You are on page 1of 12

Chemical Engineering Journal 213 (2012) 50–61

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Investigation of mixing characteristics in a packed-bed external loop airlift


bioreactor using tomography images
Mian Hamood-ur-Rehman, Yaser Dahman, Farhad Ein-Mozaffari ⇑
Department of Chemical Engineering, Ryerson University, 350 Victoria Street, Toronto, Ontario, Canada M5B 2K3

h i g h l i g h t s

" A new bioreactor, which combines advantages of both packed bed and external loop airlift bioreactors, was introduced.
" 2D and 3D tomograms were utilized to explore the mixing hydrodynamics.
" Mixing efficiency was quantified using mixing time and superficial liquid velocity.
" The effects of design parameters and operating conditions were characterized.
" Installation of the gas redistributor between two rolls of packing had a significant impact on mixing hydrodynamics.

a r t i c l e i n f o a b s t r a c t

Article history: The present research work introduces a recirculated airlift bioreactor (biocontactor), which combines
Received 3 June 2012 advantages of packed bed and external loop airlift bioreactors. This bioreactor is characterized by an
Received in revised form 27 September internal gas redistributor placed between two beds of packing in the riser while working fluid recirculates
2012
through an external downcomer. The main objective of this research work was to characterize the mixing
Accepted 27 September 2012
Available online 5 October 2012
hydrodynamics of this reactor through the non-intrusive flow visualization technique of electrical resis-
tance tomography (ERT). The tomography images were employed to examine the effects of packing, gas
redistributor, sparger configuration, gas flow rate, and liquid height on the mixing time and superficial
Keywords:
Bioreactors
liquid velocity. Results showed that the mixing time in the bioreactor was decreased with an increase
Mixing in the superficial gas velocity in the riser. It was found that at a constant riser superficial gas velocity,
Tomography the mixing time was increased and the superficial liquid velocity decreased when a cross shaped sparger
Packed bed was replaced by a circular sparger, which had the same diameter, and the same number and size of holes.
Superficial liquid velocity Results also showed that mixing time increased and the superficial liquid velocity decreased by reducing
Internal gas redistributor the liquid height in the bioreactor, in addition to installing the internal gas redistributor and packing in
the riser. The presence of the internal gas redistributor and packing reduced the cross-sectional area
available for flow and increased resistance in the liquid flow path. Results, which were obtained from this
study, can be utilized to improve mixing in external loop airlift bioreactors for wider range of
applications.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction Airlift bioreactors are encountered in the fields of aerobic fermen-


tations, waste water treatment, and other operations, where low
Airlift bioreactors have been extensively used in biotechnology shear stress is required [5–7]. The two main types of internal and
industries in recent years in a variety of arrangements and applica- external loop recirculated airlift bioreactors have been the subject
tions. This includes commercial manufacture of pharmaceuticals, of interest for different researchers. The main difference between
enzymes, fragrances, dyes and antibiotics [1–3]. The main advanta- these two types is that the internal loop airlift bioreactors consist
ges of this type of reactors are the simple construction and opera- of two concentric columns while the external loop airlift bioreac-
tion, low investment and operational cost, absence of regions of tors consist of two separate columns connected near the top and
high shear, very fine gas dispersion, enhanced mixing and mass at the bottom by two horizontal sections. This design gives a sep-
transfer performance, and relatively low power requirements [4]. arate gas disengagement section to external loop airlift bioreactors,
which helps in maximizing deaeration [5,7–13].
The external loop airlift bioreactors have low friction in riser
⇑ Corresponding author. Tel.: +1 416 979 5000x4251; fax: +1 416 979 5083. and downcomer and thus provide a better opportunity for mea-
E-mail address: fmozaffa@ryerson.ca (F. Ein-Mozaffari). surement and control of flow conditions. These bioreactors also

1385-8947/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2012.09.106
M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61 51

Nomenclature

Abc cross-sectional area of the bottom connecting section (m2) tc circulation time (s)
Atc cross-sectional area of the top connecting section (m2) Ulr superficial liquid velocity in the riser (m/s)
Ad downcomer cross sectional area (m2) Uld superficial liquid velocity in the downcomer (m/s)
Ar riser cross sectional area (m2) Ugr superficial gas velocity in the riser (m/s)
Dr diameter of riser (m)
Hr circulating liquid height in the riser (m) Greek symbols
Hbc circulating liquid length in the bottom connecting section u dimensionless factor ()
(m) egr gas holdup in the riser ()
Htc circulating liquid length in the top connecting section (m) egd gas holdup in the downcomer ()
tr liquid residence time in the riser (s)

offer effective heat transfer and temperature control. Furthermore, and inverse fluidized beds were combined to construct this novel
they can achieve independent control of the gas input-rate and li- external loop bioreactor. Loh and Liu [6] proposed a novel external
quid velocity using a throttling device at the bottom connecting loop inversed fluidized bed airlift bioreactor by installing a valve at
tube between riser and downcomer of the external loop bioreactor. the bottom between the riser and downcomer sections for the
The external-loop usually reaches nearly total gas disengagement treatment of wastewater containing high phenol concentrations.
at the top section (due to the presence of separate gas-disengage- Zhang et al. [3] designed an internal for the external loop airlift
ment section) giving rise to a higher difference in density or hydro- bioreactor. This internal contained some semicircular holes and
static pressure which results in higher circulation velocity as each hole had a tongue-like plate facing the upstream to break
compared to the internal loop airlift bioreactor and consequently the bubbles and enhance the flow redistribution. Meng et al. [24]
improves mixing and heat transfer in the reactor. On the other combined the design strategies of the external loop airlift bioreac-
hand, the external loop offers lower gas recirculation as compared tor and packed bed bioreactor. They used the woven nylon packing
to internal loop airlift, which results in lower mass transfer as a packed bed in the riser section of the airlift bioreactor. Mohan-
[3,4,14–16]. The applications of external loop airlift bioreactors ty et al. [25] investigated the hydrodynamics of a novel external
have been increasing and making them as one of the most impor- loop airlift bioreactor operating in three vertical stages. The verti-
tant bioreactor configurations. Many investigators have studied cal cylindrical column was fitted with a total of seven internals
external loop bioreactors experimentally and theoretically (four contraction disks and three expansion disks).
[2,3,17–19]. An external loop recirculated airlift bioreactor (biocontactor),
Mixing time and superficial liquid velocity play significant roles with an internal gas redistributor, which was installed between
in the hydrodynamic characterization of the airlift bioreactors. two rolls of packing in the riser, was designed and built. This de-
Mixing quality is a critical parameter, which is widely used in sign combines advantages of both packed bed and external loop
the design, modeling and operation, as well as scale up of external airlift bioreactor. The presence of two beds of packing allows for
loop airlift bioreactors. In some processes, where non-uniform pro- improving residence time and bioconversion when immobilized
files of concentration or temperature can cause side reactions or enzymes or microorganisms are utilized in variety of bioengineer-
reduce the chemical reaction rate, good mixing is very important. ing applications. To characterize the mixing hydrodynamics in this
In external loop airlift bioreactors, sparged air is used to achieve bioreactor, mixing time and superficial liquid velocity were mea-
the mixing. The mixing of gas phase is usually negligible because sured through electrical resistance tomography (ERT), which has
of the high velocity of gas bubbles in the airlift reactor. The super- been employed as a reliable and non-intrusive tool for the direct
ficial liquid velocity is also a key design parameter for the reactors, analysis of the hydrodynamics of multiphase flows in recent years
which affects the residence time of gas, mass transfer coefficient, [20,26–31]. This technology manipulates data obtained from the
turbulence, heat transfer coefficient, shear forces to which the sensors located on the periphery of the vessel in order to get the
microorganisms are exposed, and the mixing time. The cross sec- precise information on the flow regime and concentration distribu-
tional area ratio of downcomer to riser is a physical parameter, tion inside the process vessels [32]. Results obtained using the ERT
which strongly affects the superficial liquid velocity. Any were used to examine the effect of different parameters that
resistance given to the flow path of the liquid in the bioreactor also include having two packed beds in the riser; internal gas redistrib-
affect the liquid velocity in the bioreactor. utor installed between two packed-bed sections; sparger
Gumery et al. [20] utilized tomography technique to explore the configuration; gas flow rate; and liquid height on the mixing
mixing characteristics in a draft-tube airlift reactor. Jin et al. [21] hydrodynamics of the bioreactor. The new contributions of this pa-
studied the oxygen transfer in a pilot plant external air-lift bioreac- per are: (i) a new bioreactor, which combines advantages of both
tor as a function of gas holdup, oxygen transfer coefficient, super- packed and external loop airlift bioreactors, was introduced; (ii)
ficial gas velocity, and dissolved oxygen. To investigate the liquid- Tomogrphy, a reliable and non-intrusive technique, was utilized
phase mixing in the external-loop airlift reactors, Kawase et al. [22] to explore the mixing performance of this reactor; and (iii) the ef-
measured mixing time, circulation time, and downcomer linear li- fects of the design parameters and operating conditions on the
quid velocity. Different configurations of the external loop airlift mixing hydrodynamics were quantified.
bioreactors have been designed by the previous researchers. For in-
stance, Xu and Yu [23] designed a multiple airlifting membrane
bioreactor. The vessel was separated into two compartments, a 2. Experiment
tube-side aerobic compartment and a shell-side anaerobic com-
partment, by installing four sintered stainless steel filter tubes be- 2.1. Experimental setup
tween the top and the bottom sections. Han et al. [10] investigated
the hydrodynamic behavior in a new gas–liquid–solid inverse flu- A packed bed external loop recirculated airlift bioreactor with an
idization airlift bioreactor. The advantages of external loop reactors internal gas redistributor placed in the way between two rolls of
52 M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61

packing in the riser was designed and built. Polyvinyl chloride This gas redistributor was designed to have 415 holes, each of
material was used to construct this bioreactor. Optimal geometries 0.002 m, providing total open area about 3% and was installed be-
in the design of the bioreactor were considered based on results re- tween the two beds of packing in the riser (see Fig. 1). Two rolls of
ported previously in the literatures [4,6,10,33–36]. Detailed sche- fiber glass were used as the packing. Each roll was initially a sheet
matic diagram of the experimental setup is shown in Fig. 1. As of 0.234 m width; 6.3 m length and 0.0254 m thickness, which
shown in this figure, the sparger was installed at the bottom of was tightly rolled and maintained in the cylindrical shape with suf-
the riser and the air flow was controlled using a rotameter. The dif- ficient distance between each layer (1 cm). The roll of packing had
ferent sparger configurations that were examined include the cross a 0.248 m diameter, 0.234 m height, and the porosity of about 99%.
shaped and circular sparger. The superficial gas velocity in the riser
was varied between 0.011 and 0.033 m/s. An internal perforated 2.2. Electrical resistance tomography (ERT) technique
plate was utilized as the gas redistributor. The function of this gas
redistributor was to break and regenerate the gas bubbles exiting Electrical resistance tomography (ERT) technique was used in
the lower packing and before entering the upper packing in the riser. the present study to explore the mixing characteristics in the

Fig. 1. Schematic diagram of the external loop airlift bioreactor utilized in this study.
M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61 53

packed bed external loop airlift bioreactor. This instrument creates 7 was 0.117 m, and the distance between plane 7 and plane 8 was
tomograms showing the cross-sectional distribution of the electri- 0.175 m. The distance between plane 5 and the bottom of the bio-
cal conductivity of the contents within the sensing zones [37]. This reactor was 0.764 m. Plane 1, plane 2, plane 3, and plane 4 were in-
technology involves the installation of a number of electrodes on stalled in the downcomer at the same level from the bottom of the
the periphery of the vessel to be imaged [32]. The main compo- bioreactor as of plane 8, plane 7, plane 6, and plane 5, respectively.
nents of the ERT are electrodes, a data acquisition system, and a All these electrodes were connected to the ITS P2000 data acquisi-
host computer for image reconstruction. tion system (Manchester, UK) via co-axial cables. The data acquisi-
As shown in Fig. 2, sixteen equally spaced stainless steel rectan- tion system (DAS) was used to send current to the electrodes fitted
gular electrodes were fitted on the periphery of the bioreactor on the internal walls of the bioreactor. The adjacent measurement
(riser and downcomer) in one array (plane) and there were four protocol was utilized in this study [26,38], in which the electrical
planes in the downcomer and four planes in the riser. Fig. 2 also current was applied between an adjacent pair of electrodes and
shows the locations and dimensions of electrodes that were used. the voltage was measured between all other pairs of adjacent elec-
Plane 5 was installed in the middle of the axial height of the lower trodes. Other combinations of adjacent electrode pairs went
packing, while plane 6 was located in the middle of the axial dis- through the same procedure until the full rotation was obtained
tance between the top of lower packing and the bottom of the on the bioreactor. The frequency of 9,600 kHz and injection current
gas redistributor. Plane 7 was installed in the middle of the axial of 1.5 mA were employed for all experiments in this study. In this
distance between the bottom of the upper packing and the top of study, a non-iterative linear back-projection algorithm [39] was
the perforated plate while plane 8 was located in the middle of employed to create the tomograms, which showed the cross-sec-
the axial height of the upper packing. The distance between plane tional distribution of the electrical conductivity of the contents
5 and plane 6 was 0.175 m, the distance between plane 6 and plane within the same measurement plane.

Fig. 2. Position and size of the tomography electrodes.


54 M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61

2.3. Experimental procedure tc


tr ¼ ð2Þ
u
The tracer response technique was used to measure the mixing
where tc is the circulation time and it is defined as the time required
time in the bioreactor. Mixing time (tm) was calculated by subtract-
for an elemental volume of liquid to circulate through the whole
ing the time at which the tracer was injected from the time when
path-length of the bioreactor. The dimensionless factor u can be
the variation in mean conductivity values for all eight tomography
calculated using Eq. (3) and its derivation can be found in Bello [41]:
planes simultaneously became ±1% of the steady-state value. In
 
fact, the conductivity values remained stable at the steady-state 1  egd Ad Abc Hbc Atc Htc
due to complete dispersion of the tracer throughout the reactor. u¼1þ þ þ ð3Þ
1  egr Ar Ar H r Ar H r
Before the injection of the tracer, air was purged into the bioreactor
using a sparger at the bottom of the riser with constant flow rate. where egd is the gas holdup in the downcomer, egr is the gas holdup
The conductivity of the air–water mixture was monitored using in the riser, Ad is the cross-sectional area of the downcomer, Ar is the
ERT system. After 50 s of supplying air, the steady state was cross-sectional area of the riser, Abc is the cross-sectional area of the
reached and consequently the conductivity of air–water mixture bottom connecting section, Atc is the cross-sectional area of the top
became stable. After reaching the steady state, the tracer was in- connecting section, Hr is the circulating liquid height in the riser, Hbc
jected in the downcomer above the liquid level. In this study, is the circulating liquid length in the bottom connecting section and
100 ml of 4% saline solution (NaCl) was used as a tracer. The vari- Htc is the circulating liquid length in the top connecting section.
ations in mean conductivity signals generated by the ERT system Substituting values of Ad, Abc, Atc, Hr, Hbc, and Htc in Eq. (3), the gen-
were monitored on the host computer. The data collection was eral form of the dimensionless factor u for the bioreactor used in
stopped when the variations in mean conductivity values taken this study becomes:
from all four planes in the downcomer and all four planes in the
1  egd
riser simultaneously became ±1% of the steady-state value [20]. u¼1þ ½0:224 þ 0:0551 ð4Þ
The air was then discontinued and water was drained out. Same
1  egr
procedure was repeated for different gas flow rates. Fig. 3 shows Values for gas holdup in this bioreactor were measured and
the change in the mean conductivity of air–water mixture with published elsewhere [42]. The superficial liquid velocity in the
time at a riser superficial gas velocity of 1.1  102 m/s following downcomer (Uld) was calculated from a mass balance on the liquid
the injection of tracer for eight ERT planes. The mixing time was phase [43,44] using Eq. (5):
30 s for the data shown in Fig. 3.
Same experimental procedure was used to find the superficial Ar
U ld ¼ U lr ð5Þ
liquid velocity in the riser as was employed for the mixing time Ad
[6,10,40]. The riser superficial liquid velocity was calculated using
where Ulr is the superficial liquid velocity in the riser, Ad is the
the ratio of the riser liquid circulation height to the riser liquid res-
cross-sectional area of the downcomer and Ar is the cross-sectional
idence time [14,41]. The following formula was used to find the
area of the riser. Fig. 4 represents changes in conductivity with time
superficial liquid velocity in the riser:
for the cross shaped sparger at superficial gas velocity of
Hr 1.1  102 m/s in the riser and liquid height of 1.63 m for plane 2
U lr ¼ ð1  egr Þ ð1Þ (located in the downcomer) following the injection of saline solu-
tr
tion. The circulation time for the tracer was measured as the time
where Hr (measured with reference to the top of the connecting between two consecutive peaks (i.e., tc = 20 s from the data shown
tube) is the circulating liquid height in the riser, tr is the liquid res- in Fig. 4). Since the conductivity value of the second peak was much
idence time in the riser and egr is the gas holdup in the riser. The lower compared to that for the first peak, the second peak cannot be
residence time of liquid in the riser (tr) was determined by the fol- observed clearly for the given scale on y-axis in Fig. 4. Thus, to
lowing formula: obtain the circulation time accurately, we found the conductivity

tm = 30 s
Conductivity (mS/cm)

Time (s)

Fig. 3. Measurement of the mixing time at Ugr = 1.1  102 m/s for cross shaped Fig. 4. Measurement of the circulation time at Ugr = 1.1  102 m/s using ERT plane
sparger with liquid height = 1.63 m. 2 (located in downcomer) for the cross shaped sparger with liquid height = 1.63 m.
M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61 55

values at these two peaks by zooming in to get a close-up view of in the bioreactor, thus the region with higher conductivity depicts
the peaks using Excel. the higher concentration of saline solution in the bioreactor
[20,38,45]. The blue and red colors in Fig. 5 represent the lowest
and highest conductivity values, respectively.
3. Results and discussion It can be seen from Fig. 5 that the color of the tomographic
images of planes 1, 2, and 3 located in the downcomer partially
Time series of tomographic images that were recorded from 8 changed to red (higher conductivity value) at t = 2 s since the tracer
measurement planes of the ERT system (Fig. 2) are shown in was injected near plane 1. The injected tracer reached plane 4 at
Fig. 5. These images were recorded after the injection of the tracer t = 3 s and the color of the tomogram of plane 4 located in the
(i.e., 100 ml of 4% saline solution) from the top of the downcomer downcomer changed to the red color. These changes to the red col-
near plane 1 at a constant riser superficial gas velocity (i.e., or were due to the increase in conductivity of the fluid in the
Ugr = 3.3  102 m/s) using the cross shaped sparger with a con- downcomer. At t = 6 s, tomographic images recorded in planes 1,
stant liquid height (i.e., H = 1.63 m). It is a known fact that the con- 2, 3, and 4 changed to green, which represents decrease in conduc-
ductivity of saline solution is higher than the conductivity of water tivity of the solution. This can be due to the dispersion of saline

Fig. 5. Time series of tomograms collected from 8 measurement planes of ERT system (plane 1, plane 2, plane 3 and plane 4 located on downcomer and plane 5, plane 6, plane
7 and plane 8 located on riser; see Fig. 2) following the injection of saline solution into the air–water mixture using the cross shaped sparger with Ugr = 3.3  102 m/s and
liquid height = 1.63 m (without packing). Blue and red colors represent lowest and highest conductivity values. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)
56 M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61

Fig. 5. (continued)

solution in the downcomer. After passing through the downcomer, ficial gas velocity of 3.3  102 m/s using the cross shaped sparger
the saline solution entered into the riser using bottom connecting at a constant liquid height of 1.63 m in the bioreactor. The green
tube. However, since the gas holdup in the riser is much higher, the color in the downcomer at t = 0 s was due to the lower gas holdup
red regions on the tomograms of planes 5, 6, 7, and 8, located in the in the downcomer than that in the riser. Thus, the blue region was
riser from the bottom to top, were not observed. This variation in not observed in the downcomer. At t = 0 s, the saline solution was
conductivity indicates the presence of saline solution in the biore- injected into the top of the downcomer and the color in the down-
actor. The ERT images show that the mixing time under these cir- comer turned red at t = 2 s, which represents highest conductivity.
cumstances was about 17 s. At t = 10 s, the color in the downcomer changed to green, which
Fig. 6 demonstrates the vertical slice of the tomographic images represents lower conductivity. This can be due to the fact that
and the 3D tomograms of air–water mixture in the riser and down- the saline solution had left the downcomer and entered into the ri-
comer of the bioreactor. It must be mentioned that these vertical ser through the bottom connecting tube. The blue color (lower con-
slices represent the vertical area of the riser from planes 5 to 8 ductivity due to air bubbles) in the riser was slightly changed to
and the vertical area of the downcomer from planes 1 to 4. Simi- green color in Fig. 6 after 10 s, which indicates the presence of sal-
larly, images were recorded following the injection of the saline ine solution in the riser. The tracer was uniformly dispersed
solution from the top of the downcomer at a constant riser super- throughout the bioreactor after 17 s. The blue colors observed in
M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61 57

Fig. 6. Vertical slice tomographic images and 3D tomograms of the air–water mixture in riser and downcomer of the bioreactor following the injection of the saline solution
at a constant riser superficial gas velocity of 3.3  102 m/s using the cross shaped sparger at a constant liquid height of 1.63 m in the bioreactor (without packing). Blue and
red colors represent lowest and highest conductivity values. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this
article.)

Fig. 7. Effect of gas flow rate and liquid height on mixing time using cross shaped Fig. 8. Effect of sparger configuration on the mixing time with a constant liquid
sparger (without packing). height of 1.63 m in the bioreactor (without packing).
58 M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61

Fig. 9. Effect of the packing in the riser (packing height = 0.234 m) on the mixing
Fig. 12. Effect of the gas flow rate and liquid height on the superficial liquid velocity
time with a constant liquid height of 1.63 m using the cross shaped sparger.
in the riser using the cross shaped sparger.

creased with increasing superficial gas velocity in the riser.


Increasing the superficial gas velocity in the riser tends to drive
an increase in the superficial liquid velocity in the riser, which re-
duces the mixing time in the bioreactor. Fig. 7 also shows that mix-
ing time in the riser increased with decreasing liquid height in the
bioreactor. Bentifraouine et al. [35] reported that the superficial li-
quid velocity decreased when the liquid height decreased due to a
decrease in the difference between the hydrostatic pressures in the
riser and the downcomer. Liu et al. [40] also found that with
decreasing the liquid height in the bioreactor, the flowing resis-
tance in the upper connecting tube increased, which resulted in
the reduction of the liquid superficial velocity. This decrease in
the liquid superficial velocity led to a higher mixing time in the
bioreactor.
Profiles for the mixing time of the cross shaped and the circular
Fig. 10. Effect of the gas redistributor installed in the riser on the mixing time with sparger (Fig. 1) using different riser superficial gas velocities with
a constant liquid height of 1.63 m using the cross shaped sparger. constant liquid height in the bioreactor are shown in Fig. 8. These
results demonstrate that at a constant riser superficial gas velocity,
the mixing time was increased when a circular sparger was used
over a cross shaped sparger of same diameter and with same num-
ber and size of holes. The increase in mixing time was due to the
presence of more resistance in the flow. The cross shaped sparger
consists of six circular tubes of small diameter. The liquid can eas-
ily pass through the space between these circular tubes. While a
circular sparger consists of a circular plate which offers less space
between the riser and the sparger for the liquid to pass and hence
offers more resistance to flow path of the liquid. In fact, the

Fig. 11. Effect of a gas redistributor installed between two beds of packing in the
riser (height of each packing = 0.234 m) on the mixing time with a constant liquid
height of 1.63 m using the cross shaped sparger.

planes 5–8 (Fig. 6) were due to the gas holdup in the riser. In fact,
the regions with blue color (regions of lower conductivity) repre-
sent the higher gas holdup in those areas.
Fig. 7 shows the effects of superficial gas velocity and liquid
height on the mixing time in the bioreactor using the cross shaped
sparger configuration without packing and gas redistributor. It can Fig. 13. Effect of the sparger configuration on the superficial liquid velocity in the
be seen from this figure that mixing time in the bioreactor de- riser with a constant liquid height of 1.63 m in the bioreactor.
M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61 59

Fig. 16. Effect of using a gas redistributor between two beds of packing in the riser
(height of each packing = 0.234 m) on the riser superficial liquid velocity with a
Fig. 14. Effect of the packing installed in the riser (packing height = 0.234 m) on the constant liquid height of 1.63 m using a cross shaped sparger configuration.
superficial liquid velocity in the riser with a constant liquid height of 1.63 m using
the cross shaped sparger configuration.
tance in the flow path of the liquid. When an internal gas redistrib-
presence of the cross shaped sparger reduced the cross sectional utor is installed in the riser of the bioreactor, it enhances the flow
area available for flow by about 17% while the presence of the cir- resistance, which in turn decreases the liquid superficial velocity.
cular sparger reduced the cross sectional area available for flow by This decrease in liquid superficial velocity delays the time required
about 65%. It is a known fact that the increase in resistance results for mixing in the bioreactor. This delay results in an increase in
in a lower liquid superficial velocity [10,46,47], which in turn in- mixing time in the bioreactor.
creases the mixing time in the bioreactor. The effect of the gas redistributor on the mixing time in the bio-
The effect of packing in the riser on the mixing time was exam- reactor in the presence of the packing in the riser is summarized in
ined with the cross shaped sparger and constant liquid height of Fig. 11 (with a constant liquid height and using the cross shaped
1.63 m. It was observed from results in Fig. 9 that at a constant ri- sparger). Clearly, the results show an increase in mixing time when
ser superficial gas velocity, the mixing time was increased when a the gas redistributor was installed between the two rolls of pack-
packing was installed in the riser. The installation of packing in the ing at a constant superficial gas velocity in the riser. This increase
riser offers more resistance to the flow path of the liquid in the ri- in mixing time was due to the presence of more resistance in the
ser which results in the decrease in the liquid superficial velocity form of packing and a gas redistributor in the flow path of the li-
[24,36]. This decrease in liquid superficial velocity in turn increases quid in the riser. The presence of this resistance in the liquid flow
the mixing time in the bioreactor. However, it is still important to contributes to lower liquid superficial velocity [10,24,36,46,47],
install packing in a bioreactor for processes where packing is used which in turn increases the mixing time in the bioreactor.
as a support for immobilized microorganisms or enzymes [48]. Fig. 12 represents the effects of the superficial gas velocity and
Mixing time in the bioreactor was measured at different super- liquid height on the superficial liquid velocity in the riser using the
ficial gas velocities in the presence and absence of the gas redis- cross shaped sparger. It can be seen that the superficial liquid
tributor (without packing). It can be noticed from Fig. 10 that at velocity in the riser was increased with increasing the superficial
a constant riser superficial gas velocity, the mixing time in the bio- gas velocity in the riser. The superficial liquid velocity in the biore-
reactor equipped with an internal gas redistributor was higher actor depends on the hydrostatic pressure difference between the
than that in the bioreactor without the gas redistributor. Some riser and the downcomer of the bioreactor. When the gas flow rate
investigators studied the liquid velocity in the fluidized bed and in- is increased in the riser, the hydrostatic pressure difference be-
verse fluidized bed external loop airlift bioreactors [10,46,47]. They tween the riser and the downcomer increases, which in turn in-
observed that the liquid velocity in the bioreactor with fluidized creases the superficial liquid velocity in the riser. In the earlier
bed was lower than that without any fluidized bed due to the resis- investigations conducted by Han et al. [10] and Kemblowski
et al. [16], Loh and Liu [6], Mohanty et al. [25], they all agreed on
that when the gas flow rate in the bioreactor is increased, the
superficial liquid velocity also increases. It can also be seen from
Fig. 12 that the superficial liquid velocity in the riser was decreased
with decreasing liquid height in the bioreactor. At constant super-
ficial gas velocity, when the liquid height in the bioreactor is de-
creased, the resistance in the path of the liquid flowing in the
upper connecting tube of the bioreactor is increased. This increased
resistance, in turn, decreases the liquid superficial velocity in the
bioreactor [33–35,40].
At different superficial gas velocities, Fig. 13 illustrates the ef-
fect of gas sparger type on the superficial liquid velocity in the riser
at a constant liquid height in the bioreactor. It was noticed that the
riser superficial liquid velocity was decreased at constant riser
superficial gas velocity when a circular sparger was used over a
cross shaped sparger of same diameter and with same number
Fig. 15. Effect of the gas redistributor installed in the riser on the superficial liquid
and size of holes on it. This decrease in superficial liquid velocity
velocity in the riser with a constant liquid height of 1.63 m for the cross shaped in the riser is due to the fact that the circular sparger consists of
sparger configuration. a circular solid plate, which offers more resistance than its
60 M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61

counterpart (cross shaped sparger). A cross shaped sparger consists external loop airlift bioreactor equipped with an internal gas redis-
of six small diameter circular tubes. These small diameter tubes tributor between two beds of packing in the riser. To characterize
have much more space between each other for the water to pass the mixing performance of this bioreactor, ERT data were collected
easily without much resistance. In fact, the presence of the cross after the injection of the tracer to measure the mixing time and the
shaped sparger reduced the cross sectional area available for flow superficial liquid velocity in the bioreactor. To measure the distri-
by about 17% while the presence of the circular sparger reduced bution of the tracer within the riser and the downcomer, the adja-
the cross sectional area available for flow by about 65%. Increasing cent measurement strategy and linear back-projection algorithm
resistance in the flow path of the liquid contributes to lower super- were employed. The experimental results demonstrated that the
ficial liquid velocity in the bioreactor [10,46,47]. Moreover, the mixing time in the bioreactor decreased with an increase in the
holes in the cross shaped sparger are located throughout the cross superficial gas velocity in the riser. When the gas flow rate was in-
section of the sparger; however, in a circular sparger, they are lo- creased in the riser, the hydrostatic pressure difference between
cated only at the periphery of the plate. Therefore, the cross shaped the riser and the downcomer was increased, which in turn in-
sparger can disperse air more evenly across column of the riser creased the superficial liquid velocity in the riser. This increase in
than that of the circular sparger. the superficial liquid velocity in the riser decreased the mixing
Fig. 14 demonstrates the effect of packing on the superficial li- time in the bioreactor. The ERT results also revealed that the mix-
quid velocity in the riser using the cross shaped sparger with a con- ing time was increased and the superficial liquid velocity was de-
stant liquid height of 1.63 m in the bioreactor. It was observed that creased by reducing the liquid height in the bioreactor, installing
the superficial liquid velocity in the riser decreased when the pack- the internal gas redistributor and/or installation of packing in the
ing was installed in the riser at a constant riser superficial gas riser, and by using the circular sparger. This was due to the in-
velocity. The installation of packing in the riser reduces the cross crease in the resistance in the liquid flow path.
section area available for flow and offers more resistance to the
flow path of liquid, which results in the decrease in liquid superfi- Acknowledgments
cial velocity [24,36].
Fig. 15 is a typical plot showing the effect of the superficial gas The financial supports of the Natural Science and Engineering
velocity on the superficial liquid velocity in the riser, with and Research Council of Canada (NSERC), the Agriculture and Agri-Food
without the gas redistributor. It can be seen that the superficial li- Canada through the Agricultural Bioproducts Innovation Network
quid velocity in the riser of the bioreactor fitted with the internal (ABIN), and the Ryerson University are gratefully acknowledged.
gas redistributor at a constant superficial gas velocity was lower
than that measured for the bioreactor without internal gas redis- References
tributor. The lower superficial liquid velocity in the riser was due
to the presence of internal gas redisitributor, which reduced the [1] Z. Al-Qodah, W. Lafi, Modelling of antibiotics production in magneto three-
cross-sectional area available for flow and increased flow resis- phase airlift fermenter, Biochem. Eng. J. 7 (2001) 7–16.
[2] S. Sarkar, K. Mohanty, B.C. Meikap, Hydrodynamic modeling of a novel multi-
tance in the bioreactor equipped with the internal gas redistributor stage gas–liquid external loop airlift reactor, Chem. Eng. J. 145 (2008) 69–77.
[10,46,47]. [3] T. Zhang, J. Wang, T. Wang, J. Lin, Y. Jin, Effect of internal on the hydrodynamics
Fig. 16 illustrates the effect of the internal gas redistributor in- in external-loop airlift reactors, Chem. Eng. Process. 44 (2005) 81–87.
[4] F. Benyahia, L. Jones, Scale effects on hydrodynamic and mass transfer
stalled in the riser between the two rolls of packing on the super- characteristics of external loop airlift reactors, J. Chem. Technol. Biotechnol.
ficial liquid velocity in the riser with a constant liquid height of 69 (3) (1997) 301–308.
1.63 m for the cross shaped sparger. It was noticed that the riser [5] J. Klein, J. Maia, A.A. Vicente, L. Domingues, J.A. Teixeira, M. Jurascik,
Relationships between hydrodynamics and rheology of flocculating yeast
superficial liquid velocity decreased, at constant riser superficial suspensions in a high-cell-density airlift bioreactor, Biotechnol. Bioeng. 89 (4)
gas velocity, when an internal gas redistributor was installed be- (2005) 393–399.
tween two rolls of packing in the riser. A bioreactor with an inter- [6] K. Loh, J. Liu, External loop inversed fluidized bed airlift bioreactor (EIFBAB) for
treating high strength phenolic wastewater, Chem. Eng. Sci. 56 (2001) 6171–
nal gas redistributor between two beds of packing provides less 6176.
cross sectional area available for flow and offers much more resis- [7] K. Mohanty, D. Das, M.N. Biswas, Treatment of phenolic wastewater in a novel
tance in the flow path of the liquid than that without the internal multi-stage external loop airlift reactor using activated carbon, Sep. Purif.
Technol. 58 (2008) 311–319.
gas redistributor and packing. This increase in resistance in the
[8] M. Gavrilescu, R.Z. Tudose, Mixing studies in external-loop airlift reactors,
flow path of the liquid contributes to lower liquid superficial veloc- Chem. Eng. J. 66 (1997) 97–104.
ity in the bioreactor [10,36,46,47]. [9] F. Gumery, F. Ein-Mozaffari, Y. Dahman, Characteristics of local flow dynamics
In general, the superficial liquid velocity in the downcomer was and macro-mixing in airlift column reactors for reliable design and scale-up,
Int. J. Chem. Reac. Eng. 7 (2009). Review R4.
higher than the superficial liquid velocity in the riser because of [10] S.J. Han, R.B.H. Tan, K.C. Loh, Hydrodynamic behaviour in a new gas liquid solid
the smaller diameter of the downcomer. The range of superficial li- inverse fluidization airlift bioreactor, Trans. IChemE Part C 78 (2000) 207–215.
quid velocity found in the downcomer was 0.103–0.785 m/s. [11] P.M. Kilonzo, A. Margaritis, M.A. Bergougnou, Hydrodynamics and mass
transfer characteristics in an inverse internal loop airlift-driven fibrous-bed
Experimental results of this study indicated that the gas redistrib- bioreactor, Chem. Eng. J. 157 (2010) 146–160.
utor significantly decreased the superficial liquid velocity and in- [12] P.M. Kilonzo, A. Margaritis, M.A. Bergougnou, Hydrodynamic characteristics in
creased the mixing time. However, higher gas holdup can be an inverse internal-loop airlift-driven fibrous-bed bioreactor, Chem. Eng. Sci.
65 (2010) 692–707.
achieved with the installation of the gas redistributor between [13] W. Tang, L. Fan, Steady state phenol degradation in a draft-tube, gas-liquid-
two beds of packing in the bioreactor [42]. Higher gas holdup solid fluidized-bed bioreactor, AIChE J. 33 (2) (1987) 239–249.
improves the mass transfer inside the reactor by increasing the [14] R.A. Bello, C.W. Robinson, M. Moo-Young, Liquid circulation and mixing
characteristics of airlift contactors, Can. J. Chem. Eng. 62 (1984) 573–577.
interfacial mass transfer area. The installation of the gas redistrib- [15] H. Dhaouadi, S. Poncin, J.M. Hornut, G. Wild, Solid effects on hydrodynamics
utor also enhances the uniform distribution of the smaller gas bub- and heat transfer in an external loop airlift reactor, Chem. Eng. Sci. 61 (2006)
bles in the riser. 1300–1311.
[16] Z. Kemblowski, J. Przywarski, A. Diab, An average gas holdup and liquid
circulation velocity in airlift reactors with external loop, Chem. Eng. Sci. 48
(1993) 4023–4035.
4. Conclusion [17] K. Mohanty, D. Das, M.N. Biswas, Mass transfer characteristics of a novel multi-
stage external loop airlift reactor, Chem. Eng. J. 133 (2007) 257–264.
[18] K. Shimizu, S. Takada, T. Takahashi, Y. Kawase, Phenomenological simulation
In this study, electrical resistance tomography (ERT) was suc- model for gas holdups and volumetric mass transfer coefficients in external-
cessfully employed to explore the hydrodynamics of mixing in an loop airlift reactors, Chem. Eng. J. 84 (2001) 599–603.
M. Hamood-ur-Rehman et al. / Chemical Engineering Journal 213 (2012) 50–61 61

[19] F. Yazdian, S.A. Shojaosadati, M. Nosrati, M.P. Hajiabbas, E. Vasheghani- [35] C. Bentifraouine, C. Xuereb, J. Riba, Effect of gas liquid separator and liquid
Farahani, Investigation of gas properties, design, and operational parameters height on the global hydrodynamic parameters of an external loop airlift
on hydrodynamic characteristics, mass transfer, and biomass production from contactor, Chem. Eng. J. 66 (1997) 91–95.
natural gas in an external airlift loop bioreactor, Chem. Eng. Sci. 64 (2009) [36] H. Nikakhtari, G.A. Hill, Hydrodynamic and oxygen mass transfer in an
2455–2465. external loop airlift bioreactor with a packed bed, Biochem. Eng. J. 27 (2005)
[20] F. Gumery, F. Ein-Mozaffari, Y. Dahman, Macromixing hydrodynamic study in 138–145.
draft-tube airlift reactors using electrical resistance tomography, Bioprocess. [37] Z.F. Zhao, M. Mehrab, F. Ein-Mozaffari, Mixing time in an agitated multi-lamp
Biosyst. Eng. 34 (2011) 135–144. cylindrical photoreactor using electrical resistance tomography, J. Chem.
[21] B. Jin, Q. Yu, Q.X. Yan, J. Leeuwen, Characterization and improvement of Technol. Biotechnol. 83 (2008) 1676–1688.
oxygen transfer in pilot plant external air-lift bioreactor for mycelia biomass [38] L. Pakzad, F. Ein-Mozaffari, P. Chan, Using electrical resistance tomography and
production, World J. Microbiol. Biotechnol. 17 (2001) 265–272. computational fluid dynamics modeling to study the formation of cavern in
[22] Y. Kawase, N. Omori, M. Tsujimura, Liquid-phase mixing in external-loop airlift the mixing of pseudoplastic fluids possessing yield stress, Chem. Eng. Sci. 63
bioreactors, J. Chem. Technol. Biotechnol. 61 (1994) 49–55. (2008) 2508–2522.
[23] Z. Xu, J. Yu, Hydrodynamics and mass transfer in a novel multi-airlifting [39] A. Madupu, A. Mazumdar, Z. Jinsong, D. Roelant, R. Srivastava, Electrical
membrane bioreactor, Chem. Eng. Sci. 63 (2008) 1941–1949. resistance tomography for real-time mapping of the solid-liquid interface in
[24] A.X. Meng, G.A. Hill, A.K. Dalai, Hydrodynamic characteristics in an external tanks containing optically opaque fluids, Proc. SPIE – Int. Soc. Opt. Eng. 5674
loop airlift bioreactor containing a spinning sparger and a packed bed, Ind. Eng. (2005) 36–46.
Chem. Res. 41 (2002) 2124–2128. [40] M. Liu, T. Zhang, T. Wang, W. Yu, J. Wang, Experimental study and modeling on
[25] K. Mohanty, D. Das, M.N. Biswas, Hydrodynamics of a novel multi-stage liquid dispersion in external-loop airlift slurry reactors, Chem. Eng. J. 139
external loop airlift reactor, Chem. Eng. Sci. 61 (2006) 4617–4624. (2008) 523–531.
[26] S. Hosseini, D. Patel, F. Ein-Mozaffari, M. Mehrvar, Study of solid–liquid mixing [41] R.A. Bello, A characterization study of airlift contactors for application to
in agitated tanks through electrical resistance tomography, Chem. Eng. Sci. 65 fermentations. PhD Thesis, Department of Chemical Engineering, University of
(2010) 1374–1384. Waterloo, Canada, 1981.
[27] L.K. Ishkintana, C.P.J. Bennington, Gas holdup in pulp fibre suspensions: gas [42] M. Hamood-ur-Rehman, F. Ein-Mozaffari, Y. Dahman, Dynamic and local gas
voidage profiles in a batch-operated sparged tower, Chem. Eng. Sci. 65 (2010) holdup studies in external loop recirculating airlift reactor with two rolls of
2569–2578. fibreglass packing using electrical resistance tomography, J. Chem. Technol.
[28] H. Jin, M. Wang, R.A. Williams, The effect of sparger geometry on gas bubble Biotechnol., in press doi: http://dx.doi.org/10.1002/jctb.3917.
flow behaviors using electrical resistance tomography, Chin. J. Chem. Eng. 14 [43] A.H. Essadki, M. Bennajah, B. Gourich, C. Vial, M. Azzi, H. Delmas,
(1) (2006) 127–131. Electrocoagulation/electroflotation in an external-loop airlift reactor—
[29] H. Jin, S. Yang, M. Wang, R.A. Williams, Measurement of gas holdup profiles in application to the decolorization of textile dye wastewater: a case study,
a gas liquid cocurrent bubble column using electrical resistance tomography, Chem. Eng. Process. 47 (2008) 1211–1223.
Flow Meas. Instrum. 18 (2007) 191–196. [44] M. Popovic, C.W. Robinson, Mixing characteristics of external-loop airlifts:
[30] H. Jin, M. Wang, R.A. Williams, Measurement of gas holdup profiles in a gas non-newtonian systems, Chem. Eng. Sci. 48 (8) (1993) 1405–1413.
liquid cocurrent bubble column using electrical resistance tomography, Chem. [45] G.T. Bolton, C.W. Hooper, R. Mann, E.H. Stitt, Flow distribution and velocity
Eng. J. 130 (2007) 179–185. measurement in a radial flow fixed bed reactor using electrical resistance
[31] P. Tahvildarian, H. Ng, M. D’Amato, S. Drappel, F. Ein-Mozaffari, S.R. Upreti, tomography, Chem. Eng. Sci. 59 (2004) 1989–1997.
Using electrical resistance tomography images to characterize the mixing of [46] Y.X. Guo, M.N. Rather, H.C. Ti, Hydrodynamics and mass transfer studies in a
micron-sized polymeric particles in a slurry reactor, Chem. Eng. J. 172 (2011) novel external loop airlift reactor, Chem. Eng. J. 67 (1997) 205–214.
517–525. [47] K. Loh, S. Ranganath, External-loop fluidized bed airlift bioreactor (EFBAB) for
[32] R.A. Williams, M.S. Beck, Process Tomography: Principles, Techniques and the cometabolic biotransformation of 4-chlorophenol (4-cp) in the presence of
Applications, Butterworth-Heinemann, Boston, 1995. phenol, Chem. Eng. Sci. 60 (2005) 6313–6319.
[33] W.A. Al-Masry, Effect of liquid volume in the gas-separator on the [48] J. Horiuchi, K. Tabata, T. Kanno, M. Kobayashi, Continuous acetic acid
hydrodynamics of airlift reactors, J. Chem. Technol. Biotechnol. 74 (1999) production by a packed bed bioreactor employing charcoal pellets derived
931–936. from waste mushroom medium, J. Biosci. Bioeng. 89 (2) (2000) 126–130.
[34] W.A. A1-Masry, Influence of gas-separator and scale up on the hydrodynamics
of external loop circulating bubble columns, Chem. Eng. Res. Des. 82 (A3)
(2004) 381–389.

You might also like