
434442.
[155] A.S. Khare, J.B. Joshi, Effect of ne particles on gas hold-up in three-
phase sparged reactors, Chem. Eng. J. 44 (1990) 1125.
[156] P.R. Gogate, A.A.C.M. Beenackers, A.B. Pandit, Multiple-impeller sys-
tems with a special emphasis on bioreactors: a critical review, Biochem.
Eng. J. 6 (2000) 109144.
[157] R.V. Calabrese, T.P.K. Chang, P.T. Dang, Drop breakup in turbulent
stirred-tank contactors. Part I. Effect of dispersed-phase viscosity, AIChE
J. 32 (1986) 657666.
[158] H. Li, A. Prakash, Heat transfer andhydrodynamics ina three-phase slurry
bubble column, Ind. Eng. Chem. Res. 36 (1997) 46884694.
[159] K. Sch ugerl, J. L ucke, U. Oels, Bubble column reactors, in: Advances in
Biochemical Engineering, Springer, Berlin, 1977, pp. 184.
[160] S. Wachi, A.G. Jones, T.P. Elson, Flow dynamics in a draft-tube bub-
ble column using various liquids, Chem. Eng. Sci. 46 (1991) 657
663.
[161] V. Machon, J. Vieck, A.W. Nienow, Soloman, Some effects of pseudo-
plasticity on hold-up in aerated, agitated vessels, J. Chem. Eng. 19 (1980)
6774.
[162] A. Elgozali, V. Linek, M. Fialov a, O. Wein, J. Zahradnk, Inuence of
viscosity and surface tension on performance of gasliquid contactors
with ejector type gas distributor, Chem. Eng. Sci. 57 (2002) 29872994.
[163] S.D. Kim, Y. Kang, Heat and mass transfer in three-phase uidized-bed
reactorsan overview, Chem. Eng. Sci. 52 (1997) 36393660.
[164] A. Lekhal, R.V. Chaudhari, A.M. Wilhelm, H. Delmas, Gasliquid mass
transfer in gasliquidliquid dispersions, Chem. Eng. Sci. 52 (1997)
40694077.
[165] S. Yamamoto, N. Shiragami, H. Unno, H. Honda, Analysis of oxygen
transfer enhancement by oxygen carrier in the autotrophic cultivation of
Alcaligenes eutrophus under low oxygen partial pressure, J. Chem. Eng.
Jpn. 27 (1994) 449454.
[166] F.H. Garner, P.J. Haycock, Circulation in liquid drops, Proc. R. Soc. Lond.
Ser. A 252 (1959) 457475.
[167] H. Yagi, F. Yoshida, Gas absorption by Newtonian and non-Newtonian
uids in sparged agitated vessels, Ind. Eng. Chem. Proc. Des. Dev. 14
(1975) 488493.
[168] A.B. van der Meer, A.A.C.M. Beenackers, R. Burghard, N.H. Mulder, J.J.
Fok, Gas/liquid mass transfer in a four-phase stirred fermentor: effects of
organic phase hold-up and surfactant concentration, Chem. Eng. Sci. 47
(1992) 23692374.
[169] H. Yagi, F. Yoshida, Oxygen absorption in fermenters-Effects of surfac-
tants, antifoaming agents, and sterilized cells, J. Ferment. Technol. 52
(1974) 905916.
[170] J.M.T. Vasconcelos, J.M.L. Rodrigues, S.C.P. Orvalho, S.S. Alves, R.L.
Mendes, A. Reis, Effect of contaminants on mass transfer coefcients
in bubble column and airlift contactors, Chem. Eng. Sci. 58 (2003)
14311440.
[171] M. Matsumura, M. Obara, H. Yoshitome, J. Kobayashi, Oxygen equilib-
rium distribution and its transfer in an airwateroil system, J. Ferment.
Technol. 50 (1972) 742750.
[172] S. Zhao, S.G. Kuttuva, L.-K. Ju, Oxygen transfer characteristics of
multiple-phase dispersions simulating water-in-oil xanthan fermenta-
tions, Bioprocess Eng. 20 (1999) 313323.
[173] G.D. Zhang, W.F. Cai, C.J. Xu, M. Zhou, A general enhancement factor
model of the physical absorption of gases in multiphase systems, Chem.
Eng. Sci. 61 (2006) 558568.
[174] M. Nakanoh, F. Yoshida, Gas absorption by Newtonian and non-
Newtonian liquids in a bubble column, Ind. Eng. Chem. Proc. Des. Dev.
19 (1980) 190195.
[175] H. Hikita, S. Asai, K. Tanigawa, K. Segawa, M. Kitao, The volumetric
liquid-phase mass transfer coefcient in bubble columns, Chem. Eng. J.
22 (1981) 6169.
[176] V. Linek, P. Benes, A study of the mechanism of gas absorption into
oilwater emulsions, Chem. Eng. Sci. 31 (1976) 10371046.
[177] J.L. Rols, G. Goma, Enhancement of oxygen transfer rates in fermentation
using oxygen-vectors, Biotechnol. Adv. 7 (1989) 114.
[178] W.J. Bruining, G.E.H. Joosten, A.A.C.M. Beenackers, H. Hofman,
Enhancement of gasliquid mass transfer by a dispersed second liquid
phase, Chem. Eng. Sci. 41 (1986) 18731877.
K.G. Clarke, L.D.C. Correia / Biochemical Engineering Journal 39 (2008) 405429 429
[179] A.H.G. Cents, D.W.F. Brilman, G.F. Versteeg, Gas absorption in an
agitated gasliquidliquid system, Chem. Eng. Sci. 56 (2001) 1075
1083.
[180] E. Dumont, H. Delmas, Mass transfer enhancement of gas absorption
in oil-in-water systems: a review, Chem. Eng. Process. 42 (2003) 419
438.
[181] B.H. Junker, T.A. Hatton, D.I.C. Wang, Oxygen transfer enhancement in
aqueous/peruorocarbon fermentation systems. I. Experimental observa-
tions, Biotechnol. Bioeng. 35 (1990) 578585.
[182] B.H. Junker, D.I.C. Wang, T.A. Hatton, Oxygen transfer enhancement in
aqueous/peruorocarbon fermentation systems. II. Theoretical analysis,
Biotechnol. Bioeng. 35 (1990) 586597.
[183] C.J. van Ede, R. van Houten, A.A.C.M. Beenackers, Enhancement of gas
to water mass transfer rates by a dispersed organic phase, Chem. Eng.
Sci. 50 (1995) 29112922.
[184] D.W.F. Brilman, M.J.V. Goldschmidt, G.F. Versteeg, W.P.M. van Swaaij,
Heterogeneous mass transfer models for gas absorption in multiphase
systems, Chem. Eng. Sci. 55 (2000) 27932812.

Potrebbero piacerti anche