Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

148 Ind. Eng. Chem. Res.

2009, 48, 148–158

Uniform Flow in Bubble Columns


R. F. Mudde,* W. K. Harteveld,† and H. E. A. van den Akker
Kramers Laboratorium Voor Fysische Technologie and J.M. Burgers Center for Fluid Mechanics, Delft
UniVersity of Technology, Pr. Bernhardlaan 6, 2628 BW Delft, The Netherlands

Uniform bubbly flow in a 15 cm bubble column is investigated. We use a special needle sparger consisting
of 559 separate needles, uniformly distributed over the bottom. With this sparger, we can ensure that all
bubbles generated are of the same size and that the bubble injection is very uniform over the entire bottom
of the column. Detailed experiments are reported, using optical glass fibers to measure the local gas fraction
and bubble size and velocity and using laser Doppler anemometry to measure the liquid axial velocity field.
We find that the homogeneous flow regime extends up to a gas fraction of 55% well beyond the predictions
of theory. The superficial gas velocity at which the homogeneous regime looses its stability depends on the
water quality: fresh water looses its stability much earlier than old water. However, the gas fraction as a
function of the superficial gas velocity is in the homogeneous regime independent of the water quality. The
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

overall gas fraction can be described by a Richardson and Zaki type of relation or by the proposal by Garnier
et al. We have indications that, at the point of instability, the bubble size has increased to a critical value at
Downloaded via NATL CHENG KUNG UNIV on January 13, 2023 at 03:06:46 (UTC).

which the lift force reverses sign. This causes the radial gas fraction to change from flat with a small wall
peaking to core peaking provoking the instability as suggested by Lucas et al. [Chem. Eng. Technol. 2005,
60, 3609]. Alternatively, at higher gas fractions, the swarm gets denser and the bubble wakes get suppressed.
According to Fox and co-workers [Chem. Eng. Sci. 2007, 62, 3159], this causes the flow to lose its stability.

1. Introduction size. Furthermore, the critical gas fraction peaks at db ≈ 3 mm.


The predicted peak values range from as small as 10% to 100%
In bubble columns, two basic flow regimes are possible: the (i.e., always stable), depending on the closure relations used in
homogeneous and the heterogeneous regime (see, e.g., refs 3–6). the hydrodynamic description. In the same paper, also the
The latter prevails at the higher gas flow rates and is generally experimental data from the literature are collected. For air-water
characterized by transient flow, a net circulation, and a wide columns at ambient conditions, the maximum critical gas
bubble size distribution. Bubble columns operate in most cases fraction reported is 36.5%. However, the majority of data are
in the heterogeneous regime. The homogeneous regime is found around a value of 20-25% with critical superficial velocities
at low superficial gas velocities. In this regime, the bubble size of about 3-4 cm/s.
has a narrow distribution and the liquid motion is restricted to León-Becerril et al.13,14 report that, according to their stability
small scales induced by the passing bubbles. At some point, analysis, the homogeneous flow becomes unstable at R ≈ 30%
the homogeneous regime becomes unstable and a transition sets for spherical bubbles. This critical gas fraction reduces to 10%
in to the heterogeneous one. Numerous papers have appeared
to 15% for ellipsoidally shaped bubbles. Lucas and co-workers1,15
that deal with one or more of the aspects of this transition from
investigated the role of the lift force and concluded that the
homogeneous to heterogeneous.
influence of the lift force on stability is much higher than that
One way to study this transition is by constructing a gas of the turbulent dispersion force. Moreover, for small bubbles
fraction (R) vs superficial gas velocity (Usup) plot from that have a positive lift coefficient, the lift force is stabilizing,
experimental data,4,7–11 and find the inflection point in the plot. whereas for the larger bubbles, that have a negative lift
Here the transition to the heterogeneous regime sets in. coefficient, this force is destabilizing. Bhole and Joshi16 collected
Linear stability analysis is also frequently employed to a lot of literature data for pure systems as well as for systems
investigated the stability of the homogeneous regime, see e.g., with additives. A wide range of critical gas fractions was
1, 2, and 12–16. An extensive review is presented by Joshi et reported. For water, the maximum critical gas fraction is as high
al.17 In their review, they discuss the one-dimensional linear as 45.8%.11 The theoretical result of the work of Bhole and
stability of the homogeneous flow in various dispersed systems, Joshi gave a good prediction of the transition from homogeous
like fluidized beds and bubble columns. The authors present a to heterogeneous flow for a wide range of systems. Joshi12 also
relation for the maximum gas fraction in the bubble column in conducted a linear stability analysis for a 2-dimensional
the homogeneous regime. This maximum is reached when the air-water bubble column. From this work, it follows that the
gas fraction versus superficial gas velocity is at its maximum. critical gas fraction follows from a simple relation: Rcrit ) 1/(n
However, the system may become unstable before this point is + 1) with n bing the Richardson and Zaki coefficient. According
reached. The stability analysis provides the critical gas fraction to the authors, a typical value for n is 1.4, leading to a critical
at which small disturbances of the homogeneous regime get gas fraction of 42%. Finally, Monahan and Fox2 inspected
amplified. Various parameters influence the critical gas fraction instability from both horizontal and vertical modes in the bubbly
calculated from the stability analysis. For bubble sizes above 3 flow. Different from e.g. Lucas et al. they find that the instability
mm, the critical gas fraction is found to decrease with increasing is associated with the bubble induced turbulence production.
* To whom correspondence should be addressed. E-mail address: A lot of work on the transition between homogeneous and
r.f.mudde@tudelft.nl. heterogeneous flow in the bubble columns has been done by

Present address: Shell Global Solutions, Amsterdam, The Netherlands. Drahos and co-workers. Zahradnik9 already in 1979 investigated
10.1021/ie8000748 CCC: $40.75  2009 American Chemical Society
Published on Web 09/06/2008
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 149
23
this problem and found that for a perforated plate the critical et al. measured, using a hot wire, the liquid velocity in a bubble
gas fraction was higher for smaller hole diameter. They reached column of 8.0 cm diameter equipped with a needle sparger
a value of 39% for a perforated plate with a hole size do ) 0.5 consisting of 271 needles. Very uniform flows could be obtained.
mm. Locket and Kirkpatrick8 also experimentally investigated They also measured the local gas fraction using a double optical
the stability of bubble flows. They operated in a countercurrent probe. A small cocurrent liquid flow was used in the experi-
mode such that the swarms could be kept stationary. This way ments; hence, this is not literally a bubble column. They found
they managed, in a pipe of 76 mm diameter, to obtain a critical very flat gas fraction profiles for gas fractions as high as 40%.
gas fraction of 66%. However, this is not directly comparable They also gave a scaling law for the bubble rise velocity in the
to the homogeneous bubble column, as now there is a distinct swarms. It scales with R1/3, like sedimenting particles in the
liquid velocity profile close to the wall creating a lift force on dense regime. Harteveld and co-workers24 also used a needle
the bubbles. Furthermore, the liquid flow is turbulent and thus sparger to achieve very uniform inlet conditions. They reported
also a turbulent dispersion force is action on the bubbles. experiments using the laser Doppler anemometry (LDA) and
Drahos et al.18 studied pressure fluctuations in a bubble optical glassfibers and showed very uniform flow at gas fractions
column. They report a maximum hold up of 30% in their of 10%. In a later paper25 these authors reported in the same
equipment. Zahradnik et al.4 found in a 14 cm diameter bubble column very uniform flow up to a gas fraction of 25%. Kulkarni
column from residence time distribution measurements and from and Joshi26 performed 2D laser Doppler anemometry measure-
the gas fraction that for a perforated plate the homogeneous ments in a 15 cm diameter column using two different spargers:
flow could easily be obtained for superficial gas velocities up a single-point and a multipoint one. The superficial gas velocity
to 4 cm/s. They also reported that the stability of the homoge- was kept constant at 2.0 cm/s during all experiments. The results
neous flow could be significantly enhanced by adding surface were compared to CFD (computational fluid dynamics) simula-
active solutes to the liquid. They used a 14 cm diameter bubble tions of the flow. The flow showed large scale circulation already
column with different perforated plates for sparging the gas. at a gas fraction of about 7%. More importantly for the present
For water, they found a critical gas fraction of 32%, but this paper, the authors made an attempt to incorporate a real sparger
could easily be increased to 63% by adding 0.5% of ethanol. in the CFD simulations. They conclude that “still [a] lot of scope
Also for electrolytes, the critical value was higher, e.g. they exists for the development of generalized codes, possibly by
report 38% for KI. understanding the physics of the flow”. The present paper aims
Ruzicka and co-workers11 made an extensive study of the at providing detailed, reliable data on the hydrodynamics in the
transition from homogeneous to heterogeneous flow. They homogeneous regime. CFD simulations should be able to capture
provided an extensive literature review on the topic and gave a this type of flow relatively easy as the modeling of phenomena
model description for the critical prediction. From this, it such as turbulence does not play a crucial role here. Monahan
followed that the maximum gas fraction for the homogeneous et al.27 performed CFD simulations of a bubble column and
regime has a value of 55%. However, this was found for shallow paid special attention to the homogeneous regime. They showed
columns with a height of diameter ratio of 3. In a subsequent that indeed the simulations can reproduce the homogeneous
paper in 2001,19 the same group of researchers investigated the regime at relatively high superficial gas velocities.
effects of the diameter and of the mixture height on the stability. From the above, it does not become clear where the border
They reported that both an increase in the diameter and the between the two regimes is exactly located. In general, it is
height had a destabilizing effect on the homogeneous regime. found that the homogeneous flow goes through a transition
In most cases, the critical gas fraction was 30% or below. Also regime toward the heterogeneous regime at gas fraction between
this research group reported20 the effect of the liquid viscosity 15-25%. Theoretical work28 suggest that at gas fractions
on the stability: for µl ) 1-3 mPa s, no big difference was beyond 30% the homogeneous flow is inherently unstable and,
found, but for the range of 3-22 mPa s, the stability decreased in smaller diameters, a transition to slug flow sets in. However,
with increasing viscosity. Finally, in a recent paper,21 the effect stability analysis also resulted in higher predections.
of surfactants was investigated. It was found that addition of Although for industrial practice the heterogeneous regime is
e.g. CaCl2 at low concentrations stabilized the homogeneous the most important one, proper experimental data on the
regime. However, for larger concentrations a destabilizing was homogeneous regime are relevant. Nowadays, bubbly flows are
observed. In this paper, the authors conclude that the important also studied via CFD. In particular, for bubble columns, CFD
problem of the prevailing flow regime in bubble column reactors simulations of the heterogeneous regime have been reported.27,29–32
has been paid only very little attention. The flow regimes are A good test of the codes and models used is to see if a code
in most studies not specified at all, or mentioned only margin- can predict the right homogeneous flow, when the gas is
ally. If mentioned, so usually vaguely, based on the authors introduced uniformly over the bottom of the bubble column. It
intuition and experience. If specified, so usually based only on is an open question up to which gas fractions the flow can be
the visual inspection of experimental data, mainly on the shape homogeneous and whether or not CFD can predict this.
of the e(q) graph. Moreover, the onset of the instability triggering the change from
Virtually all the above-mentioned papers used rather course homogeneous to heterogeneous flow is not completely understood.
experimental techniques. The number of papers on the homo- The number of detailed experiments with well-defined and
geneous bubble regime that reported detailed hydrodynamic controlled inlet conditions is rather limited. In almost all cases,
experiments have been rather limited. Yao et al.22 preformed porous plates or perforated plates have been used. With these
experiments using a conductivity probe for the bubble size and spargers, the inlet conditions are not accurately controlled, nor
velocity and for the gas fraction and an ultrasonic Doppler are they known with high accuracy. Therefore, the present paper
anemometer also for the bubble velocity (i.e., for 2 components revisits the homogeneous bubbly flow. We aim at providing
of the velocity vector). Furthermore, they employed a hot wire detailed data of the gas fraction distribution (using glass fibers)
anemometer for the liquid velocity. They report that a large and liquid velocity (using LDA) that can, e.g., be used to test
scale liquid circulation, characteristic for the heterogeneous CFD simulations of the homogeneous regime. Our sparger is
regime, sets in above a superficial gas velocity of 4 cm/s. Garnier capable of providing a very uniform inlet condition to the bubble
150 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

Figure 1. System used to provide the bubble column with air.

Figure 2. Top view of needle sparger (left) and picture showing bubble formation with one-third of the needles in operation (right, gas fraction about 15%).

column, something that is easily implemented in computer teristics of the needles stay the same during the experiments.
simulations and can serve as a good test as discussed in the Eleven electronic mass flow meters (VPInstruments VPFlow-
work of Monahan et al.27 Furthermore, the sparger is capable mate with a range of 0-5 slm) have been used in combination
of generating a very narrow bubble size distribution and maintain with metering valves to obtain accurate control over the flow
that over a relatively wide range of superficial gas velocities. to the groups. For an accurate measurement of the total flow
Experiments like this are very limited in the literature but, in rate, another electronic mass flow meter (VPInstruments VP-
our opinion, provide a well defined “test case”. Moreover, we Flowmate with a range of 0-100 slm) has been used. A
estimate the bubble size and measure the bubble velocity using schematic representation of the setup is given in Figure 1. Note
a four point optical probe. We have carried out experiments in that upstream of each group of needles a so-called group-needle
a 15 cm diameter column and found that the uniform flow can (inner diameter 0.6 mm, length 10 cm) is placed, in order to
be present up to 55% volume fraction of the bubbles. Finally, obtain a very uniform flow to each of the groups of needles.
at the onset of instability, we find an intriguing duality in the A picture of the sparger operating with one out of every three
flow that may give a clue to the instability we observed. needles is given in Figure 2.

2. Experimental Setup 3. Experimental Techniques


All experiments are carried out in a cylindrical column of Glass Fiber Probes. Optical glass fibers are used to probe
15 cm inner diameter, filled with tap water to an ungassed height the bubble phase. The probe’s working principle is the difference
of 130 cm. A special bubble distributor is used for the injection in refraction index between glass (the probe material), water,
of the bubbles. It consists of 559 needles, with an inner diamet- and air. Light is sent into the glass fiber. At the tip, it will reflect
er of 0.8 mm and a length of 20 cm where each needle exit is back when the tip (size 200 µm) is surrounded by air. Most of
5 mm above the column bottom. More details can be found in the light will propagate into the water phase if the tip is
ref 25. The needles are placed in a triangular pattern, with a surrounded by water. By detecting the reflected light, the phase
pitch of 6 mm. The flow through the needles is controlled in present at the probe tip is detected (see, e.g., refs 33–35). Gas
small groups. In total, the needles are grouped in eleven groups. fraction profiles are measured using a single point optical probe.
The number of needles operating is a function of the required First, the time series recorded is made binary via an appropriate
gas fraction. In order to obtain a uniform bubble size, the gas threshold (thR). The latter is set at 10% of the difference between
flow rate through each needle needs to be between 1 and 3 mL/ the output value of the tip in air (sa) and that in water (sw).
s. Therefore, for gas fractions below 15%, 187 needles are used,
i.e. one out of every three. For larger gas fractions, all needles
are in operation. This split ensures that the bubbling charac- b(tj) ) { 1
0
if s(tj) > thR ) 0.1(sa - sw) + sw
else
(1)
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 151
LDA. The LDA equipment consists of a 4W Spectra-Physics
Ar+ laser and a TSI 9201 colorburst multicolor beam separator.
The beam pairs are focused using a backscatter probe with a
lens of 0.132 m focal length. Detected light is sent to a TSI
9230 colorlink. The Doppler bursts have been processed by a
TSI IFA-750 processor.
The application of LDA in bubbly flows is far from
straightforward. Upon measuring deeper in the bubbly flow, the
data rate gets strongly reduced due to blockage of the laser
beams by the bubbles. This reduction is approximately propor-
tional to exp(-c[Rl/db]), where c is a constant in the range of
1.5-3, l is the distance of the measuring volume in the bubbly
Figure 3. Typical signal of the piercing of a single bubble. The noise levels flow, R is the void fraction, and db is the bubble diameter (see,
of the probe are very low. The binary version of the signal is also given.
e.g., the work of Groen et al.37).
Apart from blocking the laser beams, the bubbles can also
cause Doppler bursts that will lead to a velocity measurement.
From a given burst, it is very difficult to tell whether it is caused
by a seeding particle or by a bubble. In the present experiments,
we used neutrally buoyant hollow glass particles with a diameter
of about 8 µm as seeding. It is discussed by Groen et al.37 that
under these conditions the number of velocity measurements
that actually comes from a bubble is negligible. It is therefore
safe to assume that the LDA measures the liquid velocity.
Typical data rates for void fractions near 10% are 150-400
Hz near the wall and 0.8 Hz in the column center. This shows
Figure 4. Effect of small departures from vertical alignment on the axial the difficulty of experiments deeper inside the column, especially
liquid velocity profile (R ) 8.1%). at the higher void fractions, where experiments in the center
Next the (local) gas fraction is obtained as follows: are no longer possible. The measuring time per point was set
at 900 s. Close to the wall a single data set contains more than
N 200 000 data points, whereas at a radial position r/R ≈ 0.5 at a
∑ b(t ) · ∆t
j s gas fraction of 30% this number has dropped to less than 5000.
j)1
R) (2) LDA data are not bias free. First, due to increased noise levels
N∆ts caused by the bubbles multiple validation, i.e. measuring the
with N the total number of samples, and ∆ts ) 10 µs, the sample burst of a single seeding particle multiple times, occurs more
time. The threshold should not be too high as this would cause often than in single phase flow. The multiple data points are
bubbles that touch the tip rather than being pierced to be filtered removed by rejecting those data points that are outside 4 times
out, thus underestimating the gas fraction. Obviously, it can not the standard deviation of the data set. Next, for all two data
be too close to 0 as then unwanted noise would cause an points within a time interval of 1 ms, the second data point is
overestimate of the gas fraction. An example of the signal (and removed. Second, a velocity bias correction is employed. This
its binary) of a single bubble passage is given in Figure 3. Note correction is based on the so-called 2D+ weighing: the weighing
that when the tip of the probe leaves the bubble the response is factor is inversely proportional to the magnitude of the velocity
very fast: the original signal and the binary step almost vector that is estimated from two of the three velocity
completely overlap. The measuring time per data point was set components. For a full discussion, see, e.g., the work of
at 900 s. The number of bubbles detected scales roughly linearly Tummers.38 These corrections amount to about 2 cm/s, reducing
with the gas fraction and is in all cases above 5000. the velocity. It brings the measured net cross-sectional flow of
The experimental error of measurements with the glass fibers the liquid down from 1 or 2 cm/s to values of 0.3 cm/s or less.
is made of two parts. The first part is the statistical error Column Alignment. Rice and co-workers39,40 have shown
associated to the measuring time. It is proportional to the square that very small deviations from vertical alignment result in the
root of the measuring time. Repeated experiments show that in creation of a large scale liquid circulation in the column and an
our case the absolute uncertainty in the gas fraction is increase in the axial dispersion of the bubble phase. Tinge and
0.001-0.003 for the gas fraction range studied. The second part Drinkenburg41 discuss the maximum inclination angle before
is a bias error due to the difficulties of piercing a bubble at the this circulation sets in. Extrapolation of their results gives a
bubble edge. This is caused by the surface tension force that maximum inclination of 0.05° in our case. Careful alignment
makes the bubble deviate from its vertical path upon colliding made it possible to have a horizontal deviation at a position of
with the glass fiber. This error depends on the approach velocity 1.5 m above the bottom of the column with respect to the vertical
of the bubble and on the swarm density surrounding the bubble: of less than 1 mm, which indicates a maximum inclination angle
the higher the velocity, the smaller the bias; the denser the of 0.04°. We have run a test with an average gas fraction of
swarm, the less room there is for the bubble to flow around the 8.1% (measured from the liquid level increase) to see the
glass fiber tip. In the literature,36,35 it is reported that this bias influence of the inclination angle on the axial liquid velocity as
caused an underestimate of 5-10% of the actual gas fraction. measured by the LDA. Results are shown in Figure 4. It is clear
In the present case, the swarm gets much denser than used in that even angles as small as 0.04° are visible. Note further that,
the studies mentioned. We expect that the bias will be smaller for “perfect” alignment, we measure a slightly nonzero mean
and estimate it to be 5%. Note that this effect causes a systematic flow. This is a consequence of a small bias in the LDA
underprediction of the gas fraction. measurements.
152 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

Table 1. Fit Parameters


V∞(m/s) n(-) Cµ(-)
RZ 0.195 1.40
Garnier 0.280 0.60

Garnier’s model.23 Both models are fitted to the data of Figure


5. The results are given in Table 1.
Clearly, we find a much lower power for the Richardson and
Zaki slip velocity than according to the original model for
particles with a single particle Reynolds number above 500.
However, the value of n ) 1.4 is in agreement with the one
mentioned by Joshi.12 Moreover, the single bubble rise velocity
Figure 5. Void fraction as a function of the superficial gas velocity with needs to be set at 0.195 m/s. This is too small: the bubbles in
time t after filling the bubble column with fresh tap water. Note that both the present experiment have an equivalent diameter of 4 mm
theoretical lines (RZ and Garnier) almost overlap.
(see section 4.4). These bubbles have a terminal rise velocity
of 0.25 m/s, significantly higher than the value needed for a
4. Gas Fraction and Liquid Velocity proper fit of the Richardson and Zaki relation to our data. If we
use the power n ) 1.4, the maximum possible gas fraction of
4.1. Overall Gas Fraction. The overall gas fraction is plotted
the homogeneous regime (Rmax ) 1/n) is 0.71,8,17 and the critical
in Figure 5. For low superficial velocities, the bubbly flow is in
value12 is Rcrit ) 1/(n + 1) ) 0.42. This underestimates our
the homogeneous regime and the gas fraction raises more or findings. Joshi17 provides a more extensive analysis and arrives
less linearly with the gas flow rate. However, at a certain point, at a different expression for the critical gas fraction:
this linearity breaks down and the gas fraction only slowly
increases with the superficial gas velocity. This marks the change Vs2Cv(1 + Cv) ) Ddbg(Rcrit + Cv) (5)
from homogeneous to heterogeneous flow.11 Where exactly this
in which Vs(R) is the slip velocity between the two phases, Cv
happens, depends on the purity of the water. As can be seen in
is the virtual mass coefficient, D is a dispersion constant given
Figure 5, for fresh tap water, we find the transition around
a value of 1.2, db is the bubble size, and g is gravity’s
30-35%. But as the water stays longer in the column, this point acceleration. If in eq 5, the Richardson and Zaki relation is
shifts until we finally can reach gas fractions of 55% before substituted with n ) 1.4 and the virtual mass is set to 0.5, then
the uniform flow becomes unstable. This is surprising as all for 3 mm sized bubbles the critical gas fraction is 0.19. For
reports so far on bubble columns have this transition much larger bubbles, this value decreases. Again, we have an
lower: at or below 30%. Moreover, linear stability analysis of underestimate of our maximum stable gas fraction found in the
the uniform flow predicts instability from 25-40% for spherical experiment.
bubbles or as low as 10-15% for ellipsoidal bubbles (see, e.g., The Garnier model uses an R1/3 dependence. This reflects
refs 28 and 13). We have measured the surface tension of the the bubble-bubble distance in the swarm. Garnier et al. report
water in the column as a function of time, using the Du Noüy that their bubbles have an equivalent diameter ranging from
Ring method. We find, that according to this method the surface 3.4 to 5.5 mm. This is in the same range as the bubble size in
tension stays constant: σ ) (71.6 ( 0.5) × 10-3 N/m. No our case. For the model of Garnier, we find a value of 0.28 m/s
decreasing trend with time has been found. for the slip velocity of a single bubble which is in reasonable
The data, presented in Figure 5, can be analyzed based on agreement with the actual velocity of a single bubble (≈ 0.25
m/s) and is the same value as reported by Garnier. However,
the slip velocity between the gas and liquid phase. Both the
the coefficient, Cµ, in the Garnier model is found to be 0.60 in
Richardson and Zaki (RZ) relation for the slip velocity and the
our experiments, rather than close to 1 as Garnier et al. report.
proposal by Garnier et al.23 are used. A difference with the Garnier experiments is that in our case
The slip velocity (Vs), defined as the bubble velocity (Vd) the liquid velocity is 0, whereas in their case, it is not. From
minus the velocity of the continuous phase (Vc), is connected their results, it can be observed that their findings are different
to the superficial velocities: for different liquid velocities: the lower the liquid velocity, the
quicker the slip velocity drops as a function of the gas fraction.
Ug Uliq In contrast, our data show a slower decrease. It is not clear why
Vs ≡ Vd - Vc f Vs(R) ) - (3) we observe such a difference. The single bubble rise velocity
R 1-R
is almost the same in both Garnier and our experiment.
The superficial liquid velocity can be set to zero as we are in Furthermore, the bubble size is quite comparable.
the homogeneous regime. Hence, no liquid circulation is induced If we use the Joshi criterion of eq 5 with the same parameters
and the local, averaged liquid velocity is zero. Thus, we obtain as given above, we find for the critical gas fraction a value of
a direct relation between the void fraction and the superficial 0.2. This is almost equal to the prediction using the Richardson
gas velocity: and Zaki expression for the slip velocity. Again, this value is
much smaller than our experimental one and it decreases with

{
increasing bubble size. Joshi et al.17 performed a parameter
V∞R(1 - R)n-1 RZ variation study on all variables in the criterion of eq 5. They
Ug ) R · Vs(R) ) (4)
V∞R(1 - CµR1⁄3) Garnier show in their paper that the dispersion coefficient, D, needs to
be as small as 0.5 to get a critical gas fraction above 0.5. This
with V∞ being the slip velocity of a single bubble, n the seems to indicate that at the high gas fraction we can reach, the
Richardson and Zaki power () 2.39 for bubbles of the present dispersion of the gas phase is reduced significantly compared
size in the original RZ model), and Cµ a parameter set to 1 in to the more dilute cases. An explanation for this may be that,
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 153
Table 2. Superficial Gas Velocities for the Various Gas Fractions
Studied
Ug (m/s) R (-) no. of needles
0.015 6.1% 187
0.017 7.6% 187
0.025 11% 187
0.032 16% 559
0.039 20% 559
0.049 25% 559

them. Moreover, some bacterial growth can not be completely


excluded over the long time that the water is in the column.)
Figure 6. Void fraction profiles in the entrance region (Ug ) 0.023 m/s). The superficial gas velocities that have been used are given in
Table 2. This table also shows the number of needles that have
for these dense systems, the motion of individual bubbles is been used. The flow rates through the needles for these
greatly reduced by the presence of all other bubbles. In essence conditions give equivalent bubble diameters in the size range
this is the well-known mutual hindrance that restricts the bubble 3.5-5.0 mm. If the superficial gas velocity is increased from
mobility. 0.015 to 0.025 m/s, the bubble size increases due to the larger
4.2. Radial Gas Fraction Profiles. Radial void fraction flow rates through the needles. If the superficial velocity is
profiles were determined with single glass fiber probes for increased further, the number of needles is tripled, the flowrate
various superficial gas velocities at various heights in the through the needles decreases, and the bubble diameter drops
column. Figure 6 gives an impression of the uniformity of gas again. As a result, the bubble diameter variation over the entire
injection that is achieved with the needle sparger. The result void fraction range is reduced and attains a local maximum value
shows that the void fraction profile contains small nonunifor- for R ) 11%.
mities at z ) 0.07 m, introduced by imperfections in the sparger, Figure 7 shows void fraction profiles for superficial gas
but that these nonuniformities quickly even out: at z ) 0.15 m, velocities of 0.015 and 0.049 m/s. Figure 8 shows more profiles
the void fraction distribution is very uniform. This shows that for the superficial gas velocities without large scale structures
the needle sparger can provide very uniform gas injection. for z ) 1.2 and 0.6 m. The figures show that the inside of the
Typically, the gas fraction for bubbles of the present size bubble column is very uniform for all superficial velocities,
(equivalent diameter about 4 mm) is underestimated by 5-10%, except close to the surface for R ) 25%. The low superficial
see, e.g., refs 33 and 35. This is due to the difficulty of piercing gas velocities show a wall peaking behavior for all heights. The
a bubble at its side: in many cases, this results in a bouncing same peaking behavior is observed for Ug > 0.035 m/s (R >
off from the tip rather than being pierced. Surface tension forces 18%) for z < 0.9 m. In higher parts of the column, the peak
are responsible for this bouncing. disappears. A small dip occurs near x/R ) 0.7 for all experiments
Small peaks are observed in the void fraction at distances of with Ug > 0.03 m/s. This small dip is due to an artifact of the
5 mm from the wall (about one bubble diameter). At smaller experiments with the probe. We have used 5 different probes
distances to the wall, the void fraction drops more or less simultaneous to speed up the measurements. The one positioned
linearly. The drop near the wall may be explained with a local at x/R ) 0.7 systematically slightly underpredicts the void
force driving bubbles away from the wall, such as that modeled fraction.
by Antal et al.42 This model is employed quite frequently in Additional void fraction profiles for more contaminated tap
modeling studies for the estimation of, e.g., void fraction profiles water of two weeks old are shown in Figure 9. For all the
for bubbly pipe flow or bubble columns.43,44 It is not clear, experiments from this point on, the ungassed liquid height was
however, if such a force is really present: the derivation by 1.0 m, unless noted otherwise. The behavior is similar to that
Antal42 is performed for small bubble Reynolds numbers. The observed in Figures 7 and 8. For the superficial gas velocities,
works by Takemura and Magnaudet45 and De Vries46 suggest that are significantly below the critical superficial gas velocity
that, instead, for larger bubble Reynolds numbers, bubbles may where the onset of vortical structures occurs (Ugcrit ≈ 0.07 m/s),
be attracted to and bounce against the wall. An alternative again wall peaking is observed for all heights (e.g., Figure 9a
explanation for the drop in void fraction near the wall may be with Ug ) 0.047 m/s). For the superficial gas velocities
the fact that bubbles cannot overlap with the wall. This way, somewhat below the critical superficial gas velocity (Figure 9b
their shape determines the variation in void fraction. A simple with Ug ) 0.066 m/s), the wall peaking disappears near the top
check with the assumption of an ellipsoidal shape shows that of the bubble column. For the case where dynamic large scale
this gives a reasonable approximation for the decrease in void structures are present (Figure 9c with Ug ) 0.076 m/s), the
fraction near the wall. familiar core-peaking void fraction profile is found. In this case,
Next, the shape of the void fraction profile is investigated an additional underestimation of the void fraction is present:
for various heights, superficial gas velocities, and water bubbles rising downward have a smaller probability of being
contamination levels. Since the gas injection is very uniform, pierced.47
the nonuniformities observed in the profiles are due to hydro- Wall peaking behavior has been reported for bubbly pipe
dynamic effects and not due to the sparger. First, results are flows by e.g. Serizawa et al.48 Similar void fraction wall peaking
shown for tap water of several days old (large vortical structures is observed in model predictions for bubble columns and pipe
occur for Ug > 0.06 m/s) and an ungassed liquid height of 1.3 m. flow by Guet et al.43 The behavior is explained by the effect of
This water is still relatively clean. (The water in the column is the lift force, which causes accumulation of small bubbles (i.e.,
in a batch and will get contaminated over time by contamination smaller than 5 mm) in the negative velocity part of the flow,
from the surroundings and, more importantly, from the air which is close to the wall. The position where the peaking is
supply; air needs to be sparged continuously as otherwise water observed in the current experiments is at 5 mm from the wall.
would seep into some of the needles, blocking air flow through The lift force acting on the bubbles at this location is significant
154 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

Figure 7. Void fraction profiles for 6.1% void fraction (Ug ) 0.015 m/s) (a) and 25% void fraction (Ug ) 0.049 m/s) (b). The tap water is several days old.

Figure 8. Void fraction profiles for various gas fractions at z ) 1.2 m (a) and z ) 0.6 m (b). The tap water is several days old.

Figure 9. Void fraction profiles for tap water which is 2-3 weeks old. Large scale structures are found for Ug > 0.07 m/s.
since these bubbles experience a shear due to the down flow 0.9 with a typical very small velocity around 0.01 m/s.
close to the wall. This downward liquid motion is caused by Downflow occurs for r/R > 0.9 and has a typical velocity of
the slightly lower gas fraction very close to the wall and thus around -0.04 m/s. The downflow is most likely caused by
a higher local mixture density. the low void fraction very close to the wall, which causes a
4.3. Axial Liquid Velocity Profiles. Mean axial liquid density difference between the inner bubble column regions
velocity profiles for various gas fractions are shown in Figures and the wall region. Most publications on bubble columns
10 and 4. The profiles show that for all superficial gas (e.g., refs 49 or 50) report that the inversion point (i.e., the
velocities, upflow occurs in the central region with r/R < point where uax,liq ) 0) is located around r/R ) 0.7, whereas
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 155
the bubble size has increased beyond a critical size. To further
investigate this, measurements of the actual bubble size have
been conducted. However, at the same time, the swarm becomes
denser and the wakes may be suppressed causing a transition
toward homogeneous flow according to the Fox mechanism.
Figure 12 shows the means and standard deviations for the
major (a) and minor (b) axis lengths of the bubbles obtained
with a photographic technique. For the horizontal diameter, clear
differences are observed between the case of “fresh” tap water
and the case of contaminated tap water. For the relatively clean
water, the horizontal bubble size is increasing for Ug < 0.05
m/s, whereas it is more or less constant for the contaminated
Figure 10. Mean axial liquid velocity profile for various void fractions.
water. In this range, the bubble shape for the contaminated water
The tap water is several days old. The ungassed liquid height is 1.3 m.
is more or less ellipsoidal, whereas the bubbles in the relatively
the present study finds the inversion point near r/R ) 0.9. “clean” water have much more irregular shapes and exhibit
This shows the very strong homogeneity for the present flow. strong shape oscillations. As soon as the horizontal diameter
The difference is most likely caused by the absence of any reaches a value around 5.8 mm in the region near the free
large scale structure in the flow. For the other studies, the surface, the increase levels off and strongly increased liquid
gas injection may not have been uniform enough to achieve downflow is observed close to the wall in the top regions of
this. The results close to the wall show no trend with Ug. the bubble column (“A” for fresh, “a” for contaminated water).
This is probably due to the presence of a strong preshift If the superficial gas velocity is increased slightly further, large
frequency bleed through in the LDA signal for measurements scale instability is observed, first in the top parts of the column
near the wall. This affects the mean velocity. In addition, (“B” for fresh water) and for a small further increase in the
since the wall location is determined by locating positions entire bubble column (“C” for fresh, “c” for contaminated
where zero velocity is found, a small error in the radial water).
alignment is created, which manifests itself strongest close The results could be explained by the reversal of the lift-
to the wall where the velocity gradient is largest. force direction as proposed by Lucas et al.1 The graph in Figure
4.4. End of Homogeneous Regime. In the previous sections, 13 reproduces the results of the stability analysis by Lucas et
wall peaking of the void fraction profiles was found, which is al.1 for a Gaussian bubble size distribution. If the standard
generally associated with a lift force driving the smaller bubbles deviation of the bubble diameter exceeds a certain critical value,
closer to the wall. Close to the wall a narrow region with liquid the flow becomes unstable. This critical standard deviation is a
down flow exists, providing a velocity gradient that can produce function of the mean bubble diameter. The mean and standard
a significant lift force. In the higher parts of the column, and deviation that were observed in the present experiments at the
for the higher superficial gas velocities approaching the critical onset of dynamic large scale structures are also plotted in the
gas velocity where the first large structures are observed, the same figure. The agreement with the result of the stability
velocity of the down flow increases. Nevertheless, for these analysis is good, although the good match is not very sensitive
conditions, the wall peaking disappears. This indicates that to the precise value of the standard deviation for the present
the lift force toward the wall is decreasing in magnitude. Since conditions. In addition, the result shows that the width of the
the actual velocity gradient is increasing, the decrease in the bubble diameter distribution is quite small in terms of lift-force
lift force is probably caused by a decrease of the lift coefficient, behavior.
which may possibly even reach a negative value. The results in
ref 51 show that the lift coefficient decreases if the horizontal When the gas flowrate is increased to about Ug ) 0.08 m/s,
bubble diameter approaches a critical value, which is ap- the uniform flow looses its stability. However, at this setpoint,
proximately 5.8 mm for air-water. The results, therefore, could the flow starts to oscillate between homogeneous and hetero-
indicate that the mechanism for the transition that was proposed geneous, with a period of some 20 s. The gas fraction as a
by Lucas et al.1 plays an important role. These authors argue function of time is for this condition given in Figure 14. After
that small, local disturbances of the state of uniform bubble reaching its highest gas fraction of about 55%, an instability
distribution get positive feedback from the lift force if the bubble quickly develops: from the free surface, the liquid starts flowing
size is larger than the critical size. Consequently, the distur- rapidly downward in the wall region and the entire flow becomes
bance grows and the homogeneous regime looses its stability. “turbulent”. Note that at these high gas fractions it is even close
Fox and co-workers2 proposed a different mechanism. From to the wall difficult to measure with the LDA.
their stability analysis, they concluded that the transition from The downward velocity is high enough to drag bubbles
homogeneous to heterogeneous flow could be caused by an downward. This can be seen in Figure 15 that shows the time
instability triggered by a reduction of the bubble induced trace of the vertical component of the liquid velocity in a point
turbulence. This reduction is a consequence of the suppression close to the wall as obtained with the LDA. After a few seconds,
of bubble wakes in the more dense swarms. The authors found the gas fraction reaches a minimum and the downward liquid
no need to link the transition to a change in lift coefficient as velocity a maximum. Then, the flow starts to relax, the gas
has been proposed by Lucas and co-workers. The mechanism fraction starts to climb, and the homogeneous flow is restored.
proposed by Fox and co-workers is interesting and requires When the gas fraction reaches 55%, the cycle repeats in a
additional experiments in which the focus should be on sawtooth manner.
retrieving turbulence data. From visual observation, we tend to conclude that the bubble
Figure 11 shows images of the bubbles close to the wall. It size is decreased during the “turbulent” stage. This makes sense
is evident from these pictures that with increasing gas fraction as the large scale liquid circulation will strip the bubbles
the bubble size increases. If this continues, then according to somewhat earlier from the needles of the sparger. Therefore,
the Lucas criterion, the uniform flow looses its stability when the bubble size decreases. This fits with the picture that the onset
156 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

Figure 11. Images of bubbles close to the wall for increasing void fraction (contaminated water of three weeks old).

Figure 14. Alternating regimes for Ug ) 0.08m/s: time series of the gas
fraction.

Figure 15. Alternating regimes for Ug ) 0.08m/s: time series of the axial
liquid velocity close to the wall.

rest of the column due to the lower hydrostatic pressure. Hence,


we can expect the first instability to occur at the top of the
column, with the bubbles moving away from the wall. The
relatively heavy layer close to the wall then starts moving
Figure 12. Evolution of mean and standard deviation of the major and minor downward, creating the large scale instability and moving more
bubble axis lengths for two levels of water contamination. The dimensions bubbles to the center at the position where it “bends inward”
were obtained with a photographic technique.
(for continuity reasons). Finally, when the liquid circulation
reaches the sparger, the bubbles formed are smaller and the lift
force regains its original sign but is now stronger and can restore
the uniform distribution. This takes several seconds as both the
small bubbles will have to move to the top of the column and,
more importantly, the liquid flow will have to decay. Once this
flow has decayed, the uniform gas distribution can be re-
established. However, at the same time, the bubbles formed at
the sparger will get bigger as they no longer feel the drag from
the liquid flow trying to pull them off. Thus slightly larger
bubbles will move upward, reaching the critical size at the top
of the column and the cycle repeats.
The wake suppression (as proposed by Fox et al.) could
Figure 13. Comparison of the maximum stable standard deviation of the equally well explain the above observation. Now the critical
bubble diameter according to ref 43 and the conditions in the current swarm density with the required turbulence reduction also starts
experiment at the onset of transition. at the top of the column, as due to the reduced hydrostatic
of the instability is caused by a reversal of the direction of the pressure, the bubble size is increased, and the gas fraction is
lift force. The bubbles at the top are slightly larger than in the slightly higher in the top region. Also with this model, the flow
Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009 157
may become in a transitional state. Now due to the circulation, (3) Deckwer, W.-D. Bubble column reactors; Wiley: Chicester, 1992.
the swarm gets less dense and the turbulence level may be (4) Zahradnı́k, J.; Fialová, M.; Røužička, M.; Drahoš, J.; Kaštánek, F.;
Thomas, N. H. Duality of gas-liquid flow regimes in bubble colomn reactors.
restored. Additional experiments are required to sort out the
Chem. Eng. Sci. 1997, 52 (21/22), 3811–3826.
cause of the instability. (5) Shaikh, A.; Al-Dahhan, M. H. A Review of FLow Regime transitions
in Bubble Columns. Int. J. Chem. Reac. Eng. 2007, 5, 1–68.
5. Concluding Remarks (6) Mudde, R. F. Gravity-Driven Bubbly Flows. Ann. ReV. Fluid Mech.
2005, 37, 393–423.
In this paper, we have reported experiments on bubbly flow (7) Reith, T.; Renken, S.; Israel, B. A. Gas hold-up and axial mixing in
in a 15 cm-diameter cylindrical column. The continuous phase the fluid phase of bubble columns. Chem. Eng. Sci. 1968, 23, 619–629.
(water) is in a batch. The air bubbles (equivalent diameter about (8) Lockett, M. J.; Kirkpatrick, R. D. Ideal bubbly flow and actual flow
in bubble columns. Trans. Inst. Chem. Eng. 1975, 53, 267–273.
3.5-5 mm) are injected via a special needle sparger that
(9) Zahradnı́k, J.; Kaštánek, F. Gas holdup in uniformly aerated bubble
generates a very uniform input of the bubbles. We employed column reactors. Chem. Eng. Commun. 1979, 3, 413–429.
optical glass fiber probes and LDA to assess the uniformity of (10) Maruama, T.; Yoshida, S.; Mizushina, T. The flow transition in a
the flow. bubble column. J. Chem. Eng. Jpn. 1981, 14, 352–357.
After careful alignment of the column, we obtained flat gas (11) Ruzicka, M.; Zahradnı́k, J.; Drahoš, J.; Thomas, N. H. Homogeneous-
fraction profiles up to gas fractions as high as 55%, with a small heterogeneous regime transition in bubble columns. Chem. Eng. Sci. 2001,
56, 4609–4626.
wall peaking. As a consequence of this flat gas fraction profile, (12) Shnip, A. I.; Kolhatkar, R. V.; Swamy, D.; Joshi, B. J. Criteria for
the liquid shows hardly any circulation: liquid velocities close the transition from the homogeneous to the heterogeneous regime in two-
to the wall are a few centimeters per second downward, and in dimensional bubble column reactors. Int. J. Multiphase Flow 1992, 18, 705–
the center, we measure typically 1-2 cm/s upward flow, which 726.
is partly a bias error in the LDA measurements. The downflow (13) León-Becerril, E.; Liné, A. Stability analysis of a bubble column.
Chem. Eng. Sci. 2001, 56, 6135–6141.
at the wall results in a shear field that will via a lift force push (14) León-Becerril, E.; Cockx, A.; Liné, A. Effect of bubble deformation
bubbles smaller than about 5-6 mm toward the wall. This could on stability and mixing in bubble columns. Chem. Eng. Sci. 2002, 57, 3283–
be responsible for the wall peak we observe. 3297.
The flow is steady: only small scale fluctuations are observed. (15) Lucas, D.; Krepper, E.; Prasser, H. M.; Manera, A. Investigations
Our findings are in contrast to predictions from linear stability on the stability of the flow characteristics in a bubble column. Chem. Eng.
Technol. 2006, 29 (9), 1066–1072.
analysis. Generally, the latter predicts a transition from the
(16) Bhole, M. R.; Joshi, J. B. Stability analysis of bubble columns:
uniform regime into a heterogeneous one for gas fractions in predictions for regime transition. Chem. Eng. Sci. 2005, 60, 4493–4507.
the range of 25-35%. However, if we let the water “age”, we (17) Joshi, J. B.; Desphande, N. S.; Dinkar, M.; Phanikumar, D. V.
find stable uniform flow up to 55%, i.e. well beyond the stable Hydrodynamic Stability of Multiphase Reactors. AdV. Chem. Eng. 2001,
regime predicted by the linear stability analysis for air-water 26, 1–130.
bubbles of the present size. In the homogeneous regime, the (18) Drahoš, J.; Zahradnı́k, J.; Punčochár, M.; Fialová, M.; Bradka, F.
Effect of operating conditions on the characteristics of pressure fluctuations
gas fraction and the superficial gas velocity are coupled by a in a bubble column. Chem. Eng. Process 1991, 29, 107–115.
Richardson and Zaki (RZ) type of relation, i.e. the slip velocity (19) Ruzicka, M.; Drahoš, J.; Fialová, M.; Thomas, N. H. Effect of
can be described by Vs ) V∞(1 - R)n-1 with V∞ ) 0.195 m/s, bubble column dimensions on flow regime transition. Chem. Eng. Sci. 2001,
which is too low for the present bubbles, and n ) 1.4. However, 56, 6117–6124.
also the relation proposed by Garnier et al.23 gives an adequate (20) Ruzicka, M.; Drahoš, J.; Mena, P. C.; Teixeira, J. A. Effect of
viscosity on homogeneous-heterogeneous flow regime transition in bubble
fit to the data with Vs ) V∞(1 - CµR1/3) with V∞ ) 0.28 m/s and columns. Chem. Eng. J. 2003, 96, 15–22.
Cµ ) 0.6. (21) Ruzicka, M.; Vecer, M. M.; Orvalho, S.; Drahoš, J. Effect of
At a gas fraction of 55%, the flow becomes unstable. surfactant on homogeneous regime stability in bubble column. Chem. Eng.
Instabilities at the top develop and propagate downward. The Sci. 2008, 63, 951–967.
nature of the flow changes drastically: the gas profile is no longer (22) Yao, B. P.; Zheng, C.; Gasche, H. E.; Hofmann, H. Bubble
behaviour and flow structure of bubble columns. Chem. Eng. Process 1991,
flat, but core peaking is found instead. This causes large density 29, 65–75.
differences within the mixture and gravity will act, making the (23) Garnier, C.; Lance, M.; Marié, J. L. Measurement of local flow
flow unstable. Our findings could be explained by the reversal characteristics in buoyancy-driven bubbly flow at high void fraction. Exp.
of the lift force once the bubble size passes a critical diameter,51 Therm. Fluid Sci. 2002, 26, 811–815.
resulting in a flow instability as proposed by Lucas et al.1 (24) Harteveld, W. K.; Mudde, R. F.; Van Den Akker, H. E. A.
However, care should be taken, as at these high gas fractions Dynamics of a bubble column: influence of gas distribution on coherent
structures. Can. J. Chem. Eng. 2003, 81 (3-4), 389–394.
the bubble-bubble distance is small and the wakes cannot (25) Harteveld, W. K.; Juliá, J. E.; Mudde, R. F.; Van Den Akker,
develop in a similar way as in the case of isolated bubbles. H. E. A. Large scale vortical structures in bubble columns for gas fractions
Therefore, the instability may also be triggered by the suppres- in the range of 5%-25%. Proceedings of the 16th International Conference
sion of the wake as proposed by Fox and co-workers.2 on Chemical and Process Engineering, Prague, Czech Republic, Aug 22–
26, 2004.
(26) Kulkarni, A. A.; Ekambara, K.; Joshi, J. B. On the development of
Acknowledgment flow pattern in a bubble column reactor: experiments and CFD. Chem. Eng.
Sci. 2007, 62, 1049–1072.
This work has been done via a grant from the Dutch (27) Monahan, S. M.; Vitankar, V. S.; Fox, R. O. CFD predictions for
Foundation for Fundamental Research on Matter (FOM) under flow-regime transitions in bubble columns. AIChE J. 2005, 51 (7), 1897–
the Dispersed Multiphase Flow programme. 1923.
(28) Biesheuvel, A.; Gorissen, J. Void fraction disturbances in a uniform
bubbly flow. Int. J. Multiphase Flow 1990, 16, 211–231.
Literature Cited (29) Rampure, M. R.; Kulkarni, A. A.; Ranade, V. V. Hydrodynamics
of bubble column reactors at high gas velocity: Experiments and compu-
(1) Lucas, D.; Prasser, H. M.; Manera, A. Influence of the lift force on tational fluid dynamics (CFD) Simulations. Ind. Eng. Chem. Res. 2007, 46
the stability of a bubble column. Chem. Eng. Technol. 2005, 60, 3609– (25), 8431–8447.
3619. (30) Dhotre, M. T.; Joshi, J. B. Design of a gas distributor: Three-
(2) Monahan, S. M.; Fox, R. O. Linear stability analysis of a two-fluid dimensional CFD simulation of a coupled system consisting of a gas
model for air-water bubble columns. Chem. Eng. Sci. 2007, 62, 3159–3177. chamber and a bubble column. Chem. Eng. J. 2007, 125 (3), 149–163.
158 Ind. Eng. Chem. Res., Vol. 48, No. 1, 2009

(31) Lucas, D.; Krepper, E.; Prasser, H. M. Investigations on the stability (42) Antal, S. P.; Lahey, R. T.; Flaherty, J. E. Analysis of phase
of the flow characteristics in a bubble column. Chem. Eng. Technol. 2005, distribution in fully developed laminar bubbly two phase flow. Int. J.
29 (9), 1066–1072. Multiphase Flow 1991, 17, 635–652.
(32) Van Baten, J. M.; Krishna, R. Scale effects on the hydrodynamics (43) Guet, S.; Ooms, G.; Oliemans, R. V. A. Simplified Two-Fluid model
of bubble columns operating in the heterogeneous flow regime. Chem. Eng. for gas-lift efficiency predictions. AIChE J. 2005, 51 (7), 1885–1896.
Res. Des. 2004, 82 (A8), 1043–1053. (44) Krepper, E.; Reddy Vanga, B. N.; Prasser, H.-M.; Lopez de
(33) Cartellier, A. Simultaneous void fraction measurement, bubble Bertodano, M. A. Experimental and numerical studies of void fraction
velocity, and size estimate using a single optical probe in gas-liquid two- distribution in rectangular bubble columns. Proceedings of the 3rd
phase flows. ReV. Sci. Instrum. 1992, 63, 5442–5453. International Symposium on Two-Phase Flow Modelling and Experimenta-
tion, Pisa, Italy, Sept 22-24, 2004.
(34) Mudde, R. F.; Saito, T. Hydrodynamical similarities between bubble
(45) Takemura, F.; Magnaudet, J. The transverse force on clean and
column and bubbly pipe flow. J. Fluid Mech. 2001, 437, 203–228.
contaminated bubbles rising near a vertical wall at moderate Reynolds
(35) Juliá, J. E.; Harteveld, W. K.; Mudde, R. F.; Van Den Akker, number. J. Fluid Mech. 2003, 495, 235–253.
H. E. A. On the accuracy of the void fraction measurements using optical (46) De Vries, P. W. G. Path and wake of a rising bubble. PhD thesis,
probes in bubbly flows. ReV. Sci. Instrum. 2005, 76 (3), 1–13. Twente University, The Netherlands, 2001.
(36) Barrau, E.; Riviere, N.; Poupot, C.; Cartellier, A. Single and double (47) Groen, J. S. Scales and structures in bubbly flows. PhD thesis, Delft
optical probes in air-water two-phase flows: real time signal processing University of Technology, The Netherlands, 2004.
and sensor performance. Int. J. of Multiphase Flow 1999, 25, 229–256. (48) Serizawa, A.; Kataoka, I.; Michiyoshi, I. Turbulent structure of air/
(37) Groen, J. S.; Mudde, R. F.; Van Den Akker, H. E. A. On the water bubble flow. Int. J. Multiphase Flow 1975, 2, 235–246.
application of LDA to bubbly flow in the wobbling regime. Exp. Fluids (49) Mudde, R. F.; Groen, J. S.; Van Den Akker, H. E. A. Liquid velocity
1999, 27, 435–449. field in a bubble column: LDA experiments. Chem. Eng. Sci. 1997, 52,
(38) Tummers, M. Investigation of a turbulent wake in an adverse 4217–4224.
pressure gradient using laser Doppler anemometry. PhD theis, Delft (50) Hills, J. H. Radial non-uniformity of velocity and voidage in a
University of Technology, The Netherlands, 1999. bubble column. Trans. Inst. Chem. Eng. 1974, 52, 1–9.
(39) Rice, R. G.; Littlefield, M. A. Dispersion coefficients for ideal (51) Tomiyama, A.; Tamai, H.; Zun, I.; Hosokawa, S. Transverse
bubbly flow in truly vertical bubble columns. Chem. Eng. Sci. 1987, 42 migration of single bubbles in simple shear flows. Chem. Eng. Sci. 2002,
(8), 2045–2053. 57, 1849–1858.
(40) Rice, R. G.; Barbe, D. T.; Geary, N. W. Correlation of nonverticality ReceiVed for reView January 16, 2008
and entrance effects in bubble columns. AIChE J. 1990, 36 (9), 1421–1424. ReVised manuscript receiVed June 24, 2008
(41) Tinge, J. T.; Drinkenburg, A. A. H. The influence of slight Accepted July 3, 2008
departures from vertical alignment on liquid dispersion and gas hold-up in
a bubble column. Chem. Eng. Sci. 1986, 41 (1), 165–169. IE8000748

You might also like