You are on page 1of 13

Biochemical Engineering Journal 39 (2008) 164–176

Design and characterisation of a miniature stirred bioreactor


system for parallel microbial fermentations
N.K. Gill a , M. Appleton b , F. Baganz a , G.J. Lye a,∗
a The Advanced Centre for Biochemical Engineering, Department of Biochemical Engineering, University College London,
Torrington Place, London, WC1E 7JE, UK
b Bioxplore, 50 Moxon Street, Barnet, Hertfordshire, EN5 5TS, UK

Received 2 August 2007; accepted 3 September 2007

Abstract
The establishment of a high productivity microbial fermentation process requires the experimental investigation of many interacting variables.
In order to speed up this procedure a novel miniature stirred bioreactor system is described which enables parallel operation of 4–16 independently
controlled fermentations. Each miniature bioreactor is of standard geometry (100 mL maximum working volume) and is fitted with a magnetically
driven six-blade miniature turbine impeller (di = 20 mm, di /dT = 1/3) operating in the range 100–2000 rpm. Aeration is achieved via a sintered sparger
at flow rates in the range of 0–2 vvm. Continuous on-line monitoring of each bioreactor is possible using miniature pH, dissolved oxygen and
temperature probes, while PC-based software enables independent bioreactor control and real-time visualisation of parameters monitored on-line. In
addition, a new optical density probe is described that enables on-line estimation of biomass growth kinetics without the need for repeated sampling
of individual bioreactors. Initial characterisation of the bioreactor involved quantification of the volumetric oxygen mass transfer coefficient as
a function of agitation and aeration rates. The maximum kL a value obtained was 0.11 s−1 . The reproducibility of E. coli TOP10 pQR239 and
B. subtilis ATCC6633 fermentations was shown in four parallel fermentations of each organism. For E. coli (1000 rpm, 1 vvm) the maximum
specific growth rate, μmax , was 0.68 ± 0.01 h−1 and the final biomass concentration obtained, Xfinal , was 3.8 ± 0.05 g L−1 . Similarly for B. subtilis
(1500 rpm, 1 vmm) μmax was 0.45 ± 0.01 h−1 and Xfinal was 9.0 ± 0.06 g L−1 . Biomass growth kinetics increased with increases in agitation and
aeration rates and the oxygen enrichment for control of DOT levels enabled μmax and Xfinal as high as 0.93 h−1 and 8.1 g L−1 respectively to be
achieved. Preliminary, scale-up studies with E. coli in the miniature bioreactor (100 mL working volume) and a laboratory scale 2 L bioreactor
(1.5 L working volume) were performed at matched kL a values. Very similar growth kinetics were observed at both scales giving μmax values of
0.94 and 0.97 h−1 , and Xfinal values of 5.3 and 5.5 g L−1 respectively. The miniature bioreactor system described here thus provides a useful tool
for the parallel evaluation and optimisation of microbial fermentation processes.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Miniature bioreactor; Parallel operation; Fermentation; On-line monitoring

1. Introduction strains product synthesis can be further enhanced by study of cul-


ture medium composition, nutrient feeding regimes and physical
The design and optimisation of industrial fermentation variables such as temperature, pH and dissolved oxygen levels.
processes requires the experimental investigation of many inter- Here large numbers of parallel shake flask or stirred-bioreactor
acting biological and physical variables. Advances in metabolic experiments must be performed because of the number of exper-
engineering and protein evolution techniques now enable the imental variables requiring investigation per strain. Finally, once
rapid creation of large libraries of recombinant microorganisms a production strain is identified, further experimental investiga-
[1]. These are normally evaluated in parallel microwell cultures tion over a range of scales is necessary to establish the operating
and a small number of promising strains identified based on sim- boundaries and robustness of the process for validation purposes
ple initial screens for product yield or activity [2]. For the chosen [2]. At this stage virtually all processes would be performed in
stirred bioreactors because of the dominance of this design in
the chemical and biopharmaceutical sectors.
∗ Corresponding author. Tel.: +44 20 7679 7942. Small scale bioreactor systems, that enable parallel and auto-
E-mail address: g.lye@ucl.ac.uk (G.J. Lye). mated operation of several fermentations simultaneously, have

1369-703X/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.bej.2007.09.001
N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176 165

In this work we report on the design, instrumentation and


Nomenclature characterisation of a novel miniature stirred bioreactor system
(maximum working volume 100 mL) that can support the par-
CL concentration of dissolved oxygen in fermenta-
allel operation of 4–16 independently controlled bioreactors.
tion broth (kg m−3 )
Each bioreactor is of standard geometry being agitated by a
Cp normalised oxygen concentration
single six-blade miniature turbine impeller. Continuous on-line
C* saturated dissolved oxygen concentration
monitoring and control of pH, dissolved oxygen and tempera-
(kg m−3 )
ture is achieved using miniature steam sterilisable probes while
dB width of baffle (mm)
a novel optical density (OD) probe allows on-line estimation
di diameter of impeller (mm)
of biomass growth kinetics. Bioreactor oxygen mass transfer
dT diameter of vessel (mm)
characteristics have been studied as a function of agitation and
DOT dissolved oxygen tension (%)
aeration rates giving a volumetric oxygen mass transfer coef-
hi distance between centre line of impeller and base
ficient, kL a, of up to 0.11 s−1 . Parallel E. coli TOP10 pQR239
of vessel
and B. subtilis ATCC6633 fermentations have been shown to
kL a volumetric oxygen mass transfer coefficient (s−1 )
be highly reproducible. In the case of E. coli TOP10 it has also
mO2 oxygen required for cell maintenance
been shown how biomass growth kinetics and yields vary as
(mol gDCW −1 h−1 )
a function of agitation and aeration conditions. Finally initial
OTRmax maximum oxygen transfer rate (mmol L−1 h−1 )
results using constant kL a as a basis for scale-up have shown
OURmax maximum oxygen uptake rate (mmol L−1 h−1 )
that results obtained in the miniature bioreactor can accurately
P power input (W)
be reproduced in a conventional laboratory scale stirred biore-
t time (s)
actor.
tm mass transfer time, 1/kL a (s)
V volume (L)
2. Materials and methods
vvm volumetric air flow per volume of broth per minute
Xfinal final biomass concentration (g L−1 )
2.1. Chemicals and microorganisms
YX/O2 yield of biomass on oxygen (g mol−1 )

Greek letters The chemicals used in this work were obtained from BDH
μmax maximum specific growth rate (h−1 ) (Dorset, UK) unless otherwise stated and were of the highest
τp probe response time (s) purity available. RO water was used for all experiments. E. coli
υS superficial gas velocity (m s−1 ) TOP10 pQR239, which expresses cyclohexanone monooxyge-
nase (CHMO) under the control of an l-arabinose promotor
[15], and B. subtilis ATCC6633 [16] were used in the fermenta-
tion studies. Cells were maintained as 40% (v/v) glycerol stock
the potential to increase the rate at which the necessary exper- solutions at −80 ◦ C.
iments are performed thus reducing fermentation development
times and costs [3]. In recent years various designs of paral- 2.2. Miniature bioreactor design and instrumentation
lel miniature bioreactor systems have been reported including
stirred tank bioreactors, bubble columns and shake flasks [4–7]. 2.2.1. Design of individual bioreactors
The key design features of many of these will be discussed in The miniature bioreactor was designed to be geometrically
detail later (Section 3.5). While each bioreactor design aims to similar to conventional laboratory scale stirred fermenters. Each
satisfy all the requirements for rapid fermentation process devel- bioreactor consisted of a borosilicate glass vessel (100 mL
opment there is normally a trade-off between throughput and the maximum working volume), allowing visual inspection of the
information content of the data obtained from each experiment contents of the bioreactor, and was sealed with a stainless steel
[2]. This is most clear when comparing related developments head plate, as shown in Fig. 1.
in microwell fermentations [6–10] with the small scale stirred Mixing of the bioreactor was achieved by a magneti-
bioreactor designs considered here. Irrespective of the approach cally driven, six-blade miniature turbine impeller (di = 20 mm,
taken for high throughput fermentation process development di /dT = 1/3) fabricated from PEEK (poly-ether ether ketone).
however, it will be important that advances in throughput are There was a distance of 20 mm between the centre line of the
matched by the creation of microscale downstream process- Rushton turbine and the base of the vessel (di /hi = 1). Eight
ing operations. In this regard the generation of quantitative cylindrical magnets (3 mm in diameter and 2 mm long) were
process data from a number of microwell-based downstream uniformly distributed and embedded into a small PEEK disc
unit operations has recently been reported [3,11,12] as has the 10 mm below the miniature turbine as shown in Fig. 1(A).
automated operation of linked process sequences [13,14]. The The whole impeller structure was mounted onto a hollow,
automated linkage of upstream and downstream operations at non-rotating stainless steel stirrer shaft that screwed into the
such scales now offers the potential for the integrated optimisa- headplate of the bioreactor. This design eliminated the need for
tion of the entire process from fermentation through to purified rotating mechanical seals and reduced the area occupied on the
product. headplate. The PEEK impeller assembly, which was magneti-
166 N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176

Fig. 2. Schematic diagram of in situ optical density probe: (A) dimensions of


probe; (B) arrangement of light source and detectors. Further details given in
Section 2.2.2.

2.2.2. On-line instrumentation of individual bioreactors


Fig. 1. Mechanical drawing showing key design features and dimensions of a Continuous on-line monitoring of each bioreactor was
single miniature bioreactor: (A) cross section through bioreactor; (B) plan view facilitated using a miniature pH probe (Hamilton Bonaduz
of head plate. Further details given in Section 2.2.1. AG, Switzerland), polargraphic oxygen electrode (Hamilton
Bonaduz AG, Switzerland), a narrow K-type thermocouple
(HEL Ltd., UK) and a novel optical probe (HEL Ltd., UK)
cally driven from below, had a stirring range of 100–2000 rpm. for optical density measurement. All probes were steam
Each bioreactor was also equipped with four equally spaced sterilisable.
removable baffles of width 6 mm (dB /dT = 1/10). Aeration of The optical probe, as shown in Fig. 2(A), had a PEEK body
the vessel was via a narrow sparger located directly beneath and was sized to fit each bioreactor headplate. The end of the
the miniature turbine. The sparger was fitted with a 15 ␮m probe, Fig. 2(B), had a white light source and two detectors that
stainless steel sinter to create finer gas bubbles and promote were each located in sealed glass tubes. The light source (LS) and
more efficient oxygen transfer. The air flow rate was manu- one of the detectors (D1) were positioned directly opposite each
ally regulated by a standard laboratory rotameter in the range other approximately 10 mm apart. D1 thus gave a direct mea-
0–200 mL min−1 (Fisher Scientific, UK) and it was possible to surement of the optical density of the broth. A second detector
use atmospheric air alone or oxygen-enriched air via a gas blend- (D2) was placed at right angles to the direction of transmit-
ing system (HEL Ltd., UK). The gas flow rate to each reactor ted light approximately 5 mm away from the beam. The output
was frequently checked in order to ensure constant gas flow of this second detector measured the amount of light scattered
rate. by the broth. The light source had a sufficiently narrow angle
As shown in Fig. 1(B), the headplate accommodated a total of of emission such that any light reaching the right-angle detec-
four probes for continuous on-line monitoring of pH, dissolved tor (D2) by direct transmission was considered negligible. In
oxygen tension (DOT), optical density (OD) and temperature as this work only the data from the optical density signal (D1) is
described in Section 2.2.2. As a result of the limited space avail- reported. A calibration curve of on-line and off-line OD values
able on the headplate a narrow opening in the centre allowed was generated for each type of fermentation completed in the
a thermocouple for temperature sensing to be located within miniature bioreactors.
the thin walled hollow stirrer shaft. All liquid additions and
sample removals were via a single multi-port on the headplate 2.2.3. Design and control of parallel bioreactor systems
which had five openings closed by self-sealing septa (HEL Ltd., To support the individual bioreactors a modular base unit
UK). Liquid additions were made via sterile hypodermic nee- was designed which securely located four miniature bioreac-
dles, securely mounted by luer lock fittings. The exhaust gas port tors and their respective peristaltic pumps (RS Components
was fitted with a stainless steel, water cooled, condenser which Ltd., UK), rotameters, electrical disc heaters and magnetic drive
prevented liquid loss from the medium by evaporation. The inlet assemblies. Cables from the various probes plugged into the
and exhaust gas lines were filtered through 0.2 ␮m filters (Fisher base unit which housed all the associated electronics creat-
Scientific, UK). ing a compact unit with a footprint of 480 by 362 mm. Up
N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176 167

to four of these modular base units can be used in a single bioreactor was inoculated with 2 mL (2%, v/v inoculum) of broth
system. from a shake flask culture (100 mL in a 1 L shaken flask) grown
A custom piece of PC-based software was written which for 14 h on the same medium at 37 ◦ C and 200 rpm on a hor-
allowed independent monitoring and control of up to 16 minia- izontal shaken platform (New Brunswick, USA). The pH was
ture bioreactors. This was based on an architecture originally controlled at 7 (±0.1) by the metered addition of 3 M NaOH and
used for control of automated chemical reactors that has been 3 M H3 PO4 .
reported to be easily programmed and simple to use [17]. Exper- In the case of B. subtilis ATCC6633 chemically defined
iments with multiple operating conditions and set points could growth media was used. The media components were
be conducted by generating an experimental plan that dictated prepared and sterilised (121 ◦ C for 20 min) in five sepa-
which piece of equipment should be running and what it was rate groups [16]. One litre of biomedia consisted of: (a)
required to do. Each experimental plan can consist of several 100 mL of a 200 g L−1 solution of d-glucose, (b) 895 mL
steps with each step being carried out automatically in sequence. of a solution of 11.2 g L−1 (NH4 )2 SO4 , 15.2 g L−1 KH2 PO4 ,
The end to each step is specified by a terminating condition, 12.1 g L−1 K2 HPO4 , 4.0 g L−1 Na2 HPO4 , and 1.1 g L−1
which is any parameter defined by the user, e.g., time, tem- antifoam (polypropylene glycol 2000), (c) 2 mL of a 246 g L−1
perature, pH, OD, DOT, etc. The software also enabled key solution of MgSO4 ·7H2 O, (d) 1 mL of a 147 g L−1 solution of
parameters such as pH, DOT, temperature and optical density CaCl2 ·2H2 O and (e) 2 mL of an acidified (pH 1) solution of trace
to be displayed in real time on interactive graphs for each biore- metals (40 g L−1 FeSO4 ·7H2 O, 5 g L−1 MnSO4 ·H2 O, 2 g L−1
actor. CoCl2 ·6H2 O, 1 g L−1 ZnSO4 ·7H2 O, 1 g L−1 MoO4 Na2 ·2H2 O,
0.5 g L−1 CuCl2 ·2H2 O, 2 g L−1 H3 BO3 ). Each miniature biore-
2.3. Characterisation of bioreactor oxygen transfer rates actor was sterilised as a complete unit (121 ◦ C for 20 min)
containing group (b) media components. After cooling the tem-
The oxygen transfer capability of the miniature bioreac- perature was maintained at 32 ◦ C and the remaining groups of
tor was assessed using the dynamic gassing out technique media components were added aseptically immediately prior to
[18]. Before each experiment 10 g L−1 NaCl was dissolved in inoculation. Each bioreactor was inoculated with 10 mL (10%,
fresh de-ionised water and the dissolved oxygen probe cali- v/v inoculum) of an actively growing culture (100 mL in a 1 L
brated between 0% and 100% saturation by sparging nitrogen shaken flask) with an optical density of approximately 2. The
and air, respectively. All experiments were carried out at a shake flask culture was grown on the same medium at 32 ◦ C
constant temperature of 37 ◦ C at predetermined stirrer speeds and 300 rpm on a horizontal shaken platform. The pH was con-
(1000–2000 rpm) and aeration rates (1–2 vvm) using either trolled at 6.8 (± 0.1) by the metered addition of 3 M NaOH and
atmospheric air or air enriched with oxygen. Assuming the liq- 3 M H3 PO4 .
uid in the bioreactor was well mixed, the volumetric oxygen For both E. coli TOP10 and B. subtilis ATCC6633 fermen-
mass transfer coefficient, kL a, was determined from the mea- tations, agitation rates in the miniature bioreactors were varied
sured dissolved oxygen–time profiles accounting for the probe between 1000 and 2000 rpm and aeration rates varied between
response time according to Eq. (1) [19]: 1 and 2 vvm (certain experiments with E. coli involved aeration
     with oxygen-enriched air maintaining the DOT at a set point
1 −t −t of 30% or 50%). Fermentations were performed in parallel on a
Cp = tm exp − τm exp (1)
tm − τ p tm τp four pot system either under identical agitation and aeration con-
ditions (to show reproducibility) or under different agitation and
where Cp is the normalised dissolved oxygen concentration mea- aeration conditions (to show parallel evaluation of fermentation
sured by the probe at time t, tm equals 1/kL a and τ p is the probe conditions).
response time (18 s at 37 ◦ C). All gassing out experiments were
performed in triplicate with the maximum coefficient of variance 2.5. Analytical techniques
for kL a determination being 6.1%.
In addition to the on-line optical density measurements (Sec-
2.4. Parallel E. coli and B. subtilis fermentations tion 2.2.2), cell growth was also monitored by taking broth
samples of up to 1 mL at regular intervals and measuring the
Two different microorganisms were used in this work. In OD600 of appropriately diluted samples (1 in 2 to 1 in 20 dilu-
the case of E. coli TOP10 pQR239 the growth media con- tions were used) off-line (Ultraspec 4000 spectrophotometer,
sisted of 10 g L−1 each of tryptone, yeast extract, glycerol, Pharmacia Biotech, USA). All biomass concentrations reported
NaCl (Sigma–Aldrich, Poole, UK) and 50 mg L−1 ampicillin here are dry cell weight (DCW) concentrations and were from
(Sigma–Aldrich, Poole, UK). Each miniature bioreactor was appropriate experimentally determined calibration curves for
sterilised as a complete unit (121 ◦ C for 20 min) with all media each microorganism and medium. The DCWs used to create
components (apart from ampicillin) and 0.2 mL L−1 of added the calibration curve were determined in triplicate from known
antifoam (polypropylene glycol 2000). After cooling the temper- volumes of fermentation broth. After centrifugation at 1300 rpm
ature was maintained at 37 ◦ C (±0.2) via an electrical disc heater for 15 min (Eppendorf AG, Germany), cell pellets were washed
positioned beneath the glass vessel. Filter sterilised (0.2 ␮m) once with 10 g L−1 NaCl solution and dried at 100 ◦ C for 24 h
ampicillin was added immediately prior to inoculation. Each in pre-weighed and dried 2 mL Eppendorf tubes.
168 N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176

3. Results and discussion

3.1. Miniature bioreactor design and operation

The basis of the miniature bioreactor design was that it should


be as geometrically similar to conventional laboratory scale
bioreactors as possible and be agitated by a standard minia-
ture turbine impeller. This was dictated by the desire to obtain
quantitative and scaleable data from each miniature bioreactor
and efficient oxygen transfer during microbial fermentations. In
practice these constraints limited the volume of each bioreactor
to a minimum of 100 mL, to ensure that the miniature impeller
was completely submerged in liquid and maximum agitation
rates of up to 2000 rpm could be achieved. In order to simplify
the assembly of each unit, and to minimise any risk of contam-
ination, the need for a rotating mechanical seal on the impeller
drive shaft was overcome by opting for a magnetically driven
stirrer. This was only possible by compromising on the design of
the impeller and the incorporation of a flat disk at the end of the
impeller drive shaft as shown in Fig. 1(A). This was necessary Fig. 3. Influence of bioreactor operating conditions on oxygen uptake kinetics
to house sufficient magnets to enable agitation at rates of up to during dynamic gassing out experiments. From left to right: (—·–·—) 2000 rpm,
2 vvm; (—·-·— ) 2000 rpm, 1 vvm; (- - -) 1500 rpm, 1 vvm; (—·—) 1000 rpm,
2000 rpm without decoupling the impeller from the magnetic
2 vvm; (—) 1000 rpm, 1.5 vvm; (—) 1000 rpm, 1 vvm. Experiments performed
drive. Even with these modifications visual observation showed at 37 ◦ C in 10 g L−1 NaCl solution as described in Section 2.3.
uniform distribution of gas bubbles throughout the entire liq-
uid volume of each vessel provided that the agitation rate was
oxygen uptake is seen to increase with both increasing agitation
1000 rpm or above.
and aeration rates. For each condition the corresponding kL a
To facilitate parallel operation, four miniature bioreactors
values were calculated using Eq. (1) and are plotted in Fig. 4 as
could be assembled and sterilised simultaneously in a typical
a function of stirrer speed. The maximum kL a was determined
laboratory autoclave. A hydrostatic pressure test indicted that
as 0.11 s−1 .
each vessel could typically withstand an applied pressure of
The results in Fig. 4 show that kL a increases linearly with
8 bar before rupturing. An initial medium sterilisation and hold
stirrer speed over the range investigated. Stirrer speed is also
test, in which DOT, pH and OD were monitored over 4 days
seen to have a more significant effect on kL a than increases
showed no signs of contamination. This was confirmed by the
in aeration particularly at aeration rates above 1.5 vvm where
absence of any colonies from medium samples withdrawn peri-
odically and grown on nutrient agar plates at 37 ◦ C for 24 h. To
further facilitate parallel and unattended operation of multiple
units each bioreactor was instrumented with pH, temperature
and dissolved oxygen probes as well as a novel on-line optical
density probe (as described in Section 2.2.2). The small area of
each bioreactor head plate meant that it was necessary to source
the smallest available probes. In order to then have sufficient
ports for liquid additions, sampling and the possibility of con-
tinuous bioreactor operation all liquid handling operations took
place via a specially designed multi-port. Finally, it was pos-
sible to independently monitor and control up to 16 miniature
bioreactors using custom-written, PC-based software.

3.2. Characterisation of bioreactor oxygen transfer


capability

Since the majority of industrial fermentations use aerobic


microorganisms the oxygen transfer capability of the miniature
bioreactor is of great interest. Parameters such as kL a are also
useful for comparing different bioreactor designs and provide a
Fig. 4. Bioreactor oxygen mass transfer coefficient (kL a) as a function of stirrer
useful criterion for scale-up. Fig. 3 shows a series of dynamic speed and aeration rate with atmospheric air: () 1 vvm; () 1.5 vvm; () 2 vvm.
gassing out experiments performed at different stirrer speeds Values of kL a calculated from the dynamic gassing out data shown in Fig. 3 using
and aeration rates for aeration with atmospheric air. The rate of Eq. (1). Solid lines fitted by linear regression.
N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176 169

the improvement in oxygen transfer is minimal. Classically the ducible if performed under identical operating conditions. Fig. 5
impeller power input and aeration rate have been correlated in shows biomass growth kinetics and dissolved oxygen tension
the literature using the well known van’t Riet correlation: (DOT) profiles for typical parallel E. coli fermentations in the
 α four-pot miniature bioreactor system. These were carried out at
Pg β a fixed impeller speed and aeration rate of 1000 rpm and 1 vvm,
kL a = K υS (2)
V respectively which gave a relatively low kL a value of 0.04 s−1
(as measured in 10 g L−1 NaCl). Considering the first fermen-
where Pg /V is the impeller gassed power per unit volume, υS is tation shown in Fig. 5 (bioreactor B1), the entire fermentation
the superficial gas velocity, K is a constant and α and β are expo- lasted a total of 540 min with the end of the exponential cell
nents in the range of 0.4 < α < 1 and 0 < β < 0.7, respectively [18]. growth phase occurring around 300 min. This coincided with
For kL a measurements in ionic solutions, the reported values of the point at which oxygen mass transfer limitation occurred and
the constants K, α and β were 2.0 × 10−3 , 0.7 and 0.2, respec- the measured DOT reached zero. The particular strain of E. coli
tively. Although the van’t Riet correlation was determined for used here is known to have a high specific oxygen demand so the
much larger vessels than used here, fitted with standard Rushton fact that the measured DOT reached zero at this kL a value is not
turbine impellers, the stronger dependency of kL a on Pg /V (and surprising [15]. The period of exponential growth was followed
hence stirrer speed) than on uS shown in Fig. 4 initially suggests by an almost linear increase in biomass concentration for a fur-
that results for the miniature bioreactor are consistent with the ther 160 min at which point the culture entered stationary phase
established theory. At present the Power number of the miniature and ceased to grow. The biomass growth kinetics and dissolved
bioreactor is not known and so absolute values of Pg /V cannot oxygen profiles for the other three bioreactors were very similar
be estimated. apart from the DOT profile in bioreactor B4 which reached zero
DOT somewhat earlier.
3.3. Parallel fermentations and in-situ monitoring Table 1 shows the key kinetic parameters derived from the
individual fermentation profiles shown in Fig. 5. The maxi-
3.3.1. Reproducibility of parallel E. coli fermentations mum specific growth rates were very similar, giving an average
A key requirement of any parallel bioreactor system is that value of 0.68 ± 0.01 h−1 as were the final biomass concentra-
cultivations in separate bioreactors should be highly repro- tions achieved where the average value was 3.8 ± 0.05 g L−1 .

Fig. 5. Parallel batch fermentation kinetics of E. coli TOP10 pQR239 grown under identical conditions: () off-line biomass concentration; (-) DOT. Experiments
performed at 1000 rpm and 1 vvm as described in Section 2.4. B1–B4 refer to bioreactors one to four, respectively.
170 N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176

Table 1
Reproducibility of batch E. coli TOP10 pQR239 fermentation kinetics from four parallel fermentations
Parameter B1 B2 B3 B4 BAV

μmax (h−1 ) 0.68 0.66 0.68 0.69 0.68 ± 0.01


Xfinal (g L−1 ) 3.7 3.8 3.8 3.8 3.8 ± 0.05
tDOT = 0 (min) 290 300 300 243 283 ± 27.2
OURmax (mmol L−1 h−1 ) 18.1 18.4 18.4 18.4 18.3 ± 0.15

Kinetic parameters derived from the fermentation profiles shown in Fig. 5. B1–B4 refer to bioreactors one to four respectively, while BAV indicates the mean values
of the kinetic parameters (error indicated represents one standard deviation).

The time at which the measured DOT reached zero was also com- Similarly, OTR can be estimated from:
parable in the majority of cases. Overall these results indicate
excellent reproducibility of the four-pot system. OTR = kL a(C∗ − CL ) (4)
At the point where the measured DOT first reaches zero, it
where C* is the saturated dissolved oxygen concentration, esti-
can be assumed that the maximum oxygen uptake rate (OURmax )
mated to be 6.9 mg L−1 [21] and CL is the actual concentration of
and the maximum oxygen transfer rate (OTRmax ) are equal. The
dissolved oxygen in the fermentation broth. At the point where
OURmax can thus be estimated from:
the measured DOT first reaches zero, CL is also zero and so Eq.
μX (4) reduces to:
OURmax = + m O2 X (3)
YX/O2 OTRmax = kL aC∗ (5)

where μ is the specific growth rate at the time point when the From Eq. (3) the average OURmax for the fermentations
DOT first reaches zero, X is the corresponding biomass concen- shown in Fig. 5 was calculated to be 18.3 ± 0.15 mmol L−1 h−1 .
tration, YX/O2 is the yield of biomass on oxygen, which was In order to estimate the corresponding OTRmax value from Eq.
taken to be 1.92 g g−1 [20] and m is the oxygen required for cell (5), bioreactor kL a values were measured by further gassing out
maintenance which was taken to be 0.003 mol O2 g−1 h−1 [20]. experiments using spent biomedia containing the antifoam PPG

Fig. 6. Parallel batch fermentation kinetics of B. subtilis ATCC6633 grown under identical conditions: () off-line biomass concentration; (-) DOT. Experiments
performed at 1500 rpm and 1 vvm as described in Section 2.4. B1–B4 refer to bioreactors one to four, respectively.
N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176 171

used during fermentation experiments. The presence of PPG respectively [16] and the contribution from cell maintenance
is known to significantly reduce kL a values [22] and the mea- is negligible. Over a short time period dCL /dt is constant and
sured value of 0.03 s−1 was not surprisingly 25% lower than the kL a was calculated. This method gave a kL a value of
the value measured in electrolyte solution. Using this lower kL a 0.062 ± 0.001 s−1 which is in excellent agreement with that
value the calculated value of OTRmax was thus estimated to be measured using the gassing out technique under the same oper-
18.2 mmol L−1 h−1 which is in excellent agreement with the ating conditions (Table 2).
calculated OURmax values.
3.3.3. Correlation of on-line and off-line optical density
3.3.2. Reproducibility of parallel B. subtilis fermentations measurements
In addition to showing the reproducibility of parallel fermen- As described in Section 2.2.2, each miniature bioreactor was
tations of E. coli, a facultative anaerobe, the flexibility of the equipped with a novel on-line OD probe (Fig. 2). This was
miniature bioreactor system was shown using parallel fermenta- designed to facilitate continuous on-line monitoring of biomass
tions of B. subtilis, a strict aerobe. Fig. 6 shows typical biomass growth kinetics which could be important given the small vol-
growth kinetics and DOT profiles for four parallel B. subtilis ume of each bioreactor and the potential supervision of up to
fermentations in the miniature bioreactors. These fermentations
were carried out at a fixed impeller speed and aeration rate of
1500 rpm and 1 vvm, respectively corresponding to a kL a value
of 0.06 s−1 . Considering the first fermentation shown in Fig. 6
(bioreactor B1) it can be seen that the entire fermentation lasted
660 min with the exponential phase of cell growth lasting until
approximately 540 min. At this point the measured DOT again
reached zero at which point the culture rapidly entered stationary
phase and ceased to grow. As found with E. coli, very similar
fermentation profiles for B. subtilis were obtained in all four
miniature bioreactors.
Table 2 shows the key kinetic parameters derived from the
individual fermentation profiles shown in Fig. 6. The maximum
specific growth rates were again very similar, giving an average
value of 0.45 ± 0.01 h−1 while the average final biomass concen-
tration achieved was 9.0 ± 0.06 g L−1 . The average time taken
for the DOT level to reduce to zero was 565 ± 33.7 min. The level
of variation seen for the B. subtilis fermentations is comparable
to that determined for the earlier E. coli work and again indicates
excellent reproducibility of the four-pot bioreactor system.
In addition to using the gassing out method to determine kL a
values (as described in Section 3.2) in the case of a strict aerobe
like B. subtilis an oxygen mass balance can also be used, where
under steady state conditions oxygen uptake and consumption
rates must balance [21], thus from Eqs. (3) and (4):
dCL
= kL a(C∗ − CL ) − OUR (6)
dt
For this particular medium and strain of B. subtilis CL
and YX/O2 have been reported to be 6.8 mg L−1 and 1.6 g g−1

Table 2
Reproducibility of batch B. subtilis ATCC6633 fermentation kinetics from four
parallel fermentations
Parameter B1 B2 B3 B4 BAV

μmax (h−1 ) 0.46 0.45 0.45 0.44 0.45 ± 0.01


Xfinal (g L−1 ) 9.0 9.0 9.0 9.1 9.0 ± 0.06
tDOT = 0 (min) 529 572 550 608 565 ± 33.7
kL a (s−1 ) 0.061 0.062 0.061 0.059 0.061 ± 0.001 Fig. 7. On-line measurement of optical density in E. coli TOP10 pQR239 fer-
mentations. (A) Comparison of biomass growth kinetics from an individual
Kinetic parameters derived from the fermentation profiles shown in Fig. 6. fermentation: () off-line optical density; (—) on-line optical density. (B) Par-
B1–B4 refer to bioreactors one to four respectively, while BAV indicates the ity plot of on-line and off-line biomass concentration data from nine identical
mean values of the kinetic parameters (error indicated represents one standard fermentations. Experiments performed at 1000 rpm and 1 vvm as described in
deviation). Section 2.4.
172 N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176

Fig. 8. Influence of agitation and aeration rates and oxygen enrichment on batch fermentation kinetics of E. coli TOP10 pQR239: () off-line biomass concentration;
(-) DOT. Experimental conditions: (A) 2000 rpm, 1 vvm; (B) 1000 rpm, 2 vvm; (C) 2000 rpm, 1 vvm and DOT controlled at 30%; (D) 2000 rpm, 1 vvm and DOT
controlled at 50%. Fermentations performed as described in Section 2.4.

16 bioreactor units in parallel by a single operator. As shown and 2000 rpm (aeration rate remained constant at 1 vvm) and
in Fig. 7(A) the growth profile generated during a single E. coli higher aeration rates of 1.5 and 2 vvm (agitation rate remained
fermentation by the on-line OD probe was virtually identical to constant at 1000 rpm). Aeration was with either atmospheric air
that from off-line OD measurements, apart from during the final or oxygen-enriched air utilising the gas-blending capability of
stages of the culture. The OD reading of the probe was initially the miniature bioreactor system.
calibrated using the sterilised culture medium. Calculated val- Fig. 8 shows examples of biomass concentration and DOT
ues of the maximum specific growth rate from the on-line and profiles under the different agitation and aeration conditions. The
off-line data were 0.67 h−1 and 0.68 h−1 , respectively indicating derived kinetic parameters from the entire series of experiments
that the on-line data can be used for quantitative analysis of cell are summarised in Table 3. For aeration with atmospheric air
growth kinetics. Fig. 7(B) shows a parity plot of on-line and off- there is a significant increase in both the maximum growth rate
line biomass concentration data collected from nine identical (0.75–0.94 h−1 ) and final biomass concentration (5.1–5.6 g L−1 )
E. coli fermentations. There is seen to be excellent correlation in all cases when compared to the values obtained with the
between the two types of biomass concentration data indicating lowest agitation and aeration conditions used previously in Sec-
the robustness of the probe after repeated cycles of sterilisation tion 3.3.1 (0.68 h−1 and 3.8 g L−1 ). These increases in cell
and fermentations. growth rate are broadly in line with the measured increases
in kL a values. The largest value of OURmax was calculated to
3.4. Influence of agitation and aeration conditions on E. be 46.8 mmol L−1 h−1 . In all cases, however, oxygen transfer
coli fermentations limitations were observed to remain with the measured DOT
reaching zero at some point during the course of the indi-
The initial E. coli and B. subtilis fermentations reported in vidual fermentations. In order to overcome this problem the
Section 3.3 were carried out at relatively low kL a values and so bioreactors could be aerated with oxygen-enriched air such
oxygen limitations were observed in both cases. To explore the that the DOT was maintained at a constant set point value.
oxygen uptake requirements of the E. coli strain further, a series For agitation at 2000 rpm and aeration at 1 vvm in this case,
of fermentations were carried at higher agitation rates of 1500 graphs C and D in Fig. 8 show that the DOT could be
N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176 173

Table 3
Variation of bioreactor oxygen mass transfer coefficient and E. coli TOP10 pQR239 fermentation kinetics as a function of agitation and aeration conditions including
the use of oxygen enrichment to control DOT levels
Parameter Agitation and aeration conditions

1000 rpm 1500 rpm 2000 rpm 1000 rpm 1000 rpm 2000 rpm 1 vmm (C) 2000 rpm 1 vmm (D)
1 vmm 1 vmm 1 vmm (A) 1.5 vmm 2 vmm (B)

Gas blending – – – – – DOT = 30% DOT = 50%


kL a (s−1 ) 0.04 0.06 0.08 0.06 0.06 – –
μmax (h−1 ) 0.68 0.83 0.94 0.75 0.79 0.86 0.93
Xfinal (g L−1 ) 3.8 5.1 5.3 5.6 5.6 7.6 8.1
OURmax (mmol L−1 h−1 ) 18.4 35.7 46.8 26.3 27.5 ND ND

(A)–(D) correspond to the fermentation profiles shown in Fig. 8. Oxygen mass transfer coefficient and values derived from Fig. 3. N/D = not determined.

maintained and controlled at a minimum level of 30 or 50% ing volumes of 35 mL. However, the oxygen transfer rates are
respectively. In the absence of oxygen limitations the maxi- significantly lower compared to other systems described in this
mum growth rate of the culture increased further to 0.93 h−1 section (0.01 s−1 ).
and the highest biomass concentration obtained was 8.1 g L−1 . In addition to stirred vessels, small scale bubble columns have
While gas blending can perhaps be avoided during the initial also be designed in recent years, high-lighting the importance
stages of fermentation process development it will be neces- and the need for a variety of bioreactor systems to address dif-
sary to implement it during the later stages when high cell ferent process applications. Currently a 200 mL bubble column
density cultures must be attained in the miniature bioreac- is available from Infors, Switzerland and is reported to achieve
tors. kL a values of up to 0.16 s−1 [29,30], with the capacity to operate
up to 16 vessels in parallel. Doig et al. [16] have characterised
3.5. Comparison with other miniature bioreactors a system with up to 48 miniature bubble columns operating in
parallel with working volumes of 2 mL, producing kL a values of
Several miniature bioreactor systems for rapid bioprocess around 0.06 s−1 . Although these systems lack the mixing capa-
development have been reported in recent years. A number of bility of a stirrer, sufficient aeration and some degree of mixing is
designs have been investigated but the recent trend has been achieved by direct sparging. These researchers have also demon-
toward mechanically stirred bioreactors as these are most widely strated good correlation with a laboratory scale bubble column
used in industry for development and large scale operation. The (100 mL) in terms of oxygen transfer and volumetric power con-
majority of the bioreactors have been designed with parallel sumption [31]. Cell cultivations were also scaled up based on
operation in mind and there is a consensus that 10’s of experi- constant kL a to bench scale stirrer bioreactor, and the results
ments be performed in parallel for the systems to be of value. were comparable.
In this respect the “bioreactor block” [23] currently gives the The miniature bioreactor systems described herein are
highest degree of parallelisation, 48, but must be operated on quickly replacing the traditionally popular shake flask for par-
the deck of a dedicated laboratory automation platform. The allel, high throughput operation, given the level of sophisticated
smaller units (1–10 mL) tend to be single use disposable items technology providing large quantities of data. However, some
while the larger ones (100–200 mL) like that described here can researchers have attempted to enhance the design of the shake
be repeatedly steam sterilised. flask to facilitate some degree of online monitoring, these
In terms of agitation and aeration systems those bioreac- include the RAMOS (HITECZang GmbH, Germany) [32] and
tors featuring mechanical agitation tend to have the highest kL a the Fedbatch-pro (DasGip,Germany) [33] systems.
values. Stirrer speeds necessarily increase dramatically as the
working volume of the bioreactors drop below about 10 mL. In
many cases non-standard gas-sparging impellers [23] and gas
Table 4
blending [24,25, this work] have been implemented to main- Comparison of E. coli TOP10 pQR239 fermentations carried out in miniature
tain sufficient oxygen transfer rates. The highest kL a reported to and conventional laboratory bioreactors using constant kL a as a basis for scale-up
date for any of the miniature bioreactors has been for the “biore- (based on aeration with atmospheric air)
actor block”, achieving 0.4 s−1 [23]. kL a values of this order Parameter Miniature bioreactor Laboratory bioreactor
are necessary to support the microbial cell densities that could
Working volume (mL) 100 1500
be achieved in approximately half of the designs that are capa-
Stirrer speed (rpm) 2000 1000
ble of fed-batch or continuous operation. Betts et al. [25] have Aeration rate (vvm) 1.00 0.67
reported kL a values in the order of 0.1 s−1 for a 10 mL stirred kL a (s−1 ) 0.08 0.08
miniature bioreactor (the detailed design for this vessel has been Inoculum (%working volume) 2.0 6.7
previously described by Lamping et al. [26]). The commercially μmax (h−1 ) 0.94 0.97
Xfinal (g L−1 ) 5.3 5.5
available Cellstation (Fluorometrix, USA) [27,28] allows up to
12 stirred bioreactors to be operated in a single run with work- Kinetic parameters derived from Fig. 9.
174 N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176

In order to asses the reliability of miniature high throughput bioreactor. Comparing the results of the miniature bioreactor and
systems as a tool for scale-down studies, it is critical to demon- the 7 L vessel concluded reasonable agreement in terms of max-
strate their ability to mimic the conditions and productivity that imum specific growth rates for E. coli DH5␣ producing plasmid
would be expected at larger scales. A majority of the minia- DNA.
ture systems that have been reported over the years have not Virtually all the designs feature the standard on-line probes
done this, however, some scale comparison studies have been expected for effective bioreactor monitoring and control though
carried out comparing the growth and productivity of Bacil- a number feature novel probes for on-line monitoring of culture
lus subtillis RB50 and the productivity of riboflavin [34] in the optical density [23,27,28, this work]. The striking trend is the
“bioreactor block” [23] and a 7 L bioreactor. Betts et al. [25] switch from standard probe technologies to fluorescent/optical
have carried out more rigorous scale-up studies using a defined probes once the bioreactor volume drops below about 10 mL.
scale-up criterion based on constant power per unit volume given All the designs for which parallel operation has been demon-
the geometric similarity between their vessel and a conventional strated feature dedicated PC-based software necessary for the
supervision of multiple units.

3.6. Scale-up from miniature to laboratory scale


bioreactors

While the miniature stirred bioreactor described here is


capable of automated parallel operation, it is important to
asses the scale-up potential of the results obtained. In this
section the relationship between miniature bioreactor results
and those obtained using the same E. coli strain in a typical
laboratory scale 2 L bioreactor (1.5 L working volume ves-
sel fitted with two top driven Rushton turbine impellers) will
be considered [15]. As a basis for predictive scale translation
experiments were initially performed at matched kL a values
of 0.08 s−1 based on direct measurements of oxygen transfer
rates at the two scales. Table 4 shows the corresponding agi-
tation and aeration conditions. For E. coli cultures at the two
scales Fig. 9(A) shows that there was good agreement between
biomass growth kinetics while Fig. 9(B) shows the correspond-
ing agreement between measured DOT profiles. Experiments
at both scales were performed without gas-blending so the
expected oxygen limitations toward the end of the exponential
growth phase are again seen. Given the 15-fold scale dif-
ference there is excellent agreement between the calculated
maximum specific growth rates and final biomass concen-
trations achieved (Table 4). Similar agreement between the
two scales was determined for matched fermentations at kL a
values over the range 0.06–0.11 s−1 (data not shown). This
gives confidence that results obtained from parallel experi-
ments in the miniature bioreactors can be rapidly translated to
the conventional scales used for fermentation process develop-
ment.

4. Conclusions

A parallel miniature bioreactor system (100 mL working


volume) designed to be geometrically similar to conventional
laboratory scale stirred bioreactors has been constructed and
evaluated in this study. The oxygen transfer characteristics of
the miniature bioreactor were first evaluated in terms of the
oxygen mass transfer coefficient, kL a, as a function of agi-
Fig. 9. Batch fermentation kinetics of E. coli TOP10 pQR239 fermentations
tation and aeration rates. Values as high as 0.11 s−1 could
carried out at miniature (100 mL) and conventional laboratory bioreactor (1.5 L)
scales: (A) off-line biomass concentration; (B) DOT. Experiments performed at be achieved comparable to those found in typical laboratory
matched kL a values as described in Section 2.4. Laboratory bioreactor data from scale bioreactors. For identical fermentations performed in the
Doig et al. [15]. four-pot bioreactor system, excellent reproducibility between
N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176 175

parallel E. coli and B. subtilis fermentations was shown in [10] G.T. John, I. Klimant, C. Wittmann, E. Heinzle, Integrated optical sens-
terms of the average calculated maximum specific growth rates ing of dissolved oxygen in microtitre plates: a novel tool for microbial
(0.68 ± 0.01 h−1 and 0.45 ± 0.01 h−1 for E. coli and B. sub- cultivation, Biotechnol. Bioeng. 81 (2003) 829–836.
[11] K. Rege, M. Pepsin, B. Falcon, L. Steele, M. Heng, High-throughput pro-
tilis, respectively). Biomass growth rates and yields for E. coli cess development for recombinant protein purification, Biotechnol. Bioeng.
could be improved, and oxygen transfer limitations overcome, 93 (2006) 618–630.
by increases in agitation and aeration rates and by the implemen- [12] N.B. Jackson, J.M. Liddell, G.J. Lye, An automated microscale technique
tation of gas blending. Finally, kinetic parameters determined for the quantitative and parallel analysis of microfiltration operations, J.
from miniature bioreactor experiments were shown to be in Membrane Sci. 276 (2006) 31–41.
[13] C. Ferreira-Torres, M. Micheletti, G.J. Lye, Microscale process evaluation
accord with those from a conventional 2 L stirred bioreac- of recombinant biocatalyst libraries: application to Baeyer-Villiger mono-
tor (1.5 L working volume) when experiments were performed oxygenase catalysed lactone synthesis, Bioprocess Biosystems Eng. 28
at matched kL a values. The ability to obtain quantitative and (2005) 83–93.
scaleable data from up to 16 miniature bioreactors in parallel [14] M. Micheletti, T. Barrett, S.D. Doig, F. Baganz, M.S. Levy, J.M.
Woodley, G.J. Lye, Fluid mixing in shaken bioreactors: impli-
makes the system described here a useful tool for the paral-
cations for scale-up predictions from microlitre scale microbial
lel optimisation of microbial fermentation processes. Current and mammalian cell cultures, Chem. Eng. Sci. 61 (2006) 2939–
experiments are aimed at a better characterisation of the engi- 2949.
neering environment within the miniature bioreactors with a [15] S.D. Doig, L.M. O’Sullivan, S. Patel, J.M. Ward, J.M. Woodley, Large
view to improved designs and predictive scale-up to larger biore- scale production of cyclohexanone monooxygenase from Escherichia coli
TOP10 pQR239, Enzyme Microb. Technol. 28 (2001) 265–274.
actors.
[16] S.D. Doig, A. Diep, F. Baganz, Characterisation of a novel miniature bubble
column bioreactor for high throughput cell cultivation, Biochem. Eng. J.
Acknowledgements 23 (2004) 97–105.
[17] M.E. Van Loo, P.E. Lengowski, Automated workstations for parallel syn-
thesis, Org. Process Res. Dev. 6 (2002) 833–840.
The authors would like to thank the UK Joint Infras-
[18] K. van’t Riet, Review of measuring methods and results in non-viscous
tructure Fund (JIF), the Science Research Investment Fund gas liquid mass transfer in stirred vessels, Ind. Eng. Chem. Process Design
(SRIF) and the Gatsby Charitable Foundation for funds Dev. 18 (1979) 357–364.
to establish the UCL Centre for Micro Biochemical Engi- [19] I.J. Dunn, A.J. Einsele, Oxygen transfer coefficients by the dynamic
neering. Financial support from the UK Engineering and method, J. Appl. Chem. Biotechnol. 25 (1975) 707–720.
[20] B. Atkinson, F. Mavituna, Biochemical Engineering and Biotechnology
Physical Sciences Research Council (EPSRC) and HEL
Handbook, second ed., Stockton press, 1999.
Ltd., in the form of an Engineering Doctorate (EngD) stu- [21] Bailey, Ollis, Biochemical Engineering Fundamentals, McGraw-Hill Inter-
dentship for Naveraj Gill, is also acknowledged. The authors national, Singapore, 1986.
are grateful to Martin Peacock for his contribution to this [22] A. Morao, C.I. Maia, M.M.R. Fonseca, J.M.T. Vasconcelos, S.S. Alves,
work. Effect of antifoam addition on gas-liquid mass transfer in stirred fermenters,
Bioprocess Eng. 20 (1999) 165–172.
[23] R. Puskeiler, K. Kaufmann, D. Weuster-Botz, Development, paralleli-
References sation, and automation of a gas inducing millilitre-scale bioreactor for
high-throughput bioprocess design (HTBD), Biotechnol. Bioeng. 89 (2005)
[1] E.G. Hibbert, F. Baganz, H.C. Hailes, J.M. Ward, G.J. Lye, J.M. Woodley, 512–523.
P.A. Dalby, Directed evolution of biocatalytic processes, Biomol. Eng. 22 [24] R. Puskeiler, A. Kusterer, G.T. John, D. Weuster-Botz, Minia-
(2005) 11–19. ture bioreactors for automated high-throughput bioprocess design
[2] S.D. Doig, F. Baganz, G.J. Lye, High throughput screening and process (HTBD): Reproducibility of parallel fed-batch cultivations with
optimisation, in: C. Ratledge, B. Kristiansen (Eds.), Basic Biotechnology, Escherichia coli, Biotechnol. Appl. Biochem. 42 (2005) 227–
third ed., Cambridge University Press, UK, 2006. 235.
[3] G.J. Lye, P. Ayazi-Shamlou, F. Baganz, P.A. Dalby, J.M. Wood- [25] J.I. Betts, S.D. Doig, F. Baganz, Charaterisation and application of a minia-
ley, Accelerated design of bioconversion processes using automated ture 10 mL stirred-tank bioreactor, showing scale-down equivalence with
microscale processing techniques, Trends Biotechnol. 21 (2003) 29– a conventional 7 L reactor, Biotechnol. Prog. 22 (2006) 681–688.
37. [26] S.R. Lamping, H. Zhang, B. Allen, P. Ayazi-Shamlou, Design of a prototype
[4] S. Kumar, C. Wittman, E. Heinzle, Minibioreactors Biotechnol. Lett. 26 miniature bioreactor for high throughput automated bioprocessing, Chem.
(2004) 1–10. Eng. Sci. 58 (2003) 747–758.
[5] D. Weuster-Botz, Parallel reactor systems for bioprocess development, [27] Y. Kostov, P. Harms, L. Randers-Eichhorn, G. Rao, Low-cost microbiore-
Adv. Biochem. Eng Biotechnol. 92 (2005) 125–143. actor for high-throughput bioprocessing, Biotechnol. Bioeng. 72 (2001)
[6] P. Fernandes, J.M.S. Cabral, Review: microlitre/millilitre shaken biore- 346–352.
actors in fermentative and biotransformation processes, Biocatal. [28] P. Harms, Y. Kostov, J.A. French, M. Soliman, M. Anjanappa, A. Ram, G.
Biotransform. 24 (2006) 237–252. Rao, Design and performance of a 24-station high throughput microbiore-
[7] J.I. Betts, F. Baganz, Review: miniature bioreactors: current practices and actor, Biotechnol. Bioeng 93 (2006) 6–13.
future opportunities, Microb. Cell Factories 5 (2006) 21. [29] S. Dilsen, W. Paul, D. Herforth, A. Sandgathe, J. Altenbach-Rehm, R.
[8] W.A. Dutez, L. Ruedi, R. Hermann, K. O’Conner, J. Büchs, B. Witholt, Freudl, C. Wandrey, D. Weuster-Botz, Evaluation of parallel operated
Methods for intense aeration, growth, storeage and replication of bacte- small-scale bubble columns for microbial process development using
rial strains in microtiter plates, Appl. Env. Microbiol. 66 (2000) 2641– Staphylococcus carnosus, J. Biotechnol. 88 (2001) 77–84.
2646. [30] D. Weuster-Botz, J. Altenbach-Rehm, A. Hawrylenko, Process-
[9] I. Elmahdi, F. Baganz, K. Dixon, T. Harrop, D. Sugden, G.J. Lye, pH control engineering characterisation of small-scale bubble columns for microbial
in microwell fermentations of S. erythraea CA340: influence on biomass process development, Bioprocess Biosystems Eng. 24 (2001) 3–11.
growth kinetics and erythromycin biosynthesis, Biochem. Eng. J. 16 (2003) [31] S.D. Doig, K. Ortiz-Ochoa, J.M. Ward, F. Baganz, Character-
299–310. ization of oxygen transfer in miniature and lab-scale bubble
176 N.K. Gill et al. / Biochemical Engineering Journal 39 (2008) 164–176

column bioreactors and comparison of microbial growth perfor- [33] D. Weuster-Botz, J. Altenbach-Rehm, M. Arnold, Parallel substrate feeding
mance based on constant kL a, Biotechnol. Prog. 21 (2005) 1175– and pH control in shaking flasks, Biochem. Eng. J. 7 (2001) 163–170.
1182. [34] B. Knorr, H. Schlieker, H.P. Hohmann, D. Weuster-Botz, Scale-down and
[32] T. Anderlei, J. Büchs, Device for sterile online measurement of the oxygen parallel operation of the riboflavin production process with Bacillus sub-
transfer rate in shaking flasks, Biochem. Eng. J. 7 (2001) 157–162. tilis, Biochem. Eng. J. 33 (2007) 263–274.

You might also like