You are on page 1of 19

International Journal of Multiphase Flow 75 (2015) 88–106

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / i j m u l fl o w

Experimental investigation of the influence of column scale, gas density


and liquid properties on gas holdup in bubble columns
Philipp Rollbusch a,⇑, Marc Becker a, Martina Ludwig a, Andrè Bieberle d, Marcus Grünewald b,
Uwe Hampel d,e, Robert Franke a,c
a
Evonik Industries AG, Paul-Baumann-Straße 1, 45772 Marl, Germany
b
Ruhr-University Bochum, Laboratory of Fluid Separation, Universitätsstraße 150, 44801 Bochum, Germany
c
Ruhr-University Bochum, Universitätsstraße 150, 44780 Bochum, Germany
d
Institute of Fluid Dynamics, Helmholtz-Zentrum Dresden – Rossendorf, Bautzner Landstr. 400, 01328 Dresden, Germany
e
Technische Universität Dresden, Germany

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

Article history: Measurements of gas holdups in bubble columns of 0.16, 0.30 and 0.33 m diameter were carried out.
Received 6 November 2014 These columns were operated in co-current flow of gas and liquid phases and in semibatch mode. The
Received in revised form 13 May 2015 column of 0.33 m diameter was operated at elevated pressures of up to 3.6 MPa. Nitrogen was employed
Accepted 15 May 2015
as the gas phase and deionized water, aqueous solutions of ethanol and acetone and pure acetone and
Available online 25 May 2015
cumene as the liquid phase. The effects of differing liquid properties, gas density (due to elevated pres-
sure), temperature, column diameter and superficial liquid velocity on gas holdup were studied. The gas
Keywords:
holdup measurements were utilized by differential pressure measurements at different positions along
Multiphase flow
Bubble column
the height of the bubble columns which allowed for the identification of axial gas holdup profiles. A
Scale-up decrease of gas holdup with increasing column diameter and an increase of gas holdup with increasing
Gas density pressure was observed. The effect of a slightly decreasing gas holdup with increasing liquid velocity
Organic solvents was found to exist at smaller column diameters. The use of organic solvents as the liquid phase resulted
in a significant increase in gas holdup compared to deionized water. It is found that published gas holdup
models are mostly unable to predict the results obtained in this study.
Ó 2015 Elsevier Ltd. All rights reserved.

Introduction semi-empirical approaches to calculate gas holdups only.


Unfortunately the predictions of available models tend to fail if
Within the chemical and petrochemical industry bubble col- they are used for scale-up purposes or to predict holdups for sys-
umns are present as multiphase reactors and contactors in a vari- tems with physical properties other than they are derived from.
ety of processes. Bubble columns are thereby utilized in various But even if the commonly investigated air/water system is exam-
modes of operation, ranging from semibatch to co- and countercur- ined with different available correlations immense variations of
rent operation with two or three phases involved. The basic con- the results are observed. The reasons for this can be found in sev-
struction of bubble columns is relatively simple, unless no eral factors. A first point to be stated is that the experimental facil-
internals are present, as they are mainly cylinders in which gas ities differ in terms of column diameter, height to diameter H/D
and liquid are brought in contact. The main features of bubble col- ratio and mode of gas distribution. There are several recommenda-
umns have been summarized by e.g. Deckwer (1985) and Kantarci tions summarized by e.g. Shah et al. (1982) concerning the mini-
et al. (2005). mum diameter (at least 0.15 m) and H/D ratio (greater than 5)
Despite the simple construction, precise prediction of the gov- which should be used in order to measure gas holdups indepen-
erning hydrodynamic parameters and the overall flow field is still dently from undesired side effects. A second point concerns the
not possible which has been pointed out by Jakobsen et al. (2005) qualities of the liquids used. Even if deionized water or tap water
recently. This can also be seen from the fact that decades of is used as the liquid phase different water qualities and accidental
researchers tried to develop models based on empirical and impurities cause differences in the experimental data. This is due
to a bubble coalescence inhibiting or promoting effect of the speci-
fic impurity. A third point accounts for the availability of experi-
⇑ Corresponding author. Tel.: +49 02365494792.
mental data especially for scale-up and gas density studies. Only
E-mail address: Philipp.Rollbusch@evonik.com (P. Rollbusch).

http://dx.doi.org/10.1016/j.ijmultiphaseflow.2015.05.009
0301-9322/Ó 2015 Elsevier Ltd. All rights reserved.
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 89

a few studies, e.g. by Forret et al. (2003), Krishna et al. (2001), Table 2
Krishna and Ellenberger (1996) and Wilkinson et al. (1992), with Density of nitrogen at various pressures.

varying column diameters are present up to this date and their p (MPa) 0.1 1 1.85 3.6
results are contradictory. Therefore even fewer gas holdup models Nitrogen
exist which account for the influence of column diameter. Density (kg/m3) 1.15 11.50 21.28 41.38
It is the purpose of this paper to present and discuss gas holdup
results obtained in three gas–liquid bubble columns of different
sizes but comparable gas distributors and liquids employed. In Table 3
addition the influence of impurities is simulated by adding small Liquid properties at different temperatures.
amounts of ethanol and acetone to the liquid phase. To discuss T (°C) 20 50 75
the effect of gas density due to elevated pressure on gas holdup
Deionized H2O
experimental studies at pressures of up to 3.6 MPa were carried Density (kg/m3) 998 988 975
out. Some other influencing parameters which are important for Viscosity (mPa s) 1 0.55 0.38
production scale bubble columns like temperature and liquid Surface tension (N/m) 0.074 0.068 0.063
superficial velocity are also examined within the studies presented. Acetone
Density (kg/m3) 767 – –
Viscosity (mPa s) 0.32 – –
Surface tension (N/m) 0.024 – –
Experimental facilities and procedures
Cumene
Density (kg/m3) 867 844 823
Experimental facilities
Viscosity (mPa s) 0.79 0.54 0.42
Surface tension (N/m) 0.028 0.025 0.022
To perform the experimental studies three bubble columns of
different diameters and heights were set up. Table 1 summarizes
the column dimensions together with their H/D ratio based on liq-
uid height.
As can be seen from Table 1 all columns are above the minimum
H/D ratio of 5 and the minimum diameter of 0.15 m mentioned by
Shah et al. (1982) to avoid any wall effects on gas holdup during
the measurements. The columns of 0.16 and 0.3 m diameter are
used to study the effect of column dimensions, superficial liquid
velocity and liquid properties on gas holdup. A third column of
0.33 m diameter is primarily used to examine the effect of a higher
gas density due to elevated pressures and the effect of temperature
on gas holdup. As the difference in diameter to the 0.30 m diameter
column is small, no remarkable effects of scale are expected.
Nitrogen was always used as the gas phase (see Table 2 for nitro-
gen densities at investigated pressure levels) and deionized water,
acetone, cumene and aqueous solutions of organic solvents as liq-
uid phase (properties related to investigated temperature levels
listed in Table 3).
All columns were operated in co-current flow of gas and liquid
phase. The gas was distributed by a perforated plate sparger with
holes of 1 mm diameter. The spargers were designed according
to the methods proposed by Ruff et al. (1976) and its dimensions
are listed in Table 4. All spargers match flow characteristics in each
column which results in a different number of openings due to the
varying column diameters and the associated flow rates.
Simplified schematics of all three facilities are given in Figs. 1–3.
Note that nearly all safety devices, valves and outlets are not
shown here to enhance clarity of the depicted experimental setups.
Safety devices include for example pressure relief valves, concen-
Fig. 1. Simplified schematic of 0.16 m diameter glass column.
tration sensors, groundings, buffer vessel level indication and auto-
matic shut-down mechanisms. Fig. 1 shows the 0.16 m diameter
column which is made of glass. measurement error). Liquids employed were deionized water,
Liquid is circulated via a pump from bottom to top of the col- aqueous solutions of ethanol, acetone and cumene. Nitrogen as
umn. At the top the liquid leaves the column through an overflow the gas phase also enters the column at the bottom and is dis-
and flows into a buffer vessel. The liquid flow rate is measured by a tributed by a perforated plate sparger. It leaves the column at the
Coriolis flow meter (Endress + Hauser, promass63a, 0.1% top from where it enters the buffer vessel to separate entrained liq-
uid from the gas. Afterward nitrogen passes through a condenser,
again to separate liquid and gas, before it enters the exhaust sys-
Table 1
tem. The amount of gas flowing through the column is measured
Column dimensions and H/D ratio.
by two gas flow meters (Krohne, H250, 1.6% measurement error),
Column diameter D (m) Liquid height H (m) H/D ratio (–) one for low and one for higher gas throughputs, to ensure a better
0.16 1.8 11.25 accuracy of the measurement. Gas and liquid superficial velocities
0.30 2.63 8.75 were varied up to 0.1 m/s and 0.01 m/s respectively. Gas holdups
0.33 3.88 11.75
are measured by glass capillaries which are connected with the
90 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

Fig. 2. Simplified schematic of 0.3 m diameter glass column.

column by PTFE hoses. The glass capillaries measure the pressure Basically the operation of this column is identical with the for-
difference caused by the gas flowing through the column by liquid mer one. Gas and liquid enter the column at the bottom and flow
level indication. To avoid inaccuracies by dynamic pressure losses co-currently to the top of the column. Liquid leaves the column
caused by the passing gas bubbles PTFE plugs with 1 mm openings via an overflow and flows into a buffer vessel to be recirculated
are installed at the bottom of each glass capillary. The level indica- by a pump. The gas passes through a series of condensers (due to
tors allow for the determination of gas holdups along the column reasons of simplification only one is shown) to get rid of entrained
axis in three 0.6 m sections which are denominated as Sections liquid and leaves through the exhaust. The liquid and gas flow rates
S1 to S3. The calculation method is provided in the later section are identical with the ones of the 0.16 m diameter column.
of this chapter. Nitrogen was used as gas phase and deionized water, aqueous
The second glass column 0f 0.3 m diameter is sketched in Fig. 2. solutions of ethanol and acetone as liquid phase. Again, gas
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 91

Fig. 3. Simplified schematic of 0.33 m diameter stainless steel column.

holdups are measured by level indication in glass capillaries which the exhaust. Liquid is circulated by a pump and its flow rate is
are damped by PTFE plugs to ensure higher accuracies as described measured by a flow meter (Krohne H250, 1.6% measurement
before. Similar to the smaller column the positions of the holdup error). It is possible to heat the liquid up to 75 °C before entering
measurements are distributed along the column axis to allow for the column by the use of a heat exchanger. Deionized water and
the measurement of the axial gas holdup distribution. cumene were employed as liquid phase. Gas holdups were mea-
The third column used in this study is pictured in Fig. 3. sured by six differential pressure transmitters (Endress + Hauser,
As this column is operable at elevated pressures the functional- Deltabar S FMD78) distributed along the column height (see
ity of this stainless steel column is somewhat different compared Fig. 3 for details and distances) to measure the axial evolution of
to the above described two glass columns. First of all liquid and dispersed phase holdup. The evaluation procedure is presented in
gas are in co-current flow and enter the column at the bottom. Section ‘Procedures and data evaluation’. If the column is to be
Nitrogen, which was used as the gas phase, is provided by a com- operated under pressure a backpressure regulator at the gas outlet
pressor and introduced to the column by a perforated plate sparger was used to adjust the pressure. The pressure itself is raised by the
(see Table 4 for details). The gas flow is measured by a gas flow use of a nitrogen gas bundle and varied between 0.1 and 3.6 MPa.
meter (Krohne, H250, 1.6% measurement error). After the gas Radial gas holdups were measured by a gamma ray CT and a
leaves the column at the top it enters a buffer vessel for phase sep- wire-mesh-sensor (WMS) which were developed by the
aration. Afterward it is cooled by a condenser and leaves through Helmholtz-Center Dresden-Rossendorf (see Bieberle et al., 2013
and Schlusemann et al., 2013 for details). As the gamma ray CT
measurements are based on radiation transmission, the section
Table 4
Sparger geometries. of measurement is constructed with a lower wall thickness of
30 mm which is sketched in Fig. 3.
Column diameter D (m) Number of holes (–) Free area (%)
For radiation based CT a radiation source is directed to an object
0.16 92 0.36 of interest and a detector measures the radiation attenuation by
0.30 352 0.85
the object of investigation. For full CT scans such radiographic pro-
0.33 352 0.65
jections must be obtained from various angular positions. The data
92 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

sets are then used as input for CT reconstruction algorithms to cal- is that level measurements in glass capillaries are used to obtain
culate the material distribution within a measuring slice or volume gas holdups. As both methods are based on pressure differences
section. In contrast to medical CT, isotopic sources with high pho- the holdup calculations are similar and presented below.
ton energy can be used for industrial CT. This enables penetration Gas holdup (eG) is usually defined as the ratio of gas volume to
of dense walls of a few centimeters. However, the higher the pho- total two or three phase volume in the column (where VG is the gas
ton energy of the isotopic source the worse is the phase contrast volume and Vl the liquid volume).
between, e.g. gas and liquid. HireCT is a transportable CT system
VG
and consists mainly of three parts: an isotopic source, a radiation eG ¼ ð2-1Þ
VG þ Vl
detector arc and a rotational unit. As isotopic source 137Cs with
an activity of 180 GBq is used emitting gamma photons with an Eq. (2-2) yields the easiest way of estimating holdups by measuring
energy of approximately 662 keV. The radiation is limited to a the clear liquid height H0 and the gassed liquid height HG.
40° wide and 8 mm height fan beam and is automatically HG  H0
moved-back into a shielding container in case of a power loss. eG ¼ ð2-2Þ
HG
The radiation detector arc consists of 320 temperature stabilized
scintillation detector elements operated in pulse mode and each As this method of measurement is prone to uncertainties because
with an active area of 2 mm in width and 4 mm in height. HG might be difficult to measure accurately due to disengaging
Projection data read-out is automatically triggered by an optical gas bubbles at the surface a manometric method was chosen for
positioning system installed below a rotational ring on which holdup measurements. If one-dimensional steady-state flow,
source and detector arc are placed on. The spatial resolution of isothermal behavior, constant properties and negligible
HireCT is about 2 mm. Note, CT scans take several minutes, thus, cross-sectional mass transfer are assumed, Eq. (2-3) represents a
only averaged phase fraction distributions can be measured flow model to calculate gas holdups.
(Bieberle et al., 2013; Hampel et al., 2007).  
Dp 4sw u2l DeG
The wire-mesh sensor consists of two planes of 64 parallel, eG ¼ 1 þ þ þ ð2-3Þ
equally distributed, stretched wires positioned orthogonally but
ql g Dh ql gD ð1  eG Þg Dh
offset by a small axial distance of approximately 2 mm. It was where ql is the liquid density, Dp pressure difference between two
especially developed for the high pressure column. Thus, spatial points, Dh the height difference, sw the wall shear stress, D the col-
resolution of about 5 mm is achieved. One wire plane is operated umn diameter, ul liquid velocity and g acceleration due to gravity.
as a transmitter plane, while the second acts as a receiver plane. According to Hills (1976) and Tang and Heindel (2006) the neglec-
The working principle is to measure the local instantaneous gas tion of inertia and shear forces is justified at low superficial liquid
holdup at the virtual crossing point of transmitter and receiver velocities (ul < 0.1 m/s) in co-current two-phase flow. If shear stress
wires. By activating each transmitter wire successively, the electri- and inertia forces are neglected, Eq. (2-3) simplifies to Eq. (2-4).
cal currents at each virtual crossing point, flowing toward the Rearranging yields Eq. (2-5), which can be used to calculate gas
receiver wires, are measured. Data sampling rates of up to holdups at the experimental conditions of this study (x refers to
10,000 Hz are possible. the axial position and Dx is the distance between two points of
To visualize the expected flow regimes in this study, the above measurement).
introduced columns and their operating conditions with respect to
Dp ¼ ql g Dxð1  eG ðxÞÞ ð2-4Þ
superficial gas velocity are marked in Fig. 4. This classification is
taken from Shah et al. (1982) and is only valid for air/water and
Dp
air/dilute alcohol systems at atmospheric pressures. It can be seen eG ðxÞ ¼ 1  ð2-5Þ
from Fig. 4 that the studied flow regime in this work is mainly the
ql g Dx
homogeneous flow regime. As level indication in capillaries is used for both glass columns, Eq.
(2-6) is used to calculate holdups for these columns (Dh refers to
Procedures and data evaluation the difference of liquid level of capillaries).
Dh
It has been stated that gas holdups in all three columns were eG ðxÞ ¼ 1  ð2-6Þ
Dx
measured by the manometric method. The difference between
the two glass columns and the pressurized stainless steel column It should be noted that it is vital to clean the capillaries after each
experimental run because Eq. (2-5) is only valid if exactly the same
fluid is present in the column and all capillaries. Gas holdups in the
pressurized stainless steel column of 0.33 m diameter could be cal-
culated with Eq. (2-4). A study by Tang and Heindel (2006) offers
another calculation method, which takes wall shear stresses into
account. Eq. (2-7) represents the proposed method of estimation
(Dp0 is the pressure difference at zero gas flow).
Dp
eG ðxÞ ¼ 1  ð2-7Þ
Dp0
As can be seen from Eq. (2-7) the proposed method requires the
measurement of differential pressures Dp0 for each operating con-
dition without sparging gas into the column. As this can be comfort-
ably realized and Eq. (2-7) yields more accurate holdup values
according to Tang and Heindel (2006), this method is chosen to
obtain the experimental holdups of this study.
To ensure a high accuracy of measurement all pressure differ-
ence readings of the stainless steel column are recorded for a per-
Fig. 4. Expected flow regimes in this study, modified from Shah et al. (1982). iod of 10 min with a frequency of 1/s after steady state conditions
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 93

are met. Steady state means in this case that no serious fluctua- breakage rate exceeds the coalescence rate of the bubbles. The
tions of the data readings are observed after startup or after chang- bubble frequency at the sparger also increases with decreasing sur-
ing the experimental conditions. These values are then averaged face tension because less momentum is needed to detach a bubble
and the holdup is calculated with Eq. (2-7). In addition, all flow from a sparger orifice and the time for bubble growth before
and other related measurements are also recorded and averaged detachment is reduced (Kulkarni and Joshi, 2005). As a result smal-
over the same period of time. As no signal processing is possible ler primary bubbles are formed at the sparger. The same is true for
for the glass column setups, all experiments are repeated to vali- the effect of a lower liquid viscosity on gas holdup. A higher viscos-
date the measurement principle. The total holdup in all three col- ity favors bubble growth and coalescence and consequently
umns is calculated by averaging the holdups of each axial section. reduces gas holdups (Urseanu et al., 2003). Another interesting
The superficial gas velocity ug is calculated at operating condition point is an observable shift of bubble shapes from larger wobbling
to account for changes in gas density. nitrogen bubbles in water to smaller spherical bubbles in acetone
and cumene as shown in Fig. 7.
This trend can also be theoretically derived from a diagram pro-
Results
posed by Clift et al. (1978) if the dimensionless Morton and Eötvös
numbers for a specific gas/liquid system are calculated. Eö and Mo
Influence of liquid properties
numbers are listed in Table 5 for assumed bubble diameters of
0.001–0.01 m (which seems to be a realistic range based on litera-
In this section the results obtained in the 0.16 m diameter col-
ture data assembled by Kulkarni and Joshi, 2005). It can be seen
umn at atmospheric pressure are presented. Dispersed phase mea-
from Table 5 that bubble shapes for acetone shift to spherical cap
surements were carried out with deionized water, aqueous
bubbles and for cumene to spherical and ellipsoidal bubbles
solutions of organic solvents, acetone and cumene as the liquid
dependent on the Reynolds number. A different bubble shape
phase and nitrogen as the gas phase. The superficial gas velocity
affects the drag force acting on a bubble and therefore also influ-
did not exceed 0.10 m/s and it can be expected that the bubble col-
ences the bubble movement within the liquid which in turn inter-
umn was mainly operated in the homogeneous flow regime and at
acts directly with gas holdup.
the beginning of the transition to heterogeneous flow. Fig. 5 shows
As acetone and cumene have comparable physical properties no
the obtained overall gas holdup values of nitrogen in deionized
significant differences in holdup might be expected. Nevertheless
water, acetone and cumene. In all three systems the gas holdup
at about 0.03 m/s an observable mismatch between gas holdups
rises with increasing superficial gas velocity.
in acetone and cumene can be noted. This difference can be attrib-
It is obvious that the increase in holdup is significantly higher in
uted either to the formation of froth at the top of the column if the
both organic solvents. The reason for this is a lower viscosity, lower
superficial gas velocity exceeds 0.03 m/s in cumene or to a change
liquid density and an about three times lower surface tension of
in flow regimes. As it is inaccurate and often impossible to distin-
acetone and cumene compared to water. A lower liquid density
guish between homogeneous and heterogeneous flow by optical
decreases the buoyancy force acting on a bubble as the density dif-
observation, a method to determine the point of regime transition
ference between both phases lowers. This together with smaller
based on measured parameters will be used. The regime transition
diameter bubbles reduces the bubble rise velocity which in turn
point can either be obtained directly from a holdup vs. superficial
increases the gas holdup. Bubble swarm velocities can be obtained
gas velocity diagram if a clear change of the holdup curve gradient
by Eq. (3-1) (Deckwer, 1985) directly from the experimental data
occurs or it can be estimated with a Wallis plot as shown in Fig. 8
(ubs is the bubble swarm velocity).
(Shaikh and Al-Dahhan Muthanna, 2007). The approximated point
ug of regime transition is the point where the measured holdup values
ubs ¼ ð3-1Þ
eg level off from the fitted drift flux curve. The fitting parameter is the
bubble rise velocity which was estimated by parameter fitting to
With Eq. (3-1) calculated swarm velocities are plotted in Fig. 6 and a be 0.47 m/s in deionized water and 0.254 m/s in acetone and
significantly lower swarm velocity is obtained for bubbles in ace- cumene. It is noteworthy that the fitted parameters agree with
tone and cumene which proves the previous statements. the calculated bubble swarm velocities in Fig. 6. With this method
Decreasing the liquid surface tension enhances bubble breakage transition holdups of 0.068 in water, 0.12 in acetone and 0.16 in
and therefore promotes the existence of smaller bubbles if the

Fig. 5. Measured gas holdups for N2/H2O, acetone and cumene. Fig. 6. Calculated bubble swarm velocities for H2O, acetone and cumene.
94 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

(a) (b) (c)

Fig. 7. Photographs of nitrogen bubbles in (a) water (b) acetone and (c) cumene.

Table 5 investigated gas/liquid system. More details are summarized in


Eötvös and Morton numbers for N2 bubbles (dB = 0.001. . .0.01 m, p = 0.1 MPa). Table 6 and it can be concluded that the facilities mainly differ in
Liquid Eö (–) log(Mo) (–) terms of H/D ratio and sparger type used. For completeness it
H2O 0.13. . .13.41 10.6 should be mentioned that Ohki and Inoue (1970) used three col-
Acetone 0.31. . .30.69 11.02 umns (0.04, 0.08 and 0.16 m in diameter) and different types of
Cumene 0.30. . .30.13 9.72 spargers. The setup listed in Table 6 is the closest to our own.
It is obvious that the results measured in this study are the low-
est in Fig. 9. To explain this behavior the following reasons can be
cumene were obtained. Translated into transition velocities this identified. First of all it is possible that very small sparger orifices
means that the flow regime starts to shift to heterogeneous flow produce very small bubbles which generate higher holdup values.
at about 0.03 m/s in water, 0.034 m/s in acetone and 0.038 m/s in A proposal of minimum diameter openings of 1 mm was given by
cumene. Wilkinson et al. (1992) in order to measure gas holdups indepen-
These values also agree with the points where the bubble dently from sparger design and not to mask other effects occurring
swarm velocities begin to rise in Fig. 6. The rising swarm velocities during the experiments. As only Letzel et al. (1999), who measured
at the point of regime transition might possibly be attributed to the the highest holdups of the studies presented here, used openings
formation of larger bubbles. At this point it can be stated that the with a diameter of less than 1 mm this alone cannot be the expla-
starting point of regime transition shifts to higher gas velocities if nation for the deviations. It should be stated that the sparger used
surface tension and viscosity of the liquids are lower than in water in this study is not completely uniformly distributing the gas until
because of less bubble coalescence. The experimentally deter- a superficial gas velocity of 0.04 m/s is reached, which might
mined regime transition velocity for water agrees well with data account for lower holdups obtained in the distributor region at
from Krishna et al. (2000a)) and Letzel et al. (1999), who experi- lower gas throughputs. As can be seen from Fig. 9, the measured
mentally determined the point of transition of an air/water system holdups begin to differentiate from literature data at gas velocities
in a 0.15 m diameter column to be at a superficial gas velocity above 0.02 m/s and it can be concluded that the effect of
slightly above 0.02 m/s and 0.03 m/s respectively. An interesting non-uniform gas distribution on overall holdup is not too large. A
point is that the transition holdups measured by Krishna et al. second factor listed in Table 6 is the aspect ratio of the column.
(2000a) and Letzel et al. (1999) are around 10% higher than those A study by Ruzicka et al. (2001) revealed that at constant column
obtained in this study. Fig. 9 compares the measured overall gas diameter, overall gas holdups decrease with increasing column
holdups of these authors with the ones obtained in this work and height. If this would be the explanation than Krishna et al.
data from Grund et al. (1992) and Ohki and Inoue (1970). (1991) should have obtained higher results than Ohki and Inoue
The criteria for selecting these authors were a comparable (1970) or Grund et al. (1992). Because Fig. 8 indicates a change
experimental setup with respect to column diameter and

Fig. 9. Overall gas holdups of this study compared with data from Krishna et al.
Fig. 8. Estimation of regime transition holdup for N2/H2O, acetone and cumene. (1991), Letzel et al. (1999), Grund et al. (1992) and Ohki and Inoue (1970).
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 95

Table 6
Experimental setups of publications depicted in Fig. 9.

Author D (m) H/D (–) Gas/liquid Sparger type


Krishna et al. (1991) 0.16 7.5 N2/H2O Ring sparger, 37  2 mm
Letzel et al. (1999) 0.15 8.13 N2/H2O Perforated plate, 200  0.5 mm
Grund et al. (1992) 0.15 28.66 Air/H2O Perforated plate, 7  2.3 mm
Ohki and Inoue (1970) 0.16 18.75 Air/H2O Perforated plate, 37  2 mm
This study 0.15 11.25 N2/H2O Perforated plate 92  1 mm

consequence a higher gas holdup is found in coalescence hindered


systems. This surface tension gradient also increases the drag act-
ing on the bubble and thus slowing down its rise velocity (Shah
et al., 1985). Additionally solvent addition might cause a stiff bub-
ble surface which in turn affects the bubble drag coefficient which
in turn influences bubble movement.
Fig. 10 shows that gas holdup in a 1 vol.% aqueous ethanol solu-
tion is 2.2 times higher than in pure deionized water. The addition
of 0.1 vol.% ethanol and same amounts of acetone to water also
causes a remarkable increase of the measured holdups. No signifi-
cant difference of measured holdups is observable for 0.1 vol.%
ethanol and acetone. The large discrepancy between ethanol and
acetone concentrations of 1 vol.% above 0.03 m/s superficial gas
velocity might be related to evaporation and entrainment of vola-
tile acetone during the measurement. As the quantities of added
solvents are very low the downstream condensers were not
expected to condense entrained acetone or ethanol.
There are certain safety requirements to be met if large quanti-
Fig. 10. Effect of solvent addition to deionized water on gas holdup. ties of organic liquids are used for experimentation. That is why it
might be desirable to substitute these pure liquids with less dan-
gerous media like aqueous mixtures with low organic content as
of flow regime there is no relationship between sparger effectivity discussed above. Fig. 11 proves that this approach is inappropriate.
and the estimated point of regime transition. In the event of full Gas holdups measured in acetone are lower than in aqueous solu-
sparger effectivity the measured holdups should still rise linearly tions of 0.1 and 1.0 vol.% acetone. This can be referred to the com-
without a change of slope. plex interaction of liquid properties as discussed above. The
One parameter which seems to be underestimated is the quality mechanism of inhibition of bubble coalescence due to addition of
of the water used for the measurements. This has recently been small amounts of organic solvents can primarily be related to bub-
stated in a publication by Kemoun et al. (2001) as well. If tap water ble–liquid interphase phenomena while the overall mechanism of
is used the liquid qualities usually differ from location to location. bubble generation, movement, coalescence and breakup in pure
But even if deionized water is used the qualities may differ from liquids must be explained by a more complex interaction of liquid
day to day since the ion exchangers used to deionize the water properties in general.
become less effective with time and need to be renewed. In addi-
tion it is often not possible or at least very difficult to clean pilot
Influence of scale and liquid velocity
scale facilities. Consequently remnants of used liquids and dirt
may be present and affect each experiment. This seems to be plau-
After discussing the influences of pure liquid properties and
sible as small amounts of surfactants already decisively influence
binary liquid mixtures on gas holdup two other parameters are
the gas holdup which was proven by a number of publications
(Camarasa et al., 1999; Keitel and Onken, 1982; Krishna et al.,
2000a,b; McLaughlin, 2003; Nguyen et al., 2012; Posarac and
Tekic, 1987; Prince and Blanch, 1990; Syeda and Reza, 2011;
Zahradnik et al., 1995). The effect of addition of small amounts of
solvents to deionized water on gas holdup is shown in Fig. 10.
An increase in gas holdup can already be observed for low sol-
vent concentrations of 0.1 vol.% in deionized water. The reason for
this is a coalescence inhibiting effect of polar solvents in water.
After absorbing at the phase interface the hydrophilic polar part
of a solvent will orient toward the water while the hydrophobic
part will arrange itself toward the bubble. This orientation results
in a repelling effect of approaching bubbles and thus hinders bub-
ble coalescence (Keitel and Onken, 1982). Another statement asso-
ciated with coalescence hindrance due to addition of solvents
implies that a local surface tension gradient occurs at the inter-
phase of dispersed and continuous phase which prevents bubble
coalescence (Syeda and Reza, 2011) which is also related to the
Marangoni effect. If less coalescence takes place, the bubble size
distribution is expected to shift to smaller bubbles and as a Fig. 11. Comparison of aqueous acetone solutions with pure acetone.
96 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

of primary concern when designing bubbling multiphase contac-


tors. It has been stated that bubble column reactors are often used
for slow gas–liquid reactions and that these reactors are operated
continuously. Therefore the influence of liquid velocity on dis-
persed phase holdup needs to be examined. Furthermore the influ-
ence of column diameter on gas holdup will be discussed in this
chapter in order to be able to gain insights into holdup behavior
at larger reactor scales. The analysis of both parameters is neces-
sary as the few available publications state different opinions
about their influence on gas holdup.
As an example Wilkinson et al. (1992) and Forret et al. (2003)
state no significant influence of column diameter on gas holdup.
A constraint for the validity of the statement above given by
Wilkinson et al. (1992) is that the column diameter needs to be lar-
ger than 0.15 m. On the other hand, Krishna et al. (2001), Krishna
and Ellenberger (1996) and Botton et al. (1978) found a decrease
in gas holdup with increasing column diameter in both homoge-
neous and heterogeneous flow regime. Botton et al. (1978) state
additionally that this is true for their experimental results if col- Fig. 12. Influence of column diameter on gas holdup.
umn diameters of less than 0.25 m are used. Table 7 summarizes
the experimental conditions of each group. According to Ruzicka et al. (2001), columns of increased height at
As the flow conditions of this study are limited to maximum a fixed diameter have lower gas holdups as their shorter relatives.
superficial gas velocities of about 0.12 m/s, the comparison of the This means that the column height is the main cause for the differ-
present results will be restricted to that point. It is noteworthy that ences between gas holdups measured in the 0.3 and 0.33 m bubble
Wilkinson et al. (1992) examined gas holdups at elevated pres- column. While the holdups of the 0.16 and 0.30 m diameter col-
sures, too. This will be discussed in a later section. Forret et al. umn have a difference of 1.5–2% at superficial gas velocities below
(2003) disclosed only one holdup value for each column diameter. 0.04 m/s, the comparison between the 0.30 and 0.33 m diameter
Thus a comparison of their results with other authors seems not column reveals that gas holdups start to differ at superficial gas
expedient. From the above mentioned authors only Botton et al. velocities above 0.02 m/s. If both glass columns are compared
(1978) operated two of their columns with non-stagnant fluids one can also see that the difference between holdups above super-
but concentrated on very high gas throughputs. ficial gas velocities of 0.04 m/s, which is in the region of flow
There are nearly no deviations of gas holdups observable in regime transition, seems to be constant if measurement errors
both columns used by Wilkinson et al. (1992). The data of are considered. The fluctuation of holdups occurs because the bub-
Krishna et al. (2001) show a significant decrease of gas holdup if ble size distribution developed at these conditions becomes more
the column diameter is enlarged from 0.1 to 0.15 m. An increase non-uniform than in the homogeneous flow regime which in turn
of lower magnitude of gas holdup can be noted if the column diam- intensifies bubble coalescence resulting in a non-linear relation-
eter is further enlarged from 0.15 to 0.38 m. The larger increase in ship of gas holdup and superficial gas velocity. Despite of the same
the first case might be referred to occurring wall effects which trends the magnitudes of the results of the present study differ
affect gas holdups in columns of diameters less than 0.15 m. widely from the ones presented by Krishna et al. (2001). This can
Dropping holdups in larger diameter columns occur according to be attributed to the initial liquid height of 1 m which was main-
Krishna et al. (2001) due to larger liquid circulations in columns tained by Krishna et al. (2001) and resulted in significant lower col-
of greater diameters. umn aspect ratios than in this publication. Another experimental
The measurements of this study show a similar trend of gas condition contributing to higher holdups is a sparger with smaller
holdup with respect to column diameter like the data presented orifice openings (0.5 mm) used by Krishna et al. (2001). The differ-
by Krishna et al. (2001). Holdups tend to decrease by about 1.5– ence in gas holdups of both studies is about the same magnitude as
2% if the column diameter is increased from 0.16 to 0.30 m diam- the difference of H/D ratios. As the column dimensions with
eter. Surprisingly it is found that holdup continuingly decreases if respect to diameter and spargers do not exactly match, this of
the diameter of the column is further increased to a value of 0.33 m course remains to be a speculative relationship. Anyhow, gas
(Fig. 12). holdup is defined as in Eq. (2-1). When aspect ratios are decreased
This behavior can be explained as the aspect ratio of the 0.33 m by lowering the liquid level, the total volume also decreases. If the
diameter column is of the same magnitude as the 0.16 m diameter same amounts of gas are introduced to achieve the same superfi-
column, while the 0.3 m diameter column is of lower H/D ratio. cial gas velocity the fraction of gas in the column will be higher

Table 7
Experimental conditions of literature studies on diameter influence on gas holdup.

Author D (m) ug (m/s) Gas/liquid


H/D (–) ul (m/s)
Wilkinson et al. (1992) 0.15, 0.23 0–0.3 N2/H2O, n-Heptane, mono-ethylene glycol
8, 5.22 –
Forret et al. (2003) 0.15, 0.40, 1.0 0.05–0.2 Air/H2O
>4 –
Krishna et al. (2001) and Krishna 0.1, 0.174, 0.19, 0.38, 0.63 0–0.866 Air, He, Ar, SF6/H2O, tetradecane, paraffin oil,
and Ellenberger (1996) 0.8–13 – polyacrylamide solutions
Botton et al. (1978) 0.02, 0.075, 0.25, 0.48 0–14 Air/H2O, aqueous glycol solutions,
16, 57.33, 8.8, 4.16 0–0.025 H2O + surface active agent
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 97

because of the same volume of gas in a smaller volume of liquid.


There might be two concurring effects of sparger and column
aspect ratio which should be further studied in detail.
A clear influence of column diameter on gas holdup can also be
seen if acetone or cumene is employed as the liquid phase. Because
of safety reasons cumene could not be used in the 0.3 m diameter
glass column. Acetone was not used in the steel column because
the seals of some equipment devices were not acetone resistant.
Therefore column diameter effects on gas holdups in acetone are
examined in both glass columns, while cumene was examined in
the 0.16 m diameter glass column and the larger steel column. A
larger decrease in gas holdups in both acetone (Fig. 13) and
cumene (Fig. 14) especially at higher superficial gas velocities of
up to 4% is noted as column diameter increases. Additionally mea-
surements of holdups in acetone done by Öztürk et al. (1987) in a
0.095 m diameter column are plotted in Fig. 13. Despite of the
smaller column diameter used for their measurements the results
obtained are within the range of the presented holdups of this
study obtained in a 0.16 m diameter column. Öztürk et al. (1987) Fig. 14. Influence of column diameter on nitrogen holdup in cumene.
used a single orifice sparger of 3 mm diameter and measured gas
holdups by comparing the ungassed liquid height (which was
0.85 m, resulting in an aspect ratio of 8.95) to the gassed liquid
level. Keeping the discussion above about column diameter and
aspect ratio in mind one would expect higher holdups in the
0.095 m diameter column used by Öztürk et al. (1987) compared
to the present results. A possible explanation could be the less
effective single orifice sparger used by the authors which possibly
distributes bubbles non-uniformly over the column cross section
resulting in a longer sparger inlet zone and thus a lower overall
gas holdup.
Almost an identical behavior is observed when nitrogen hold-
ups in cumene are compared with measurements by Matsubara
et al. (2010), who examined gas holdups in cumene in a column
of 0.3 m diameter (Fig. 14). Matsubara et al.’s (2010) results are
about 2% higher relative to the results obtained in this investiga-
tion. This difference might be attributed to a lower H/D aspect ratio
of their column compared to the 0.33 m diameter steel column.
According to the authors their column had an aspect ratio of 5, Fig. 15. Variation of superficial liquid velocity, 0.16 m diameter glass column.
related to aerated liquid height. Consequently the unaerated aspect
ratio is even lower. As stated and shown above lower aspect ratios
cause higher holdups. Anyhow Matsubara et al.’ (2010) results change in column diameter from 0.16 to 0.33 m seems not to influ-
show a similar trend like the results obtained in the 0.16 and ence the point of regime transition.
0.33 m diameter column with a change in slope at about A parameter not yet discussed while comparing the present
0.04 m/s superficial gas velocity. This change indicates the begin- results with literature data is the superficial liquid velocity.
ning of flow regime transition as pointed out earlier (Fig. 8). A Semibatch operation is often encountered in publications, but is
almost never applied as the mode of operation of industrial pro-
duction units. Nevertheless, the number of publications dealing
with this topic is limited as well as publications which examined
organic solvents. Usually low liquid velocities are found in bubble
columns because their main purpose is to carry out slow multi-
phase reactions. It is generally imaginable that co-current flow of
liquid and gas tends to reduce and countercurrent flow increases
gas holdup as bubbles are either accelerated by liquid motion
(co-current) or decelerated (countercurrent). As liquid velocities
are low to achieve residence times in the magnitude of hours, its
influence on holdup is often thought to be negligible. Akita and
Yoshida (1973) examined the influence of liquid velocity on gas
holdup in a 0.152 m diameter column and found no relationship
between these parameters as gas holdup did not change with
superficial liquid velocity. On the other hand, Shah et al. (2012)
noted a slight decrease of gas holdup in an empty and packed bub-
ble column of both 0.29 m diameter with an aspect ratio of 6.9.
Liquid velocities were varied up to 0.002 m/s in countercurrent
operation of gas and liquid. The decrease in gas holdup was attrib-
uted to an increase in friction force between liquid and gas which,
Fig. 13. Influence of column diameter on nitrogen holdup in acetone. according to the authors, enhances bubble coalescence.
98 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

Measurements of gas holdup with variation of superficial liquid Table 8


velocities in this study seem to confirm that there is no influence of Relative change of liquid properties with temperature, reference 20 °C.

liquid velocity on gas holdup within the range of parameters stud- DT (°C) 30 55
ied and the corresponding accuracy of measurement. Fig. 15 shows H2O
the results with consideration of liquid velocity in the 0.16 m Dq (%) 1.0 2.3
diameter glass column. Fig. 16 shows results of the greater 0.3 m Dl (%) 45 62
diameter glass column and Fig. 17 the effect of superficial liquid Dr (%) 8.10 14.86

velocity on gas holdup in the steel column. Cumene


It is evident that there is no distinct relationship between gas Dq (%) 2.7 5.0
Dl (%) 31.64 46.83
holdup and superficial liquid velocity observable at all column Dr (%) 10.7 21.4
dimensions and liquids studied. Variations of holdups with liquid
velocity are within the measurement errors. Hills (1976) men-
tioned that if the applied superficial liquid velocity is low com-
pared to the bubble rise velocities, no impact of liquid velocity
on gas holdup is expected as the acceleration of the bubbles in
co-current operation of both phases will be negligible. A compar-
ison between the bubble swarm velocities in Fig. 6 with the applied
liquid velocities shows that bubble swarm velocities are up to 280
times higher than the liquid velocities. Therefore it is not surpris-
ing that a potential dependence of both parameters will not be
found at these operating conditions, although they are realistic
with respect to liquid circulation rates in production plants.

Fig. 18. Influence of temperature on gas holdup.

Influence of temperature

The influence of temperature mainly affects liquid viscosity,


density and surface tension. As all three properties change with
temperature, however with different magnitudes as shown in
Table 8, it is thought to be difficult to isolate the effect of one prop-
erty on gas holdup due to a temperature increase. The property
most affected by temperature is liquid viscosity (Table 8, DT is
the temperature difference).
Smaller bubbles and slightly reduced bubble coalescence at
lower viscosity are expected. Surface tension and liquid density
Fig. 16. Variation of superficial liquid velocity, 0.30 m diameter glass column. changes only a little with rising temperature in the parameter
range examined here. Fig. 18 shows results obtained in the steel
column for liquid phase temperatures of 20, 50 and 75 °C of water
and 20 and 75 °C of cumene.
No clear relationship between temperature or rather liquid vis-
cosity and gas holdup is observable. If cumene was used the mea-
sured holdups did not deviate from each other at both temperature
levels. As water becomes less viscous with rising temperature an
increase of holdup from 20 to 50 and 75 °C would be expected.
Fig. 18 shows a more diffuse behavior because holdups measured
at 50 °C are lower than the ones at 20 °C while the results at
75 °C are higher than at 20 °C. These differences can be related to
impurities present in the liquid phase and the accuracy of mea-
surement. Kulkarni and Joshi (2005) stated that the results with
respect to viscosity obtained so far are contradictory, ranging from
no-influence to slight influence on bubble size with rising viscosity.
The observation of smaller holdups with rising viscosity was
already mentioned earlier where gas holdup measurements of
Wilkinson et al. (1992) in mono-ethylene glycol are compared with
water. In addition, Urseanu et al. (2003) found a significant
increase in gas holdups with lower liquid viscosities for high vis-
Fig. 17. Variation of superficial liquid velocity, 0.33 m diameter steel column. cous media (0.07–0.55 Pa s). Obviously the cited authors used
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 99

liquids far more viscous than the ones employed in this study.
Although a notable decrease in viscosity occurs at the conditions
examined, compared to the viscosity range studied by Urseanu
et al. (2003) or reviewed by Kulkarni and Joshi (2005) these differ-
ences appear to be negligible.
However this finding is of importance for the experiments
reported here because no possibility to cool the liquid was
installed in all three experimental facilities. As the pump always
introduces some heat into the column it was not possible to keep
the liquid temperature exactly constant at the desired level. The
temperature increased at a rate of about 1 K/10 min which was
exactly the time needed to acquire one measurement point.
Approximately 6–8 K temperature difference should be considered
to complete one experimental run. As discussed above this compli-
cation does not influence the results of this study as the changes in
gas holdup with rising temperature are small.

Influence of pressure
Fig. 19. Pressure effect on gas holdup, N2/H2O.
Often production of chemicals in multiphase reactors takes
place at elevated pressure. However, the effect of gas density on
gas holdup at higher pressure has not been studied as extensive
as one would expect it considering the importance of this param-
eter for bubble column design. It is most generally agreed that
gas holdup rises if pressure is increased. This has been experimen-
tally verified by Wilkinson and v. Dierendonck (1990), Letzel et al.
(1999) and Clark (1990) to name a few. A brief survey of other
studies can be found in Rollbusch et al. (2013b). The reasons for
increased holdups at elevated pressure can be found in the forma-
tion of smaller bubbles (Schäfer et al., 2002) because of enhanced
bubble breakup and less buoyancy force as a result of a lower dif-
ference in phase densities. Also for operation at higher gas
throughputs the point of regime transition is shifted to higher
superficial gas velocities (Wilkinson et al., 1992) because less large
bubbles form at these conditions.
In this study 4 different pressure levels (0.1, 1.0, 1.85 and
3.6 MPa) were investigated for the system nitrogen/deionized
water and 3 levels (0.1, 1.85 and 3.6 MPa) for nitrogen/cumene
in the 0.33 m diameter bubble column. Superficial gas velocities
Fig. 20. Pressure effect on gas holdup, N2/cumene.
were limited to a maximum of 0.05 m/s because low gas holdups
were of interest for this study. The superficial gas velocities were
calculated at the corresponding operating pressure. Generation of
lower holdups was also necessary to test some of the developed Letzel et al. (1999) also measured gas holdups at similar experi-
measurement techniques in this specific project. A laser endo- mental conditions, except the smaller column diameter, and found
scopic measurement technique developed by ILA (Intelligent no influence of pressure on holdup below 0.05 m/s superficial gas
Laser Applications GmbH, Jülich) was used to characterize bubble velocity. Further comparison with literature data is difficult as
size and velocity. High holdups or high bubble loads would have most results are focused on higher superficial gas velocities or
permitted the use of this measurement technique as it is based were measured in columns of very different geometry.
on evaluation of photographs. On the other hand it is quite difficult Fortunately Weber (2002) published two gas holdup data points
to establish higher superficial gas velocities at this scale and oper- of a commercial cumene oxidizer bubble column (diameter:
ating condition. The use of gas bundles was not feasible because 4.6 m, p: 0.7 MPa) which can be extracted and compared with
the necessary operating time of the tested tomographic measure- the results of this study (Fig. 21). If the 1.85 MPa points are linearly
ment device was about 10 min for one operating point and the extrapolated, which is justified as the column was operated in the
needed amount of gas for one complete experimental run cannot homogeneous regime and the extrapolation does not exceed the
be provided by gas bundles. limits of expected homogeneous flow, one observes that both
The effect of pressure on gas holdup is shown in Fig. 19 for holdup values of the industrial plant lie above the measurements
nitrogen holdups in deionized water and in Fig. 20 for holdups of at 0.1 MPa and slightly beneath those obtained at 1.85 MPa.
nitrogen in cumene. Keeping the earlier discussion about the data presented in Fig. 20
Gas holdup also seems to be a function of pressure at these con- in mind mainly three aspects can be identified for the observed
ditions which is contrary to the findings published by Pohorecki variation in holdups. Obviously the pressure is 1.05 MPa lower as
et al. (1999) and Pohorecki et al. (2001) who conducted studies during the experiments presented here which should result in
in a similar column (height: 4 m, diameter: 0.3 m) and comparable lower holdups. Another point is that the industrial column is much
operating conditions to study the effect of pressure on holdup in larger in diameter than the facility used here and that the cumene
water and cyclohexane. They found no dependence between used in the production plant should be considered as a reaction
holdup and pressure at all and mentioned that besides liquid prop- mixture mainly consisting of cumene. As pointed out earlier the
erties the superficial gas velocity is mainly affecting holdup values. cumene used in this study had a purity of 99% according to the
100 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

is noted. The relative holdup increase in water from 0.1 to 3.6 MPa
is about 500% while the relative increase in cumene is about 125%.
As reasons for this mainly two arguments can be identified. First,
the primary bubble size at atmospheric conditions in cumene or
any other organic liquid is smaller than in water because of the dif-
ferent liquid properties which influence bubble formation, coales-
cence and breakup (see the above discussion on liquid properties
for details). The effect of increased gas density on bubble size is
therefore less pronounced in organic liquids than in water. On
the other hand the decrease in surface tension if pressure is
increased from 0.1 MPa to 3.6 MPa listed in Table 9 is more distinct
in water, about 6.2%, than in cumene, which is about 3.1%. As sur-
face tension is directly influencing bubble breakup and coalescence
its relative change with pressure might also account for different
rates of holdup increase due to pressure.

Axial evolution and radial distribution of gas holdup


Fig. 21. Comparison of own measurements with industrial plant data at 0.7 MPa
published by Weber (2002). Most publications focus on integral or radial holdup measure-
ments. Only a few describe the axial distribution of gas holdup
along the column height. Information about axial holdup distribu-
product data sheet. As all three contributions to the deviations
tion can be vital to characterize sparger inlet zones, as sparger
interact with each other the lower pressure of the oxidizer might
properties such as orifice diameter and free surface area affect
be considered as the main reason. Furthermore the gas sparger of
holdup and size of this region, or the effect of internals on gas
the cumene oxidizer is oriented downwards (Weber, 2006) which
holdup. Pressure reduces along the column height because the liq-
may affect the initial bubble movement. Anyway, Fig. 21 proves
uid head reduces. Deckwer (1976) already pointed out to correct
that the conditions examined during the present experiments are
the superficial gas velocity according to the liquid height because
realistic with respect to conditions found in industrial production
of this circumstance. Of course reduced pressure interacts with
units.
bubble properties which are known to affect gas holdup.
Despite the lack of available experimental data some conclu-
Consequently at least a minor increase in gas holdup with column
sions can be drawn from the figures above and compared with
height is expected. Brauer (1971) defined three zones of varying
argumentations derived from other studies. It has already been sta-
holdup magnitudes. According to Brauer (1971) a sparger inlet
ted that the measured holdups increase with rising pressure in
zone near the bottom of the column with lowest holdups is fol-
water and cumene as well. This is the result of increased bubble
lowed by a zone in which bubble breakup and coalescence are in
breakage which leads to the existence of smaller bubbles at ele-
equilibrium in the middle of the column. Holdups rise until the
vated pressure than compared to atmospheric conditions
second zone is reached. From there gas holdup is constant until
(Idogawa et al., 1986, 1987; Jekat, 1975; Jiang et al., 1995; Letzel
the third one begins. This zone is near the top of the column where
et al., 1997). Bubble breakup at pressures above atmospheric might
gas disengages and highest holdups are to be found.
be enhanced because of a more pronounced propagation of insta-
Experimental evidence for this behavior is given by Jin et al.
bilities at the phase interface (Letzel et al., 1997). Of course buoy-
(2005) who measured axial holdup profiles with pressure differ-
ancy force reduces significantly as gas density increases with
ence and gamma-ray devices in a 6.6 m high bubble column of
pressure (Table 2 lists nitrogen densities for conditions established
0.3 m diameter. Water or acetic acid was employed as the liquid
here) which results in slower bubble rise velocities and therefore
phase and the pressure was as high as 1.0 MPa. Jin et al. (2005)
higher bubble residence times. Liquid surface tension also lowers
observed a sharp ascent of holdups with column height and the
slightly with increasing pressure (Table 9) and promotes bubble
forming of a foam layer at the top of the column. However the
breakup. Of course one should not forget that during the experi-
authors established very high superficial gas velocities between
ments presented liquid impurities might play a special role. The
0.1 and 0.4 m/s. Kumar et al. (1997) examined axial holdups with
bubble column was cleaned and dried as well as possible after
a gamma-CT device and noted increasing holdups along the col-
switching liquids (water or cumene) but it is always possible that
umn height. They attributed this result to the formation of larger
remnants of the used liquids remained inside the facility as it is
primary bubbles at the sparger which breakup as they travel to
quite difficult to completely clean experimental setups of this
the top of the column. Consequently more bubbles with smaller
dimension. As noted earlier, the cumene used during this experi-
diameter will be found at higher elevations than at the sparger
ments had a purity of 99% by delivery.
causing higher holdups.
Comparing the data in Fig. 19 with the ones of Fig. 20 a larger
The results presented here show a similar trend. The lowest
relative increase of holdups with pressure in water than in cumene
holdups are found near the bottom e.g. sparger of the column. At
0.1 MPa (Fig. 22) a zone with slightly increasing gas holdup can
be observed until 2.83 m of column height are reached. This
Table 9 increase lies within the error of measurement and should therefore
Measured surface tensions of cumene and water at various pressures and 35 °C, data
be treated carefully. Beyond 2.83 m liquid level a sudden decrease
provided by Eurotechnica GmbH.
of holdup occurs. The reason for this might be a significant reduc-
p (MPa) Deionized water (N/m) Cumene (N/m) tion of pressure at atmospheric conditions due to less liquid head
0.10 0.0715 0.026 which causes increased bubble coalescence. At the top of the col-
1.10 0.0698 – umn (3.88 m) a sharp increase of gas holdup is noted which is
1.85 0.0687 0.0255
due to gas bubble disengagement and the forming of a foam like
3.60 0.0671 0.0252
layer.
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 101

Fig. 22. Gas holdups along the column height, N2/H2O, p = 0.1 MPa. Fig. 24. Gas holdups along the column height, N2/cumene, p = 0.1 MPa.

tend to grow because of the reduced liquid head and therefore


occupy more volume at a constant number of bubbles. A foam
layer is expected to exist at the top of the column because this
was observed during the measurements in the glass column of
0.16 m diameter.
Surprisingly a different result is obtained at pressurized condi-
tions (Fig. 25). After the expected increase of gas holdup above
1.16 m it seems to stay constant within the accuracy of measure-
ment until 2.83 m are reached. After that a sharp increase occurs
before the measured holdups tend to decrease at the gas disen-
gagement zone. A possible explanation might be the earlier obser-
vation of a strong tendency of foaming in cumene at the top of the
column. At atmospheric conditions this foam might contribute to
higher measured gas holdups at the disengagement zone while
the foam layer might collapse at pressurized conditions resulting
in an abrupt decrease of holdup.
At this point some interesting remarks about radial holdup pro-
files seem appropriate. Through cooperation with the
Helmholtz-Center Dresden-Rossendorf it is possible to compare
Fig. 23. Gas holdups along the column height, N2/H2O, p = 3.6 MPa. gas holdups measured by pressure difference sensors with the ones
measured with a wire-mesh sensor and a gamma-CT (see Sectio
n ‘Experimental facilities’ for details) in the 0.33 m diameter col-
If the results at atmospheric conditions are compared to results umn. The results are shown in Fig. 26.
obtained at 3.6 MPa (Fig. 23) one observes that the zone between One can see that all three methods of measurement deliver
1.63 and 2.83 m consists of constant holdups and that the coalesc- comparable results. Deviations occur mainly due to the fact that
ing partition above 2.83 m is missing as gas holdups steadily both WMS and Gamma-CT deliver local holdups while the pressure
increase beyond 2.83 m liquid level. Because system pressure
influences bubble breakup, as discussed in the previous chapter,
the breakup rate is possibly faster than the coalescence rate and
therefore more small bubbles are present in the column which
enhances gas holdup. The foam like layer at the top of the column
might also be increased because a larger number of bubbles disen-
gage. A missing coalescence zone between 2.83 m and 3.3 m liquid
height might be explained by a reduction of liquid head in conjunc-
tion with the effects of elevated pressure. At atmospheric condi-
tions bubbles tend to grow and coalesce as they travel upward
the column because of a lower liquid head. At pressurized condi-
tions the change in pressure due to less liquid head is low com-
pared to the overall pressure of 3.6 MPa. Despite of that bubbles
moving upward in the column might slightly grow but do not coa-
lesce. This means that a larger number of bubbles is present in the
column at pressurized than at atmospheric conditions which
causes the observed increase of gas holdup along the column axis.
The same result is obtained in cumene at atmospheric pressure
(Fig. 24). Gas holdup increases continuously toward the column.
Additionally this effect seems to be more pronounced at higher
superficial gas velocities. Nitrogen bubbles do not coalesce that
much in cumene compared to bubbles in water. Nevertheless they Fig. 25. Gas holdups along the column height, N2/cumene, p = 3.6 MPa.
102 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

Fig. 28. Comparison of gas holdup correlations by Akita and Yoshida (1973), Hikita
Fig. 26. Validation of gas holdups obtained by pressure difference measurements et al. (1980), Hughmark (1967), Joshi et al. (1998), Mersmann (1977), Reilly et al.
with wire-mesh sensor and gamma-CT measurements. (1986), Joshi and Sharma (1979), Wilkinson et al. (1992), Idogawa et al. (1987).

difference measurements shown here are overall holdups. The gen-


to 400% occur within the presented correlations. Even at low
eral difference between WMS and Gamma-CT is again caused by
superficial gas velocities in the homogeneous flow regime, where
differing water qualities. Obviously gas bubbles in water seem to
a linear dependency of holdup and gas throughput is expected,
concentrate in the middle of the column which causes a radial dif-
very large differences prevail. Because bubble column hydrody-
ference in gas holdup. This is quite the opposite behavior if com-
namics are very sensitive on column geometry, sparger design,
pared to Gamma-CT measurements done in cumene (Fig. 27). In
gas and liquid properties it is very difficult to identify suitable
cumene nearly no radial holdup profile exists. This is explained
equations for predicting gas holdups (and other hydrodynamic
by the existence of smaller bubbles which are evenly distributed
parameters as well). Some of the depicted correlations are based
along the radial coordinate than the bubbles formed in water
on the principles of dimensional analysis (Akita and Yoshida,
(which are of broader size distribution).
1973; Hikita et al., 1980; Idogawa et al., 1987). Wilkinson et al.
(1992) considered changing flow regimes and consequently large
Prediction of gas holdups and small bubble holdups. However whether the design equation
is based on engineering fundamentals like dimensional analysis
Precise prediction of gas holdup is essential for the design of or theoretical considerations it most commonly fails if it is used
bubble column reactors and contactors. Gas holdup directly deter- for setups other than it is derived from and matching experimental
mines reactor size and interfacial area and is furthermore con- and calculated results are to be considered as flukes. With the
nected to liquid backmixing and heat and mass transfer rates. points mentioned of the result discussion above it is obviously dif-
Thus gas holdup is one of the most important design parameters ficult enough to establish comparable experimental conditions as
and should be estimated as accurately as possible in order to avoid even water is not comparable without being extra cautious with
designs which might lead to ineffective reactor operation respect to impurities and general water qualities. On the other
(Rollbusch et al., 2013a). hand, correlations suited for the prediction of holdups in organic
A vast number of empirical correlations exist to calculate the material are very scarce.
amount of gas holdup. In addition some semi-theoretical equations From an industrial point of view an equation to predict gas
based on idealized model assumptions have also been published. holdup must not only be reliable with respect to accuracy of hold-
Fig. 28 shows an example calculation of gas holdup in the 0.33 m ups in water. Furthermore this correlation needs to be able to pre-
diameter column at atmospheric pressure and water as liquid dict gas holdup under consideration of column diameter, different
phase with various correlations. Obviously large deviations of up liquid properties (as water is most often not of interest for indus-
trial production plants) and of course gas density. Almost no corre-
lations exist which fulfill these requirements. Krishna et al. (2001)
screened available correlations with the same scope as this study
and identified two possible equations, namely Akita and Yoshida
(1973) and Zehner (1982, 1989).
Only the Zehner (1989) correlation is able to predict the trends
observed in the results presented here and by Krishna et al. (2001).
Despite of having the column diameter as an input parameter the
equation derived by Akita and Yoshida (1973) does not predict any
influence of column diameter. Interestingly, Zehner’s (1989) corre-
lation predicted a decrease of holdups with column diameter
which is about the same magnitude as observed in the present
experiments and even Krishna et al.’s (2001) results. Because of
that the correlation proposed by Zehner (1989) will be examined
more closely. Zehner’s (1989) correlation is based on an improved
circulation cell model originally suggested by Joshi and Sharma
(1979). The original model describes bubble column hydrodynam-
Fig. 27. Gamma-CT measurements in deionized water and cumene compared to ics as circulating cells of vertical alignment. Zehner (1982) adapted
pressure difference measurements. this model and substituted the circulation cells with crosswise
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 103

aligned roller cells. According to Zehner (1982), this has the advan- Table 10
tage that the centerline velocity of the liquid phase is always direc- Measured and calculated bubble velocities, pressure as indicated in brackets.

ted upwards and the liquid velocity near the wall is directed Deionized H2O
downwards which has been experimentally confirmed by several ub (m/s) measured 0.3681 (0.1 MPa) 0.3605 (5 MPa) –
authors (e.g. Wu and Al-Dahhan, 2001). The downwards moving ub (m/s) calculated 0.328 (0.1 MPa) 0.327 (1.85 MPa) 0.325 (3.6 MPa)
Deviation (%) 10.89 9.29 –
liquid decelerates and entrains some bubbles while the upwards
moving liquid contains bubbles moving in the opposite direction. Cumene
ub (m/s) measured 0.2664 (0.1 MPa) 0.2601 (2 MPa) 0.2567 (4 MPa)
As a result a difference in gas holdups occurs which causes a pres- ub (m/s) calculated 0.2667 (0.1 MPa) 0.265 (1.85 MPa) 0.263 (3.6 MPa)
sure difference which is relieved by pressure losses due to liquid Deviation (%) 0.11 1.85 2.39
movement. The resulting correlation to predict gas holdups (Eq.
(3-2)) is then based on the liquid centerline velocity ul0, which
can be calculated with Eq. (3-3), and the velocity of the largest compared with the ones calculated with Eq. (3-5) in Table 10. As
stable bubble which should according to Zehner (1989) be calcu- one can see Eq. (3-5) predicts bubble velocities of nitrogen in
lated with Eq. (3-4). cumene with outstanding accuracy. About 10% deviation between
u =u calculation and measurement of bubble velocities in water are
g bs
eG ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
   ffi ð3-2Þ obtained. As discussed earlier, water seems to be more difficult to
2=3
ug ul0
1þ4 ubs ubs
characterize than organic material with respect to coalescence
behavior and possible impurities or slightly different water quali-
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi ties are regarded as the reason for the larger deviations.
3 1 ql  qg If Eq. (3-5) is used to predict the bubble velocity and conse-
ul0 ¼ ug gD ð3-3Þ
2:5 ql quently subsets of the measured gas holdups within a given accu-
racy the Zehner (1989) correlation and the measurements are in

rgðql  qg Þ0:25 satisfactory agreement, which is shown and discussed below. The
ubs ¼ 1:55 ð3-4Þ data in Table 10 will be published in a later publication by one of
q2l our project partners together with a detailed description of the
endoscopic measurement method.
The above presented equations inherit all parameters which were
It is possible to predict gas holdups with Eq. (3-2) within a given
identified as important with respect to gas holdup during the exper-
accuracy for subsets of the experimental results presented here. It
imental runs. Included are superficial gas velocity, liquid density
was not possible to reproduce all experimental results with Eq.
and surface tension, reactor diameter and gas density to account
(3-2). A possible reason might be the presence of tracer substances
for the pressure influence on gas holdup. Unfortunately the calcu-
and therefore impurities which affect the coalescence behavior of
lated do not match the measured holdups. This is shown in
bubbles during the experiments.
Fig. 29. A general overestimation of the predicted holdups can be
Figs. 30 and 31 show that Eq. (3-2) is able to predict the
observed. The possible reason for this is the calculated bubble
decrease in holdup with column diameter in deionized water and
velocity ubs. For bubbles in water at atmospheric conditions a value
acetone. In addition the gas holdup at atmospheric conditions in
0f 0.25 m/s is predicted by the given Eq. (3-4). Zehner (1989) stated
cumene of the 0.33 m diameter column (Fig. 32) is also accurately
that this equation is taken from Mersmann (1977). Actually a
predicted by the proposed correlation. However it fails to predict
slightly different equation for the bubble velocity is found in
holdups in cumene for the 0.16 m diameter column. The reason
(Mersmann, 1977) with a prefactor of 2 instead of 1.55 (Eq. (3-5),
for this is the formation of a large foam layer during the experi-
r means surface tension of the liquid phase).
ments in the 0.16 m column. This effect is not considered by the

rgðql  qg Þ0:25 equations used to predict gas holdups and consequently the corre-
ub ¼ 2 ð3-5Þ lation underestimates nitrogen holdups in cumene. Larger devia-
q2l
tions occur when holdups are predicted in water because of
Measurements of bubble velocities were carried out at the possible impurities present in the experimental facility during
Technical University of Hamburg-Harburg and are listed and the measurements. On the other hand it is more difficult to

Fig. 29. Comparison of predicted holdups with measured values. Fig. 30. Parity plot measured and predicted holdups N2/H2O.
104 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

Fig. 31. Parity plot measured and predicted holdups N2/acetone. Fig. 34. Parity plot for various pressures, measured and predicted holdups N2/
cumene.

The modified Zehner correlation is also able to predict subsets


of holdups at higher pressures than atmospheric in organic liquids
and in deionized water. Figs. 33 and 34 show the corresponding
comparisons between prediction and experimental results. The
results at 3.6 MPa depicted in Fig. 33 are completely underesti-
mated. As previously discussed the addition of small tracer sub-
stances was necessary and additionally water qualities might
have changed due to the presence of surfactants. Consequently it
is hard to evaluate measurements done in deionized water and
to compare them with predictions. More important is the applica-
bility of the proposed equation for holdups in organic liquids at
elevated pressures. As can be seen from Fig. 34 the experimental
holdups in cumene at elevated pressure can be predicted within
reasonable accuracy by the modified Zehner correlation.

Conclusions

Fig. 32. Parity plot measured and predicted holdups N2/cumene. The effect of various operating and design parameters on gas
holdup in two phase bubble columns was experimentally verified
and a correlation was proposed to calculate holdups at the exam-
ined conditions. Studies were carried out in three columns of vary-
ing diameter and height to diameter ratios with deionized water,
acetone, cumene and aqueous ethanol and acetone solutions. It
was found that gas holdups decrease with increasing column
diameter and height to diameter ratio. Low superficial liquid veloc-
ities do not affect gas holdup whereas increased gas density dras-
tically increases holdup. The increase of holdups in deionized
water is higher than in cumene because of a larger initial bubble
size and a more pronounced reduction of surface tension due to
elevated pressure. The use of organic solvents as liquid phase
material has shown that decreased surface tension and liquid den-
sity results in higher holdups than in deionized water. The addition
of small amounts of aqueous ethanol and acetone solutions
increased holdups dramatically due to coalescence inhibition. A
comparison between the measured holdups of the aqueous solu-
tions and pure organic liquids revealed that aqueous solutions
are not suitable as substitutes for organic substances. Regarding
the effect of temperature no dependency was found. This is mainly
Fig. 33. Parity plot for various pressures, measured and predicted holdups N2/H2O.
because liquids of low viscosities were examined and no effect of
decreasing viscosity due to higher temperatures was observed. It
was found that holdups slightly increase with column height and
measure holdups at gas fluxes of low magnitude which is the rea- that three zones along the column axis can be defined. A sparger
son for larger deviations between experiment and prediction at inlet zone, a zone of near constant gas holdup where equilibrium
very low superficial gas velocity. between breakup and coalescence exists and a gas disengagement
P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106 105

zone were identified. To predict gas holdups a modified form of the Forret, A., Schweitzer, J.M., Gauthier, T., Krishna, R., Schweich, D., 2003. Influence of
scale on the hydrodynamics of bubble column reactors: an experimental study
Zehner (1989) correlation is proposed. It was shown that this equa-
in columns of 0.1, 0.4 and 1.0 m diameters. Chem. Eng. Sci. 58, 719–724.
tion is able to predict the effect of column diameter, liquid proper- Grund, G., Schumpe, A., Deckwer, W.D., 1992. Gas-liquid mass transfer in a bubble
ties and pressure on gas holdup at conditions studied here. column with organic liquids. Chem. Eng. Sci. 47, 3509–3516.
To further validate the ability to reliably predict gas holdups Hampel, U., Bieberle, A., Hoppe, D., Kronenberg, J., Schleicher, E., Sühnel, T.,
Zimmermann, F., Zippe, C., 2007. High resolution gamma ray tomography
results at higher superficial gas velocity will be necessary. The scanner for flow measurement and non-destructive testing applications. Rev.
parameter range of this study was suited to chemical processes Sci. Instrum. 78.
operating at low superficial gas velocities in the homogeneous flow Hikita, H., Asai, S., Tanigawa, K., Segawa, K., Kitao, M., 1980. Gas hold-up in bubble
columns. Chem. Eng. J. 20, 59–67.
regime. If the proposed correlation is used to predict holdups in the Hills, J.H., 1976. The operation of a bubble column at high throughputs: I. Gas
heterogeneous regime one should be cautious. Nevertheless the holdup measurements. Chem. Eng. J. 12, 89–99.
proposed correlation here relies on a simplified flow model, bubble Hughmark, G.A., 1967. Holdup and mass transfer in bubble columns. Ind. Eng.
Chem. Process Des. Dev. 6, 218–220.
and liquid centerline velocities and does not inherit any fitting Idogawa, K., Ikeda, K., Fukuda, T., Morooka, S., 1986. Behavior of bubbles of the air-
parameters. Therefore it seems promising to predict holdups using water system in a column under high pressure. Int. Chem. Eng. 26, 468–474.
the suggested correlation. Idogawa, K., Ikeda, K., Fukuda, T., Morooka, S., 1987. Effect of gas and liquid
properties on the behavior of bubbles in a column under high pressure. Int.
Besides validating correlations the generated experimental Chem. Eng. 27, 93–99.
results, especially the axial distribution of gas holdups, might be Jakobsen, H.A., Lindborg, H., Dorao, C.A., 2005. Modeling of bubble column reactors:
useful to validate CFD models. This is of special interest as data progress and limitations. Ind. Eng. Chem. Res. 44, 5107–5151.
Jekat, H., 1975. Messung von Blasengrößenverteilungen in Druckblasensäulen im
measured at processing conditions and technical scales to validate
Bereich von 1 bis 100 bar. Fachbereich für Maschinenwesen, Technical
models are hard to find. University of Munich, Munich.
It was pointed out additionally that bubble column hydrody- Jiang, P., Lin, T.J., Luo, X., Fan, L.S., 1995. Flow visualization of high pressure (21 MPa)
namics are only comparable if identical experimental setups are bubble column: bubble characteristics. Chem. Eng. Res. Des. 73, 269–274.
Jin, H., Yang, S., Guo, Z., He, G., Tong, Z., 2005. The axial distribution of holdups in an
used. Therefore it is extremely difficult to compare own results industrial-scale bubble column with evaluated pressure using c-ray
with literature data. Even if the column dimension and the liquid attenuation approach. Chem. Eng. J. 115, 45–50.
phase are identical, deviations in the sparger design hamper com- Joshi, J.B., Sharma, M.M., 1979. A circulation cell model for bubble columns. Trans.
Inst. Chem. Eng. 57, 244–251.
parability. It seems not promising to compare holdups in water Joshi, J.B., Parasu Veera, U., Prasad, C.V., Phanikumar, D.V., Deshpande, N.S., Thorat,
because water qualities differ too much and are sensitive to surfac- B.N., 1998. Gas hold-up structures in bubble column reactors. Proc. Indian
tants. Another point is that water as liquid phase is mostly not of National Sci. Acad. 64A, 441–567.
Kantarci, N., Borak, F., Ulgen, K.O., 2005. Bubble column reactors. Process Biochem.
interest for industrial needs besides water treatment and biological 40, 2263–2283.
production processes. Organic liquids are processed at pressurized Keitel, G., Onken, U., 1982. Inhibition of bubble coalescence by solutes in air/water
conditions and therefore future experiments should concentrate on dispersions. Chem. Eng. Sci. 37, 1635–1638.
Kemoun, A., Cheng Ong, B., Gupta, P., Al-Dahhan, M.H., Dudukovic, M.P., 2001. Gas
this subject. However the use of organic material requires elabo- holdup in bubble columns at elevated pressure via computed tomography. Int. J.
rate safety measures and the operation of pressurized vessels Multiph. Flow 27, 929–946.
makes things not easier as a certain infrastructure is required to Krishna, R., Ellenberger, J., 1996. Gas holdup in bubble column reactors operating in
the churn-turbulent flow regime. AIChE J. 42, 2627–2634.
run them.
Krishna, R., Wilkinson, P.M., Van Dierendonck, L.L., 1991. A model for gas holdup in
bubble columns incorporating the influence of gas density on flow regime
transitions. Chem. Eng. Sci. 46, 2491–2496.
Krishna, R., Dreher, A.J., Urseanu, M.I., 2000a. Influence of alcohol addition on gas
Acknowledgements hold-up in bubble columns: development of a scale up model. Int. Commun.
Heat Mass Transfer 27, 465–472.
The ‘‘Low carbon dioxide emitting chemical processes for future Krishna, R., Urseanu, M.I., Dreher, A.J., 2000b. Gas hold-up in bubble columns:
influence of alcohol addition versus operation at elevated pressures. Chem. Eng.
industries: Multiscale Modelling of Multi-Phase Reactors’’ project – Process. 39, 371–378.
Multi-Phase – is funded by the German Federal Ministry of Krishna, R., v. Baten, J.M., Urseanu, M.I., 2001. Scale effects on the hydrodynamics of
Education and Research (Grant: 033RC1102A). bubble columns operating in the homogeneous flow regime. Chem. Eng.
Technol. 24, 451–458.
The authors would like to thank Melanie Bothe (Technical
Kulkarni, A.A., Joshi, J.B., 2005. Bubble formation and bubble rise velocity in gas-
University Hamburg-Harburg) for providing the experimental liquid systems: a review. Ind. Eng. Chem. Res. 44, 5873–5931.
results listed in Table 10. Kumar, S.B., Moslemian, D., Duduković, M.P., 1997. Gas-holdup measurements in
bubble columns using computed tomography. AIChE J. 43, 1414–1425.
Letzel, H.M., Schouten, J.C., Van Den Bleek, C.M., Krishna, R., 1997. Influence of
elevated pressure on the stability of bubbly flows. Chem. Eng. Sci. 52, 3733–
References 3739.
Letzel, H.M., Schouten, J.C., Krishna, R., van den Bleek, C.M., 1999. Gas holdup and
mass transfer in bubble column reactors operated at elevated pressure. Chem.
Akita, K., Yoshida, F., 1973. Gas holdup and volumetric mass transfer coefficient in
Eng. Sci. 54, 2237–2246.
bubble columns. Effects of liquid properties. Ind. Eng. Chem. Process Des. Dev.
Matsubara, H., Naito, K., Kuwamoto, H., Sakaguchi, T., 2010. Influence of operating
12, 76–80.
pressure on gas holdup and flow regime transition in a bubble column. J. Chem.
Bieberle, A., Härting, H.-U., Rabha, S., Schubert, M., Hampel, U., 2013. Gamma-ray
Eng. Jpn. 43, 829–832.
computed tomography for imaging of multiphase flows. Chem. Ing. Tech. 85,
McLaughlin, J.B., 2003. The hydrodynamics of two-phase flows with surfactants.
1002–1011.
Multiphase Sci. Technol. 15, 365–371.
Botton, R., Cosserat, D., Charpentier, J.C., 1978. Influence of column diameter and
Mersmann, A., 1977. Auslegung und Maßstabsvergrößerung von Blasen- und
high gas throughputs on the operation of a bubble column. Chem. Eng. J.
Tropfensäulen. Chem. Ing. Tech. 49, 679–691.
(Lausanne) 16, 107–115.
Nguyen, P.T., Hampton, M.A., Nguyen, A.V., Birkett, G.R., 2012. The influence of gas
Brauer, H., 1971. Grundlagen der Einphasen- und Mehrphasenströmungen. Verlag
velocity, salt type and concentration on transition concentration for bubble
Sauerländer, Aarau und Frankfurt am Main.
coalescence inhibition and gas holdup. Chem. Eng. Res. Des. 90, 33–39.
Camarasa, E., Vial, C., Poncin, S., Wild, G., Midoux, N., Bouillard, J., 1999. Influence of
Ohki, Y., Inoue, H., 1970. Longitudinal mixing of the liquid phase in bubble columns.
coalescence behaviour of the liquid and of gas sparging on hydrodynamics and
Chem. Eng. Sci. 25, 1–16.
bubble characteristics in a bubble column. Chem. Eng. Process. 38, 329–344.
Öztürk, S.S., Schumpe, A., Deckwer, W.D., 1987. Organic liquids in a bubble column:
Clark, K.N., 1990. The effect of high pressure and temperature on phase
holdups and mass transfer coefficients. AIChE J. 33, 1473–1480.
distributions in a bubble column. Chem. Eng. Sci. 45, 2301–2307.
Pohorecki, R., Moniuk, W., Zdrójkowski, A., 1999. Hydrodynamics of a bubble
Clift, R., Grace, J.R., Weber, M.E., 1978. Bubbles, Drops, and Particles. Academic
column under elevated pressure. Chem. Eng. Sci. 54, 5187–5193.
Press.
Pohorecki, R., Moniuk, W., Zdrójkowski, A., Bielski, P., 2001. Hydrodynamics of a
Deckwer, W.D., 1976. Non-isobaric bubble columns with variable gas velocity.
pilot plant bubble column under elevated temperature and pressure. Chem.
Chem. Eng. Sci. 31, 309–317.
Eng. Sci. 56, 1167–1174.
Deckwer, W.D., 1985. Reaktionstechnik in Blasensäulen, 1 ed. Salle+Sauerländer,
Frankfurt am Main.
106 P. Rollbusch et al. / International Journal of Multiphase Flow 75 (2015) 88–106

Posarac, D., Tekic, M.N., 1987. Gas holdup and volumetric mass transfer coefficient Shaikh, A., Al-Dahhan Muthanna, H., 2007. A review on flow regime transition in
in bubble columns with dilute alcohol solutions. AIChE J. 33, 497–499. bubble columns. Int. J. Chem. React. Eng. 5.
Prince, M.J., Blanch, H.W., 1990. Bubble coalescence and break-up in air-sparged Syeda, S.R., Reza, M.J., 2011. Effect of surface tension gradient on gas hold-up
bubble columns. AIChE J. 36, 1485–1499. enhancement in aqueous solutions of electrolytes. Chem. Eng. Res. Des. 89,
Reilly, I.G., Scott, D.S., De Bruijn, T., Jain, A., Piskorz, J., 1986. A correlation for gas 2552–2559.
holdup in turbulent coalescing bubble columns. Can. J. Chem. Eng. 64, 705–717. Tang, C., Heindel, T.J., 2006. Estimating gas holdup via pressure difference
Rollbusch, P., Becker, M., Ludwig, M., Grünewald, M., 2013a. Shortcut-Modellierung measurements in a cocurrent bubble column. Int. J. Multiph. Flow 32, 850–863.
von Blasensäulenreaktoren. Chem. Ing. Tech. 85, 1425–1425. Urseanu, M.I., Guit, R.P.M., Stankiewicz, A., van Kranenburg, G., Lommen, J.H.G.M.,
Rollbusch, P., Tuinier, M., Becker, M., Ludwig, M., Grünewald, M., Franke, R., 2013b. 2003. Influence of operating pressure on the gas hold-up in bubble columns for
Hydrodynamics of high-pressure bubble columns. Chem. Eng. Technol. 36, high viscous media. Chem. Eng. Sci. 58, 697–704.
1603–1607. Weber, M., 2002. Large bubble columns for the oxidation of cumene in phenol
Ruff, K., Pilhofer, T., Mersmann, A., 1976. Vollständige Durchströmung von processes. Chem. Eng. Technol. 25, 553–558.
Lochböden bei der Fluid-Dispergierung. Chem. Ing. Tech. 48, 759–764. Weber, M., 2006. Some safety aspects on the design of sparger systems for the
Ruzicka, M.C., Drahoš, J., Fialová, M., Thomas, N.H., 2001. Effect of bubble column oxidation of organic liquids. Process Saf. Prog. 25, 326–330.
dimensions on flow regime transition. Chem. Eng. Sci. 56, 6117–6124. Wilkinson, P.M., v. Dierendonck, L.L., 1990. Pressure and gas density effects on
Schäfer, R., Merten, C., Eigenberger, G., 2002. Bubble size distributions in a bubble bubble break-up and gas hold-up in bubble columns. Chem. Eng. Sci. 45, 2309–
column reactor under industrial conditions. Exp. Thermal Fluid Sci. 26, 595– 2315.
604. Wilkinson, P., Spek, A., van Dierendonck, L., 1992. Design parameters estimation for
Schlusemann, L., Zheng, G., Grünewald, M., 2013. Messung der Phasenverteilung in scale-up of high-pressure bubble columns. AIChE J. 38, 544–554.
Blasensäulen. Chem. Ing. Tech. 85, 997–1001. Wu, Y., Al-Dahhan, M.H., 2001. Prediction of axial liquid velocity profile in bubble
Shah, Y.T., Kelkar, B.G., Godbole, S.P., Deckwer, W.D., 1982. Design parameters columns. Chem. Eng. Sci. 56, 1127–1130.
estimations for bubble column reactors. AIChE J. 28, 353–379. Zahradnik, J., Fialova, M., Kastanek, F., Green, K.D., Thomas, N.H., 1995. Effect of
Shah, Y.T., Joseph, S., Smith, D.N., Ruether, J.A., 1985. On the behavior of the gas electrolytes on bubble coalescence and gas holdup in bubble column reactors.
phase in a bubble column with ethanol-water mixtures. Ind. Eng. Chem. Process Chem. Eng. Res. Des. 73, 341–346.
Des. Dev. 24, 1140–1148. Zehner, P., 1982. Momentum, mass, and heat transfer in bubble columns. 1. Flow
Shah, M., Kiss, A.A., Zondervan, E., Van Der Schaaf, J., De Haan, A.B., 2012. Gas model of bubble columns and liquid velocity. VT, Verfahrenstech. 16, 347–351.
holdup, axial dispersion, and mass transfer studies in bubble columns. Ind. Eng. Zehner, P., 1989. Multiphase flows in gas–liquid reactors. DECHEMA-Monogr. 114,
Chem. Res. 51, 14268–14278. 215–232.

You might also like