You are on page 1of 28

CFD simulation of a piolet scale bubble column with fixed and poly dispersity approach

Introduction
Gas–liquid contacting is an important unit operation and has wide applications in several chemical,
petrochemical, pharmaceutical and biochemical processes. The process is carried out using suitable
gas–liquid contactors [1-4]. The main equipment widely used to achieve this purpose are bubble column
reactors. Their Simple design without any moving mechanical parts makes them an interesting choice
for multiphase engineering application. Beside mechanical advantages, bubble columns provide large
contact area between phases which leads to an effective mixing and improved heat and mass transfer
characteristics [1-3, 5, 6].

Despite having a simple mechanical structure, they pose complex flow structure which is still poorly
understood [4, 7]. The complex behavior is due to various interrelated parameters such as local and
global flow pattern and holdup and turbulence which are connected to operating and design variables
such as pressure, temperature, gas flowrate, sparger design, column diameter, column height etc. [1].
So, to correctly design and scale up bubble columns, one should be able to predict this complex behavior
precisely. In general, experimental fluid dynamics (EFD) have been the main tool to study any
multiphase flow which involves extraction of a detailed knowledge of the flow pattern through
systematic experimental procedures and measurement techniques [2].

While experimental investigations are usually performed at small scales, the design must be done for
larger industrial scale. At larger scales, experimental data acquisition becomes more challenging and
expensive and in most cases the data obtained from small experimental scale cannot well represent the
industrial case. It is, therefore, essential to develop accurate computational models to understand such
complex interactions and to facilitate design and scale-up.

In recent decade, Computational Fluid Dynamics (CFD) became one of the essential tool in solving and
analyzing problems that involve fluid flow [8, 9]. Over the last few years considerable research effort
has been focused on obtaining CFD models and approaches that have good prediction capabilities with
minor computational costs. There have been number of approaches to numerically predict multiphase
behavior which can be found in books and literatures such as direct numerical methods (VOF, level set,
front tacking), Eulerain-Eulerain and Eulerain-lagrangian approaches.

In direct numerical simulations (DNS) the interface between separate phases are reconstructed and
tracked which makes them computationally expensive methods which cannot be easily applied to large
and industrial scale problems. At larger scales, an averaged description by the Eulerain-Eulerain model
is the most attractive solution.

1
Each phase will be treated as an Eulerain phase and conservation equations will be solved for each
phase separately. Instead of interface reconstruction, a weighted volume-fraction will be calculated
which represents the ensemble averaged probability of occurrence for each phase at a certain point in
time and space [9, 10]. Since in this averaging procedure, the small length scales associated with the
phase interface are eliminated, this application in large length scales become feasible. However, the
loss of information on the precise location of interfaces, which determines the exchange terms between
phases, requires additional closure models to determine bubble-liquid and bubble–bubble interactions
which is described by source/sink terms in the balance equations using closure models [9]. A closure
model is set of models that describes interactions such as heat, momentum and mass transfer between
phases. These closure models are of outmost importance since the predictive capability of Euler-Euler
approach is highly relies on a proper choice of closure relations [11]. Large part of the studies on
adiabatic bubbly flow have focused on the forces acting on the bubbles which are usually a function of
bubble size, turbulence and rates for bubble coalescence and breakup [8].

In this study, an Eulerain-Eulerain approach is considered to numerically investigate a piolet scale air
blown bubble column. The procedures to acquire the experimental data are explained and CFD
simulations are performed at homogenous region at different gas superficial velocities.

1. Experimental Data

1.1. Experimental Operating Condition


The experiments were performed on a pilot scale bubble column at fluid dynamic laboratory of
Politecnico di Milano. Figure 1 shows a scheme of the facility which consists a vertical pipe made of
transparent material (plexiglass) with internal and external diameter of 250 and 240 mm.

Figure 1. schematic view of the experimented bubble column and surrounding equipment

2
The column is filled with water and the air is distributed through a sparger at the bottom of the column.
The spargers are different in type and shape. Kulkarni et. al. [12-14] have investigated different sparger
and evaluate their performances and have classified different sparger into sieve plate sparger, wheel
sparger, radial sparger and spider sparger. The sparger in this experiment is a spider sparger with 2 rows
of 6 branches at different level. The bottom level is blocked with simple tape in order to have just one
active level. Figure 2 shows the experimented sparger. Each pipe will have 6 holes and 4 holes are
located at the centre of the sparger.

Figure 2. Sparger geomentry used in the study

The water head is 3 m above the sparger holes. In this set up the liquid is stalled, and air is distributed
upward and is discharged at the upper end of the column. The column is operating at constant
temperature of 298.15 K and atmospheric pressure.

1.2. Global Gas Holdup

The most important hydrodynamic parameter in any bubbly flow system with any configuration and
application, is gas holdup. It plays a vital role in bubble column transport processes and gas-liquid
contacting and has been topic of extensive studies [15]. Holdup indicates the volume fraction of gas
phase and is used to describe mean residence time of the gas phase in the vessel which is governed by
average bubble size, population of bubbles and bubble velocity. Holdup governs the velocity or flow
field, turbulence characteristics in the individual phases and the energy dissipation rates [16]. Higher
holdup means increase in the interfacial area between the gas and liquid and enhancement in heat and
mass transfer rate and it is mainly desired to maximize the holdup during an operation [15]. Global gas
holdup 𝜀𝑔 is simply defined as ratio of gas phase volume to total volume:
𝑉𝑔
𝜀𝑔 =
𝑉𝑡𝑜𝑡

In this experiment the gas holdup was measured visually by noting the difference between level with
and without aeration using the following equation:
𝐻𝑡𝑜𝑡 −𝐻𝑙
𝜀𝑔 = 𝐻𝑡𝑜𝑡

3
where 𝐻𝑙 is the initial height of the liquid in the column before gas intake and 𝐻𝑡𝑜𝑡 is the mixture height
after gas intake in the column. For each flowrate, the measurement is taken after 1 minute, to reach
steady state condition. Also due to a wavy free surface at the measurement point, a minimum and a
maximum height was recorded and final 𝐻𝑡𝑜𝑡 is the average of 𝐻𝑚𝑖𝑛 and 𝐻𝑚𝑎𝑥.
The gas holdup is obtained for different flowrate from 10 Nl/min to 300 Nl/min (values read on the
rotameter) to determine the transition point from homogenous to heterogenous regime. In this study the
homogenous regime is studied by choosing 6 different flowrates in this region as shown in Table 1
below.

Table 1. Flowrate specification considered in the study

Volumetric flow mass flow Superficial gas Holdup


rate velocity
(Qread) (Ug)
[lit/min] [kg/s] m/s -
10 0.0002343 0.003814 1.23
20 0.000468601 0.007628 2.54
30 0.000704243 0.011464 4.00
40 0.000942559 0.015344 5.40
50 0.001182642 0.019252 6.86
60 0.001429775 0.023275 8.11

1.3. Bubble Size Distribution (BSD)

Image analysis technique is applied to acquire BSD. Two regions were selected to take pictures and
sampling bubbles as shown in Figure 3. One region is near sparger which is from sparger holes to 5 cm
above the sparger and the other region is at the height of 2.5 m above the sparger which is developed
region. The developed region was chosen based on a previous study of Besagni et. al [17] which have
used the same bubble column configuration as used in this study but with a different sparger and liquids.
Also based on Thorat et al [18] at h/dc>5, the holdup profile is fully developed so BSD can be
considered approximately constant (h is height of column and dc is diameter of the column).

Figure 3. Regions to obtain BSD

4
The photographs were taken by illuminating the flow with uniform, diffused white light and capturing
the image with a Canon Camera (speed of shutter 1/4000 - 1/5000 s, focouse 2.8 - 4.0, ISO = 800). The
camera was fixed on a stand, approximately 8 cm away from the observed area. The test section of the
bubble column was located between the camera and light source that was placed behind a diffuser to
evenly distribute the light. To compensate for bubble distortion due to water and curvature of the pipe,
a box is placed around the desired sections and filled with water as done by [19, 20]. Figure 4 shows
the explained procedure.

Figure 4. camera and light source installation for image acquisition

The images were further processed with Photoshop Software in order to better distinguish bubble
boundaries. The bubbles were sampled using MATLAB code based on the approach developed by
Aloufi [21]. This method has been also used and approved by Besagni et. al. in different studies [17,
20, 22]. The shape of a bubble is not necessarily spherical but could be approximated as being an oblate
ellipsoid. Therefore, the volume and surface area of each bubble were calculated by considering the
bubble as an oblate ellipsoid. This assumption was also considered in different other studies [17, 19,
20, 22-24]. The BSD is drastically different in these two regions due to breakup and coalesces
phenomena along the bubble column so in analysis of bubble size in bubble columns, one must
distinguish between bubble-size distribution just after bubble formation at the sparger and further away
from the sparger [25]. Figure 5 shows the BSD for flowrate of 10 for both developed and sparger region.

0.08
Developed Region
0.07
Sparger Region
0.06

0.05
Frequency

0.04

0.03

0.02

0.01

0
11.25

12.75

14.25

15.75

17.25

18.75

20.25

21.75
0

12

13.5

15

18

21
0.75
1.5
2.25

3.75
4.5
5.25

6.75
7.5
8.25

9.75
10.5

16.5

19.5

22.5

db [mm]

Figure 5. BSD at the developed region and sparger for flow rate of 10

5
1.4. Mass Flow Rate Distribution at Sparger
At low flow rates, it was noticed that the bubble production in the sparger is not uniform. This
phenomenon occurs when the pressure of the flow is not enough to overcome the weight of the water
column and will create a non-uniformity in mass distribution along the sparger with de-active (or not
fully active) holes which might lead to a reverse flow of water inside the sparger. However, all holes
were fully active at higher flowrate (50 and 60) as can be seen in Figure 6. A deeper analysis was made
to understand this inhomogeneous behavior. The performed experiment allows to calculate the amount
of produced bubble in each hole and to calculate the relative productivity which hereafter will be called
eta (η). η will be used to consider each hole activity and will be used as a coefficient to determine the
mass flowrate coming out of each hole to set up a proper CFD case. Each branch of sparger is labeled
from 1 to 6. Figure 2 shows a view of the structure from the top.

Figure 6. masss flowrate inhomogenity at the sparger

At each flowrate, seven different high-quality videos were taken. One for each branch and one for the
central body holes. The videos are analyzed frame by frame. The high number of frame per second
allow to catch the bubble detachment from the sparger holes. For each video, at least 1200 frames are
examined, (i.e. 5 seconds). Then, the bubble produced per unit of time is calculated, the result obtained
are shown in Figure 7. The η index is calculated with the assumption that all produced bubbles have
the same volume and it is the amount of bubble produced for a specific hole divided by total number of
generated bubble at the sparger. With this index, it is also possible to make an estimation of the mass
flowrate and velocity of the gas passing through each hole. The results are listed in Table 2.

6
Figure 7. bubble generation for different holes per second for differennt superfcial velocity correspnd
to 10 to 50 lit/min

Table 2. η and velocity values for each sparger holes

η coefficient Velocity [m/s]


Branch Hole Ug [m/s] Branch Hole U [m/s]
0,0038 0,0076 0,0113 0,0154 0,0193 0,0038 0,0076 0,0113 0,0154 0,0193
1 a 0,044 0,030 0,027 0,028 0,025 1 a 1,73 2,36 3,15 4,43 5,08

b 0,044 0,030 0,033 0,028 0,025 b 1,73 2,36 3,84 4,43 5,08

c 0,044 0,030 0,033 0,028 0,025 c 1,73 2,36 3,84 4,43 5,08

d 0,044 0,047 0,033 0,028 0,025 d 1,73 3,71 3,84 4,43 5,08

e 0,044 0,030 0,027 0,028 0,025 e 1,73 2,36 3,15 4,43 5,08

f 0,044 0,030 0,027 0,028 0,025 f 1,73 2,36 3,15 4,43 5,08

2 a 0,044 0,030 0,033 0,028 0,025 2 a 1,73 2,36 3,84 4,43 5,08
b 1,73 3,71 3,84 4,43 5,08
b 0,044 0,047 0,033 0,028 0,025
c 1,73 3,71 3,84 4,43 5,08
c 0,044 0,047 0,033 0,028 0,025
d 1,73 3,71 3,84 4,43 5,08
d 0,044 0,047 0,033 0,028 0,025
e 1,73 3,71 3,84 4,43 5,08
e 0,044 0,047 0,033 0,028 0,025
f 1,73 3,71 3,84 4,43 5,08
f 0,044 0,047 0,033 0,028 0,025
3 a 0,59 2,36 3,84 4,43 5,08
3 a 0,015 0,030 0,033 0,028 0,025
b 0,59 2,36 3,84 4,43 5,08
b 0,015 0,030 0,033 0,028 0,025
c 0 2,36 3,15 4,43 5,08
c 0,000 0,030 0,027 0,028 0,025
d 0 0,89 3,15 4,43 5,08
d 0,000 0,011 0,027 0,028 0,025
e 0 0,89 1,05 4,43 5,08
e 0,000 0,011 0,009 0,028 0,025
f 0 0 1,05 2,95 3,66
f 0,000 0,000 0,009 0,019 0,018
4 a 0,59 2,36 3,84 4,43 5,08
4 a 0,015 0,030 0,033 0,028 0,025
b 0,59 2,36 3,84 4,43 5,08
b 0,015 0,030 0,033 0,028 0,025
c 0,59 0,89 3,84 4,43 5,08
c 0,015 0,011 0,033 0,028 0,025
d 0 0,89 3,15 4,43 5,08
d 0,000 0,011 0,027 0,028 0,025
e 0 0,89 3,15 4,43 5,08
e 0,000 0,011 0,027 0,028 0,025
f 0 0 1,05 2,95 5,08
f 0,000 0,000 0,009 0,019 0,025
5 a 0 0,89 1,05 4,43 5,08
5 a 0,000 0,011 0,009 0,028 0,025
b 0 0,89 1,05 2,95 5,08
b 0,000 0,011 0,009 0,019 0,025
c 0 0 1,05 2,95 5,08
c 0,000 0,000 0,009 0,019 0,025
d 0 0 1,05 1,11 5,08
d 0,000 0,000 0,009 0,007 0,025
e 0 0 0 0 5,08
e 0,000 0,000 0,000 0,000 0,025

7
f 0,000 0,000 0,000 0,000 0,018 f 0 0 0 0 3,66
6 a 0,015 0,030 0,027 0,028 0,025 6 a 0,59 2,36 3,15 4,43 5,08
b 0,015 0,030 0,033 0,028 0,025 b 0,59 2,36 3,84 4,43 5,08
c 0,015 0,011 0,027 0,028 0,025 c 0,59 0,89 3,15 4,43 5,08
d 0,000 0,030 0,027 0,028 0,025 d 0 2,36 3,15 4,43 5,08
e 0,000 0,011 0,027 0,028 0,025 e 0 0,89 3,15 4,43 5,08
f 0,000 0,011 0,027 0,028 0,025 f 0 0,89 3,15 4,43 5,08
Centre N 0,088 0,047 0,033 0,028 0,025 Centre N 3,47 3,71 3,84 4,43 5,08
E 0,088 0,047 0,033 0,028 0,025 E 3,47 3,71 3,84 4,43 5,08
S 0,088 0,047 0,033 0,028 0,025 S 3,47 3,71 3,84 4,43 5,08
W 0,088 0,047 0,033 0,028 0,025 W 3,47 3,71 3,84 4,43 5,08

2. CFD Modelling
An Eulerian-Eulerian approach which is implemented in Ansys CFX 18.2 software is used for
numerically investigate the bubble column. The detailed discussion for numerical setup are presented
in below sections.

2.1. Geometry and Mesh

The bubble column is simulated as a 3D vertical tube with the same configuration as experimental one.
The mesh size is 12 mm in horizontal direction and 12 to 25 mm in vertical direction with approximately
150,000 cells. Near the sparger region tetra cells are used while for other regions hexahedral cells
compose the mesh. The grid independency study is performed by Besagni et. al. [4] which have used
the same geometry, with different operating condition and sparger.

2.2. Inlet Mass Flow

Sparger holes are simulated as source points. For each source point a mass and velocity source are added
according to the evaluated η coefficient and hole velocity as discussed in section 1.4. Two different
cases are considered regarding mass flow distribution at the sparger. One would be non-uniform
distribution using η values and one would be uniform distribution in which each hole will have η value
of 0.025 and the velocity for each hole is set to be 0.24 m/s which is estimated terminal velocity of
bubbles in the studied bubble column.

2.3. Fluid Properties

Water properties are extracted at P=1 atm and T=25 °C. The air is considered to be an ideal and
incompressible gas. The pressure along the bubble column and along the path of the air flow is not
constant and depends on the bubble position and can be calculated as:

𝑃 = 𝑃𝑎𝑡𝑚 + 𝜌𝑤𝑎𝑡𝑒𝑟 𝑔ℎ

As we come up along the column, the water static pressure is reduced and at the outlet of the column
the pressure is atmospheric. So basically, density is not constant and average value is needed. In

8
simulations the value of density is calculated by considering value of pressure at the middle of the
column with h=1.5 m. With this pressure the density of the air is around 1.35 kg/m3. A surface tension
of 0.072 is considered for air-water system. Table 3 summarizes the fluids properties.

Table 3. Fluid property used in numerical calculations

Fluid Density (kg/m3) Viscosity (kg/ms) Surface tension (N/m)


Air 1.35 1.85 10-5 -
Water 998 8.9 10-4 -
Air-Water - - 0.072

2.4. Governing Equations


In any numerical simulation of fluid flow, a set of conservation of mass, momentum and the turbulence
equations will be solved for each phase and will be coupled through phase interaction models as shown
in below equation.

Mass conservation equation and momentum balance of phase i is given by

𝜕 →
(𝛼𝑖 𝜌𝑖 ) + 𝛻 · (𝛼𝑖 𝜌𝑖 𝑢𝑖 ) = 0
𝜕𝑡

𝜕 → →→ → 𝐷𝑟𝑎𝑔 𝑙𝑖𝑓𝑡
(𝛼𝑖 𝜌𝑖 𝑢𝑖 ) + 𝛻 · (𝛼𝑖 𝜌𝑖 𝑢𝑖 𝑢𝑖 ) = 𝛼𝑖 𝛻𝑝 + 𝛻𝜏𝑖 + 𝛼𝑖 𝜌𝑖 𝑔 + 𝐹𝑖 + 𝐹𝑖 + 𝐹𝑖𝑉𝑀 + 𝐹𝑖𝑤𝑎𝑙𝑙 + 𝐹𝑖𝑇𝐷
𝜕𝑡

where p is pressure, g is gravity acceleration, αi, ρg, μg , ui and τi are the void fraction, density, viscosity,
Drag
velocity and the stress tensor of the ith phase, Fi , Filift , FiVM , Fiwall , FiTD are drag, lift, virtual mass,
wall lubrication and turbulent dispersion force respectively. The sum of force component need to be
zero in order to achieve momentum conservation. The force components will be discussed later. In case
of turbulent flow, more sets of equations for turbulence parameter need to be solved. Commonly used
models for two-phase flows in Eulerian-Eulerian framework are k−ω and k−ϵ. In this study
SST k−ω model, which is obtained by a blending of k−ϵ and k−ω models, is used and conservation
equation for k and ω are as follow:

𝜕 𝜇𝑡𝑢𝑟𝑏
(𝛼𝐿 𝜌𝐿 𝑘𝐿 ) + 𝛻 · (𝛼𝐿 𝜌𝐿 𝐮𝐿 𝑘𝐿 ) = 𝛻 · (𝛼𝐿 (𝜇𝐿𝑚𝑜𝑙 + 𝐿 ) 𝛻𝑘𝐿 ) + 𝑃𝑘 − 𝑌𝑘 + 𝑆𝐿𝑘
𝜕𝑡 𝜎𝑘

𝜕 𝜇𝑡𝑢𝑟𝑏
(𝛼𝐿 𝜌𝐿 𝜔𝐿 ) + 𝛻 · (𝛼𝐿 𝜌𝐿 𝐮𝐿 𝜔𝐿 ) = 𝛻 · (𝛼𝐿 (𝜇𝐿𝑚𝑜𝑙 + 𝐿 ) 𝛻𝜔𝐿 ) + 𝑃𝜔 − 𝑌𝜔 + 𝑆𝐿𝜀 + 𝐷𝜔
𝜕𝑡 𝜎𝜔

The terms 𝑃𝑘 and 𝑃𝜔 represent the generation of k respective ω. The terms 𝑌𝑘 and 𝑌𝜔 represent the
dissipation of k and ω, respectively. The cross-diffusion term 𝐷𝜔 arises due to the blending. The

9
terms 𝑆𝐿𝑘 and 𝑆𝐿𝜀 are source terms, which represent the bubble induced turbulence modelling and will
be discussed later.

2.5. Bubble Forces


Eulerian- Eulerian approach provides an averaged description of flow phenomena with extra models to
describe interaction between separated phases. The interaction between phases can be momentum
(force), mass and heat exchange. Here we are considering an isothermal system without mass transfer
phenomena. So basically, the only phase interaction will be through momentum exchange to balance
the existing forces. To take account for momentum exchange a set of forces known as drag, virtual
mass, lift, wall, turbulent dispersion forces and turbulence exchange are considered.

There are no theoretical expressions for the mentioned models and experimental correlations are widely
used to describe each force and their related parameters. They are mainly derived using important
dimensionless numbers such as Reynolds number, Eötvös number and Morton number. Different
studies are conducted to experimentally derive correlations and they have been implemented in
commercial and non-commercial CFD software.

2.5.1. Drag Force


The drag force reflects the resistance opposing bubble motion relative to the surrounding liquid. The
corresponding gas-phase momentum source is given by

3
𝑭𝑑𝑟𝑎𝑔 = − 𝐶 𝜌 𝛼 |𝒖 − 𝒖𝐿 |(𝒖𝐺 − 𝒖𝐿 )
4𝑑𝐵 𝐷 𝐿 𝐺 𝐺

The drag coefficient CD depends strongly on the Reynolds number Re and for deformable bubbles also
on the Eötvös number Eo. A correlation distinguishing different shape regimes has been suggested
by Ishii and Zuber [26], namely

CD = max(CD,sphere , min(CD,ellipse , CD,cap )),

24 2 8
CD,sphere = (1 + 0.1Re0.75 ). CD,ellipse = √EoCD,cap =
Re 3 3

2.5.2. Wall Force


A bubble translating next to a wall in an otherwise quiescent liquid also experiences a lateral force
commonly called wall force and it has the general form of

2
𝑭𝑤𝑎𝑙𝑙 = 𝐶 𝜌 𝛼 |𝒖 − 𝒖𝐿 |2 𝒚
̂
𝑑𝐵 𝑊 𝐿 𝐺 𝐺

10
where 𝐲̂ is the unit normal perpendicular to the wall pointing into the fluid. The dimensionless wall
force coefficient CW depends on the distance to the wall y and is expected to be positive, so the bubble
is driven away from the wall. In the limit of small Morton number, the correlation for 𝐶𝑊 can be
considered as follow:
𝑑𝐵
𝐶𝑊 (𝑦) = 0.0217𝐸𝑜 × ( )2
2𝑦

It is derived from the data of Hosokawa et al [27].

2.5.3. Turbulent Dispersion Force


The turbulent dispersion force describes the effect of the turbulent fluctuations of liquid velocity on the
bubbles. Burns et. al. [28] derived an explicit expression by Favre averaging the drag force as:

3 𝛼𝐺 𝜇𝐿𝑡𝑢𝑟𝑏 1 1
𝑭𝑑𝑖𝑠𝑝 = − 𝐶𝐷 |𝒖𝐺 − 𝒖𝐿 | ( + )𝑔𝑟𝑎𝑑𝛼𝐺
4 𝑑𝐵 𝜎𝑇𝐷 𝛼𝐿 𝛼𝐺

In analogy to molecular diffusion, σTD is referred to as a Schmidt number and a value of 𝜎𝑇𝐷 =0.9 is
typically used.

2.5.4. Virtual Mass Force


When a bubble is accelerated, a certain amount of liquid has to be set into motion and pushed away as
well. This may be expressed as a force acting on the bubble as

𝐷𝐺 𝒖𝐺 𝐷𝐿 𝒖𝐿
𝑭𝑉𝑀 = −𝐶𝑉𝑀 𝜌𝐿 𝛼𝐺 ( − )
𝐷𝑡 𝐷𝑡

where DG/Dt and DL/Dt denote material derivatives with respect to the velocity of the indicated phase.
For the virtual mass coefficient, a value of CVM=0.5 have been applied.

2.5.5. Two-Phase Turbulence (Bubble Induced Turbulence)


A rising bubble, even though having a small mass, can agitate the liquid flow across its path and create
turbulence. This phenomenon is called Bubble induced turbulence (BIT). Due to the small density and
small spatial scales of the dispersed gas it suffices to consider this induced turbulence only in the
continuous liquid phase for bubbly flows. There are couple of approaches to consider this phenomenon.
The most widely used approach is proposed by Sato [29, 30] to add an extra contribution to the effective
viscosity. Another approach is to add a bubble-induced source terms directly to the generation terms (k,
ε and ω) in the equations of the turbulence model.

For the k source, it is assumed that all energy lost by the bubble due to drag is converted to turbulent
kinetic energy in the wake of the bubble. Hence, the k-source becomes

𝑑𝑟𝑎𝑔
𝑆𝐿𝑘 = 𝐹𝐿 ⋅ (𝑢𝐺 − 𝑢𝐿 )

11
For the ε-source, a similar heuristic is used as for the single-phase model, namely the k-source is divided
by some time scale τ so that
𝑆𝐿𝑘
𝑆𝐿𝜀 = 𝐶𝜀𝐵
𝜏

For 𝐶𝜀𝐵 a constant value of 1 is considered. Based on the dimensional analysis, it is possible to define
different time scale using two length scale (bubble size dB and eddy size ℓ=k 32
L /εL ) and two velocity

scale (relative velocity urel and square root of the turbulent kinetic energy √kL). Table 4 shows these
combinations.

Table 4. time scale for 𝑆𝐿𝜀 , used in different references

Bubble time scale Turbulence time scale reference


dB/urel ℓ/√kL = kL/εL = 1/ωL [31]
(𝐝𝟐𝐁 /𝛆𝐋 )𝟏𝟑 - [32]
dB/√kL ℓ/urel [8]

For SST model, the ε-source is transformed to an equivalent ω-source which gives

1 𝜀 𝜔𝐿 𝑘
𝑆𝐿𝜔 = 𝑆 − 𝑆
𝐶𝜇 𝑘𝐿 𝐿 𝑘𝐿 𝐿

This ω-source is used independently of the blending function in the SST model since it should be
effective throughout the fluid domain . In this study the approach described by Rzehak et. al. [8] will
be applied.

2.5.6. Lift Force and Critical Bubble Diameter


A bubble moving in an unbounded shear flow experiences a force perpendicular to the direction of its
motion. The momentum source corresponding to this shear lift force, often simply referred to as lift
force, can be calculated as:

𝐹𝑙𝑖𝑓𝑡 = −𝐶𝐿 𝜌𝐿 𝛼𝐺 (𝑢𝐺 − 𝑢𝐿 ) × 𝑟𝑜𝑡(𝑢𝐿 ),

There is a dependency of the lift force with the both bubble size and bubble shape in a flow with low
viscosity (low Morton number) systems under turbulent flow conditions [33]. For small and spherical
bubble, the shear lift coefficient CL is positive so that the lift force acts in the direction of decreasing
liquid velocity (negative velocity gradient) which in current case would be toward the wall. The
direction of the lift force changes its sign if a substantial deformation or size increase of the bubble
occurs. From the observation of the trajectories of single air bubbles rising in simple shear flow of a
glycerol water solution Tomiyama [34] proposed following correlation for the lift coefficient:

12
min[0.288tanh(0.121Re), f(Eo⊥ )] Eo⊥ < 4
{ f(Eo⊥ ) 4 < Eo⊥ < 10
−0.27 10 < Eo⊥

With 𝑓(𝐸𝑜⊥ ) = 0.00105𝐸𝑜⊥3 − 0.0159𝐸𝑜⊥2 − 0.0204𝐸𝑜⊥ + 0.474

This coefficient depends on the modified Eötvös number given by

𝑔(𝜌𝐿 − 𝜌𝐺 )𝑑⊥2
𝐸𝑜⊥ =
𝜎

where d⊥ is the maximum horizontal dimension of the bubble (major axis). It is calculated using an
empirical correlation the by Wellek et. al. [35] with the following equation (Wellek shape correlation):

3
𝑑⊥ = 𝑑𝐵 √1 + 0.163𝐸𝑜 0.757

where Eo is the usual Eötvös number.

Here the equivalent spherical diameter of the bubble is used to predict the major axis of the bubble
which will be used as characteristic length for lift coefficient prediction. As mentioned before, the lift
coefficient sign depends on both size and deformation of the bubbles, that is why major axis is
considered to be a good characteristic of both bubble shape and size. The small differences in the bubble
shape are crucial for the lift force, in particular for the sign change. Using Wellek correlation, Tomiyama
model predicts the sign change at a bubble size of 5.8 mm for air-water system. This size is also called
critical bubble diameter and is particularly important in CFD simulations and is used to divide the
population of bubbles into small and large groups and subgroups to take into account the polydispersity
nature of bubbly flows. Any bubble with a diameter higher than 5.8 is considered to be a large bubble
and vice versa. Large bubbles will have negative CL and tend to stay in centre of the flow while small
bubbles will be mainly concentrated near the wall region dude to positive lift coefficient.

2.6. Boundary Condition and Bubble Size Distribution

By having bubble samples and BSD three different setup of inlet boundary conditions are considered
as discussed below.

2.6.1. Monodisperse Approach


In monodisperse approach all bubbles, injected through the sparger are considered of the same size.
Using experimental data and by considering population of bubbles an average diameter can be obtained
known as sauter mean diameter with below formulation.

13
∑𝑛𝑖=1 𝑑𝑒,𝑖
3
𝑆𝑀𝐷 =
∑𝑛𝑖=1 𝑑𝑒,𝑖
2

The data for monodisperse bubbles for each flowrate are presented in Table 5. As it can be seen for all
flows the SMD is around 5.15 mm. So basically, for the studied system it can be concluded that in
homogeneous regime the average diameter for any flowrate can be considered 5.15 mm.

2.6.2. Bidisperse
Here the bubbles are divided into two groups of small and large size bubbles. The classification will be
made based on the critical bubble diameter where the lift coefficient changes sign from positive to
negative for small and large bubbles respectively. Table 5 shows this classification for each flowrate.

Table 5. Data for Bidisperse appaoch based on different bubble critical diameter

𝒅𝒄𝒓𝒊𝒕𝒊𝒄𝒂𝒍 = 5.8 mm
10 20 30 40 50 60
Group SMD VF SMD VF SMD VF SMD VF SMD VF SMD VF
Bi-Disperse Small 4.08 0.53 4.12 0.52 4.33 0.58 4.42 0.64 4.48 0.60 4.62 0.64
Large 7.26 0.47 7.29 0.48 6.97 0.42 6.66 0.36 6.76 0.40 6.83 0.36

10 20 30 40 50 60
Group SMD VF SMD VF SMD VF SMD VF SMD VF SMD VF
Mono-Disperse All 5.14 1 5.21 1 5.15 1 5.036 1 5.18 1 5.22 1

In case of bidisperse flow, a separate gas phase is defined for each velocity groups so basically two sets
of different mass and momentum is solved with two distinct velocity groups. The difference of
bidisperse with monodisperse approach will be further explained in discussion section.

One important issue in bubble selection especially in large diameter bubble column, is possibility of
different distribution of bubbles at the regions near the wall and central plane in developed region since
in a bubbly flow, small bubbles tend to move near the wall due to positive lift coefficient and larger
bubbles tend to flow at the centre of the column due to negative lift coefficient. So apparently, for
accurate acquisition of BSD we need BSD near the wall region and at the centre of the column. For
taking photos at different regions the camera focus is changed. The main problem which occurs for
selecting bubbles at the central plane is overlapping bubbles. The bubbles near the wall will block the
view and will not allow the correct selection of bubbles. This issue become more severe at higher
flowrate and it is almost impossible to select bubbles at flowrate above 30.

This problem might be overcome by using another tool like optical probe to include BSD at the centre.
However, as will be discussed later, for CFD simulation purposes the data from centre can be excluded
especially at homogeneous region and BSD at the wall would be a good representative of the whole
column BSD. In fact, one of the purpose of this study is to put this assumption into practice to see the

14
validity. Figure 8 shows the BSD and VFD (volume fraction distribution) for centre and wall for
flowrates of 10.

BSD 10 VF 10
0.08
0.1 0.07
0.09
0.08 0.06
0.07
0.05
0.06
0.05 0.04
0.04
0.03
0.03
0.02 0.02
0.01
0 0.01

wall center 50% Centre _ 50% Wall

Figure 8. BSD for different flowrate for cental and wall plane.

Visually, the only difference between cases is a peak at lower classes of BSD at the wall region. This
is due to the existence of small bubbles near the wall zone which are pushed to the wall due to their
positive lift force. Even though the frequency of these bubbles is high but their effect on volume fraction
distribution, which is an important CFD input data, can be ignored since the diameter is so small
compared to other bubbles. To better understand the differences, it is useful to compare the sauter mean
diameter (SMD) obtained from BSD at the wall and centre. Again, mono and bi-disperse approach are
considered here and as it can be seen in Figure 9 there are almost no difference between overall SMD
and SMDsmall and a slight difference between SMDlarge for centre, wall and their 50-50 combination. So
basically, the BSD at the wall can be considered a good representative of the whole column BSD. This
conclusion agrees with other studies [17, 20] .

10 20
8.000 7.26 8.00 7.29 7.09 7.20
7.00 7.13
7.000 7.00

6.000 5.14 4.91 5.02 6.00 5.21 5.01 5.11


5.000 5.00
Size (mm)
Size (mm)

4.08 4.11 4.09 4.12 4.17 4.15


4.000 4.00

3.000 3.00

2.000 2.00

1.000 1.00

0.000 0.00
SMD_small SMD_large SMDall SMD_small SMD_large SMDall

Figure 9. statestical comparison of BSD for centre and wall and their combiantaions

2.6.3. Polydisperse Flow


The bubble size distribution in real flows is not monodisperse or Bidisperse but it has rathe a
polydisperse nature and is changed along the column height due to coalescence and breakup phenomena

15
(Figure 2). This has an additional effect on mutual bubble interactions and in order to take into account
these phenomena, population balance modelling (PBM) is used in CFD simulations.

In CFX software there are two approaches to use PBM. One is MUSIG approach originally developed
by Lo [36]. Here different size groups are considered which will have a common velocity field. So
basically, one can imagine it as a monodisperse velocity group but multiple size groups. However, using
MUSIG model, the radial separation of small and large bubbles which is due to non-drag forces, cannot
be predicted especially when heterogeneous bubble motion becomes important.

The inhomogeneous multiple size group (iMUSIG) model introduced by Krepper et. al. [37] assigns
the bubble size classes used in the MUSIG model to different velocity groups. Each velocity group,
therefore, has its own velocity field (each with its own set of mass and momentum equations). In fact,
bi-disperse approach can be considered as the simplest case of polydisperse flow with two velocity
group, one size group for each, and no breakup and coalescence involved. This division is important to
describe effects like the bubble size-dependent movement of the gas phase caused by the lift force. It is
possible to consider bubble breakup and coalesces using both MUSIG and iMUSIG model. Considering
this, mass exchange would occur between sub-size groups and in case the size ranges exceed, mass
transfer terms between different velocity groups will occur. In this study Luo and Svendsen breakup
model [38] and Prince and Blanch coalesces model [39] are considered. Since mass transfer must be
included into the calculations, more source terms will appear in basic equations which are explained in
section 2.4. These differences will be applied only for gashouse phases and only mass transfer between
bubbles are considered without any mass transfer between liquid and gas phase.

The continuity equation for the gaseous dispersed phase i can then be written as
𝜕
⃗ 𝑖 ) = 𝑆𝑖
(𝛼 𝜌 ) + 𝛻 ⋅ (𝛼𝑖 𝜌g 𝑈
𝜕𝑡 𝑖 g

the momentum equation for the ith gaseous phase will become:
𝜕
⃗ ) + 𝛻 ⋅ (𝛼𝑖 𝜌g 𝑈
(𝛼 𝜌 𝑈 ⃗𝑖 ×𝑈
⃗ 𝑖 ) = 𝛻 ⋅ (𝛼𝑖 𝜇g (𝛻𝑈 ⃗ 𝑖 )T )) − 𝛼𝑖 𝛻𝑝 + 𝛼𝑖 𝜌g 𝑔 + 𝐹𝑖 + 𝑆𝑀𝑖
⃗ 𝑖 + (𝛻𝑈
𝜕𝑡 𝑖 g 𝑖

with

𝐹𝑖 = 𝐹𝑖,D + 𝐹𝑖,L + 𝐹𝑖,W + 𝐹𝑖,TD

The force terms are same as explained before. The term 𝑆𝑀𝑗 represents the transfer of gaseous phase
momentum between different velocity groups due to bubble breakup and coalescence processes that
causes bubbles of certain size to switch to a different velocity group. Additionally, for each sub-size
fraction, j (j = 1,…,Mi) in the velocity group i, the continuity equation for fjαi has to be solved.

16
𝜕
⃗ 𝑖 ) = 𝑆𝑗𝑖
(𝑓 𝛼 𝜌 ) + 𝛻 ⋅ (𝑓𝑗 𝛼𝑖 𝜌g 𝑈
𝜕𝑡 𝑗 𝑖 g

The source terms Sij represent the local transfer of gaseous phase mass due to bubble breakup and
coalescence processes. They can be assigned to Sk, which are the elements of the population balance
model. Note that in the above equations the index j extends over the range 1to N and the index k over
the range 1, …., ∑𝑁
𝑖=1 𝑀𝑖 . The population balance equations have then the form:

𝑆𝑘 = 𝐵𝑘,B − 𝐷𝑘,B + 𝐵𝑘,C − 𝐷𝑘,C

where Bk,B and Bk,C are the respective bubble birth rates due to breakup of larger bubbles from size
group k and coalescence of smaller bubbles to size group k. Dk,B and Dk,C are the respective bubble death
rates due to breakup of bubbles from size group k into smaller bubbles and the coalescence of bubbles
from size group k with other bubbles to even larger one [37].

In this study, a simplified model consisting of two velocity groups and overall 18 size group are
considered. The data at sparger region will be used here which generates bubbles with diameters larger
than 6 mm (for all flowrates). The size fractions for each velocity subgroups are shown in Table 6.

Table 6. size fraction and VF for polydisperse case

Size fraction
Groups 𝑑𝑚𝑒𝑎𝑛 10 20 30 40 50 60
Size Group 1 0.5 0.000 0.000 0.000 0.000 0.000 0.000
SMAL GROUP

Size Group 2 1.5 0.000 0.000 0.000 0.000 0.000 0.000


Size Group 3 2.5 0.000 0.000 0.000 0.000 0.000 0.000
Size Group 4 3.5 0.000 0.000 0.000 0.000 0.000 0.000
Size Group 5 4.5 0.000 0.000 0.000 0.000 0.000 0.000
Size Group 6 5.5 0.000 1.000 1.000 0.000 0.000 0.000
Size Group 7 7 0.022 0.006 0.008 0.010 0.006 0.014
Size Group 8 9 0.045 0.027 0.033 0.025 0.010 0.018
Size Group 9 11 0.099 0.063 0.061 0.075 0.039 0.043
Size Group 10 13 0.249 0.124 0.177 0.142 0.090 0.077
LARGE GROUP

Size Group 11 15 0.361 0.245 0.247 0.215 0.164 0.168


Size Group 12 17 0.127 0.271 0.225 0.165 0.213 0.223
Size Group 13 19 0.067 0.111 0.111 0.163 0.114 0.200
Size Group 14 21 0.030 0.062 0.034 0.090 0.189 0.095
Size Group 15 23 0.000 0.065 0.060 0.026 0.047 0.053
Size Group 16 25 0.000 0.000 0.019 0.068 0.080 0.046
Size Group 17 27 0.000 0.026 0.024 0.021 0.050 0.029
Size Group 18 29 0.000 0.000 0.000 0.000 0.000 0.036
VF_small - 0 0.0002 0.0008 0 0 0
VF_large - 1 0.9998 0.9992 1 1 1

17
2.7. Summary of Models and Studied Setups
CFX commercial codes (v. 18.2) is used to simulate the cases. The timestep is considered to be adaptive
and restriction of maximum CFL number lower than 0.5 is considered for calculation of time step. The
number of 150,000 mesh element are considered which is fairly fine mesh. Table 7 shows the summary
of considered models (baseline model).

Table 7. smmary of models used in CFD simualtion

Turbulence model SST k-ꙍ Model


Lift Coefficient Tomiyama
Drag Coefficient Ishii Zuber
Virtual Mass Force Included with 0.5 coefficient
Wall lubrication force Hosokawa
Turbulent dispersion force Farve Average Drag force developed by burns et. al.
Turbulence transfer and BIT Krepper et. al.

Table 8 shows different cases which are considered to compare the effect of mass flowrate at sparger
(uniform and non-uniform) and bubble size dispersity (Mono, Bi and polydisperse).

Table 8. Summary of cases used for comparison

Size approach CFD input data Mass flowrate Breakup Model Coalesces Model
Case 1 bi Sparger Non-uniform - -
Case 2 Mono and bi developed Non-uniform - -
Case 3 bi developed Uniform - -
Case 4 Poly Sparger Non-uniform Luo and Svendsen Prince and Blanch

3. Result and discussion


3.1. Case 1
In first attempt the simulations are performed using the data obtained from the BSD at sparger. As it
can be seen the gas holdup is highly under predicted. This is mainly due to large bubbles which are
generated at sparger. In the absence of break up and coalesces models this bubbles with the same size
will travel upward without any mass interaction. Large bubbles tend to create an unstable flow structure
which lead to a lower global holdup. The smaller the bubble, the more stable will be the flow, the higher
will be the holdup. As it can be seen from Figure 5 all bubbles are larger than 6 mm and average size
of bubble populations is between 13 to 16 mm depending on the flowrate (higher flowrate leads to larger
bubble generation). Also, the sparger case can be considered as mono disperse flow since all bubbles
are larger than bubble critical diameter of 5.8 mm. So basically, the data from sparger when no
breakup and coalesces are considered are not appropriate for the simulation.

18
9

Holdup (%)
5

0
0 10 20 30 40 50 60 70

Flowrate (lit/min)
simualtion Experimental

Figure 10. Global holdup prediction with data obtained from sparger BSD

3.2. Case 2
As it can be seen there is no difference between mono and Bidisperse approach however, the global
holdup is overpredicted for both approaches. The main reason for this overprediction could be the
assumption of critical bubble diameter since the volume fraction of small and large bubble groups are
based on the critical bubble diameter of 5.8 mm as predicted by Tomiyama model and Wellek shape
correlation. Probably higher amount of bubbles is considered in small groups and as mentioned before
smaller bubbles tend to stabilize the regime which ultimately leads to higher holdup.

Tommiyama
11
10
9
8
7
6 monoDisp
Holdup (%)

5 Experimental
4
BiDisp
3
2
1
0
0 10 20 30 40 50 60 70
Flowrate (lit/min)

Figure 11. holdup prediction for Case 2

Two more simulations, this time using BSD from centre of the column for flowrates of 10 and 20 are
performed. The results (in last two rows of Table 9) shows almost no differences in holdup perdition
and again proves the validity of using solo wall data to obtain BSD and other parameters.

The main difference between mono and bidisperse approach is different prediction of radial gas void
fraction profile as it can be seen in Figure 12 below.

19
Table 9. numerical results (case 2) with error with respect to the experimental data

Flowrate (lit/min) Tomiyama Error (%) Experimental


10 1.37 11.4 1.23
20 2.92 14.96 2.54
30 4.64 16.0 4.00
40 6.42 18.8 5.40
50 8.15 18.8 6.86
60 9.97 22.9 8.11
10 Centre 1.39 13.2 1.23
20 Centre 2.94 15.9 2.54

0.11

0.09
10-monodisperse

20-monodisperse

0.07 30-monodisperse

40-monodisperse
Void Fraction

50-monodisperse
0.05
10-bidisperse

20-bidisperse

30-bidisperse
0.03
40-bidisperse

50-bidisperse

0.01

-0.12 -0.07 -0.02 0.03 0.08


-0.01
R (m)

Figure 12. radial gas void fraction distribution, same model different size approach – height of 2.75
m from the sparger

Mono disperse approach predict flatter profile at the centre of the column as it is more visible in Figure
13 which shows the radial distribution for flowrate of 10. This difference becomes negligible at higher
flowrates especially for 40 and 50. This difference can also be observed in Figure14 which shows
contour of gas volume fraction at cross section of the column at height of 2.75 m.

20
0.025

0.02
Void Fraction

0.015

0.01

0.005

0
-0.12 -0.07 -0.02 0.03 0.08
R (m)

Tomiyama-10-mono Tomiyama-10-bi

Figure 13. radial gas void fraction distribution same model different size approach - – height of 2.75
m from the sparger

The reason can be explained by considering radial distribution of small and large velocity groups and
their sizes separately.

For monodisperse cases as can be seen in Table 5, for all flowrates the mean bubble size used in
simulations is around 5.15 mm. As mentioned before all bubbles smaller than 5.8 mm will be considered
as small bubble with positive lift coefficient and will be pushed away from the centre. So, a typical
behaviour for small bubbles would be a radial void fraction profile with a peak near the wall as it can
be seen in Figure 13 for mono disperse case.

However, for Bidisperse case the situation is different, and two groups of small and large bubbles are
simulated with different behaviour. For example, Figure 15 shows the radial gas void volume faction
for flowrate of 10. Again, the small group (4.08 mm) shows a peak near the wall however, large bubble
group which have a size higher than 5.88 mm (7.26 mm) will have a negative lift coefficient and tend
to stay at the centre which lead to a peak at the centre. At higher flowrates the difference between small
and large group size will reduce and both groups tend to approach 5.8 mm in size and a flatter profile.
So, the peak for large bubble group will reduce and small group profile tend to become flatter and as a
result the overall profile will be flatter as it can be seen in Figure 16 for flowrate of 50.

21
Bidisperse – flowrate 10 monodisperse – flowrate 10

Bidisperse – flowrate 50 monodisperse – flowrate 50


Figure 14. Contour of gas volume fraction in cross section of the column at height of 2.75 m

0.025

0.02
Void Fraction (-)

0.015
small group
0.01 large group

0.005 overall

0
-0.15 -0.1 -0.05 0 0.05 0.1 0.15
R (m)

Figure 15. radial gas volume fraction distribution for flowrate of 10 – height of 2.75 m from the sparger

22
0.12

0.1

Void Fraction (-)


0.08

0.06 large group


0.04 small group
0.02 overall

0
-0.15 -0.1 -0.05 0 0.05 0.1 0.15
R (m)

Figure 16. radial gas volume fraction distribution for flowrate of 50 – height of 2.75 m from the sparger

Observing global holdup and radial void fraction profile prediction, it can be concluded that global
holdup is not a sensitive parameter and cannot be consider as a solo parameter to validate a CFD model.
As it can be seen, even though the global holdup is predicted the same for mono and bidisperse
approach, void profiles are predicted differently.

3.3. Case 3
The calculation of Eta coefficient (η) (see section 2.3) is very time-consuming process so here the results
of simulations with uniform mass distribution at sparger with value of η =0.025 and velocity of 0.24
m/s (which is bubble terminal velocity) for each hole is presented.

No specific differences are observed for both global holdup and radial void fraction profile prediction
as it can be seen in Figure 17 and 18 and Table 10.

Tommiyama
9

6
Holdup (%)

0
0 10 20 30 40 50 60
Flowrate (lit/min)
Non-uniform Mass Dist experimental Uniform Mass Dist

Figure 17. CFD result comparison for uniform and non-uniform mass flow distribution at sparger

23
Table 10. Global holdup for uniform and uniform and non-uniform mass flow distribution at sparger

Flowrate Uniform Non-uniform Difference


(lit/min) (%)
10 1.45 1.44 0.6%
20 3.00 3.058 1.9%
30 4.70 4.703 0.1%
40 6.45 6.44 0.2%
50 8.22 8.19 0.4%

0.045
0.04
0.035
0.03
0.025
0.02
0.015
0.01
0.005
0
-0.12 -0.06 0 0.06 0.12

Non Uniform_Overall Non Uniform_Large Non Uniform_Small


Uniform_Overall Uniform_Large Uniform_Small

Figure 18. Radial gas void fraction for uniform and nonuniform mass distribution at sparger for
flowrate of 20.

The main reason for this independency can be explained by observing the flow structure in the column
for uniform and non-uniform mass flow distribution.

Figure 19 shows the velocity streamline at flowrate of 10. As it can be seen, for nonuniform distribution,
due to lateral flow in sparger, a large vortex is formed in the side which no bubbles are forming. These
vortices will destabilize the flow and can lead to a lower gas hold up. The phenomenon occurs near the
sparger and as it can be seen the vortex no longer exist above a specific region and near the developed
region. So basically, this vortex would affect hydrodynamics near sparger region to destabilise the flow
structure, but since the column is long enough, the flow will become stable again as it moves away from
the sparger. Also, the operating point is at homogeneous regime with mainly small bubbles. These small
bubbles have an important role in stabilizing the flow and healing the effect of vortices generated near
the sparger.

24
`
Figure 19. velocity streamline A: uniform mass flow, B: non-uniform mass flow

3.4. Case 4
In this case a polydisperse approach considering breakup and coalesces phenomena is simulated. The
global holdup prediction is presented in Figure 20 and Table 10 below. As it can be seen the global
holdup is again over predicted and compared to mono and bidisperse cases, the overprediction is even
higher.

9
8
7
6
5
Holdup (%)

4
3
2
1
0
0 10 20 30 40 50 60
Flowrate (lit/min)
Experimental Tomiyam + breakup&Coalesces Tomiyama No Breakup&coalesces

Figure 20. global holdup prediction for the case with/without breakup & coalesces models

25
Table 11. holdup prediction and errors

Flowrate with breakup&coalesces Error No breakup&coalesces Error experimental


(lit/min) (%) (%)
10 1.50 21.95% 1.39 13.17 1.23
20 3.10 22.05% 2.94 15.92 2.54
30 4.78 19.58% 4.64 16.05 4.00
40 6.60 22.22% 6.42 18.82 5.40
50 8.31 21.13% 8.15 18.81 6.86

As it can be seen the minimum and maximum error is 20 and 24% for cases with breakup&coalesces
models which are higher than minimum and maximum of 13 to 22% for bidisperse cases.

The reason can be explained by comparing simulated volume fraction distribution (VFD) and size
fractions with experimental ones at developed region. Figure 20 shows VFD at height of 2.75 m. As it
can be seen the volume fraction of small classes (1,2 and 3) are highly overpredicted by breakup model.
With this VFD, higher number of small bubbles are considered in the flow and as mentioned before
smaller bubbles tend to stabilize the flow and increase the holdup.

The same overprediction by models are reported by Liao et. al [11] who have compared the
implemented break up and coalesces in CFX (used in this study) with their own developed models.
Based on this observation it can be concluded that Luo and Svendsen breakup model implemented in
CFX are overpredicting the breakup phenomena and cannot be considered as a reliable model for large
scale applications.

flowrate 10 lit/min flowrate 10 lit/min

0.7 0.35
Transient average Conservative
Transient average Conservative Size

0.6 0.3
0.5 0.25
Volume Fraction

0.4 0.2
0.3 Simulation 0.15 Simulation
Fraction

0.2 experimental
0.1
Experimental
0.1 0.05
0 0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 1 2 3 4 5 6 7 8 9 101112131415161718
Size Group Size Group

26
flowrate 20 lit/min flowrate 20 lit/min
0.8
0.4
Transient average Conservative Size

Transient average Conservative


0.7
0.6 0.3

Volume Fraction
0.5
0.4 0.2
Fraction

Simulation Simulation
0.3
0.2
experimental 0.1 Experimental

0.1
0
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 1 2 3 4 5 6 7 8 9 101112131415161718

Size Group Size Group

Figure 21. simulated and experimental volume fraction distribution and size fractions for each size
groups

8. Conclusion
Numerical investigation of a large-scale bubble column is performed using multiphase Eulerian-
Eulerian approach for air-water system. The outcome of the study can be concluded in the following
bullet points. Experimental data are acquired and processed to obtain the necessary CFD inputs. More
importantly using BSD data three different approaches namely mono, bi and polydisperse for numerical
setup are investigated and results are compared.

The outcome of the current study can be summarized in the following bullet points:
1. BSD at the wall region of bubble column is a good representative of the whole column
particularly at homogenous regime.

2. In homogenous regime, even though the global holdup is predicted the same for mono and
bidisperse approach, but radial void profiles are predicted differently. So basically, global
holdup is not a sensitive parameter and cannot be consider as a solo parameter to validate a
CFD model.

3. BSD at the sparger should not be used when using fix dispersity approach (mono and
bidisperse) and without break up and coalesces models, instead the BSD at developed region
must be used to acquire CFD input data.

4. Final CFD simulation results are independent of mass flow distribution at the sparger for a
specific flowrate in homogenous regime and if the bubble column is long enough. No specific
differences are observed for global holdup and radial void fraction profile prediction.

5. In case of poly disperse approach, Luo and Svendsen breakup model is over estimating the
breakup phenomena leading to higher size fraction in small diameter bubble groups and
ultimately higher holdup prediction. More numerical study is needed to precisely predict bubble
behaviour in polydisperse flow and to improve prediction of breakup and coalesces phenomena.

27
References

1. Ekambara, K. and M.T. Dhotre, CFD simulation of bubble column. Nuclear Engineering and Design, 2010. 240(5): p. 963-969.
2. Ekambara, K., M.T. Dhotre, and J.B. Joshi, CFD simulations of bubble column reactors: 1D, 2D and 3D approach. Chemical Engineering Science, 2005. 60(23): p. 6733-6746.
3. Saleh, S.N., et al., CFD assesment of uniform bubbly flow in a bubble column. Journal of Petroleum Science and Engineering, 2018. 161: p. 96-107.
4. Besagni, G., et al., Computational Fluid-Dynamic modeling of the pseudo-homogeneous flow regime in large-scale bubble columns. Chemical Engineering Science, 2017.
160: p. 144-160.
5. Guan, X. and N. Yang, Bubble properties measurement in bubble columns: From homogeneous to heterogeneous regime. Chemical Engineering Research and Design, 2017.
127: p. 103-112.
6. Trivedi, R., T. Renganathan, and K. Krishnaiah, Hydrodynamics of countercurrent bubble column: Experiments and predictions. Chemical Engineering Journal, 2018.
7. Besagni, G., L. Gallazzini, and F. Inzoli, On the scale-up criteria for bubble columns. Petroleum, 2017.
8. Rzehak, R. and E. Krepper, CFD modeling of bubble-induced turbulence. International Journal of Multiphase Flow, 2013. 55: p. 138-155.
9. Rzehak, R., et al., Unified modeling of bubbly flows in pipes, bubble columns, and airlift columns. Chemical Engineering Science, 2017. 157: p. 147-158.
10. Ziegenhein, T., R. Rzehak, and D. Lucas, Transient simulation for large scale flow in bubble columns. Chemical Engineering Science, 2015. 122: p. 1-13.
11. Liao, Y., et al., Baseline closure model for dispersed bubbly flow: Bubble coalescence and breakup. Chemical Engineering Science, 2015. 122: p. 336-349.
12. Kulkarni, A.V. and J.B. Joshi, Design and selection of sparger for bubble column reactor. Part II: Optimum sparger type and design. Chemical Engineering Research and
Design, 2011. 89(10): p. 1986-1995.
13. Kulkarni, A.V. and J.B. Joshi, Design and selection of sparger for bubble column reactor. Part I: Performance of different spargers. Chemical Engineering Research and
Design, 2011. 89(10): p. 1972-1985.
14. Kulkarni, A.V., Design of a Pipe/Ring Type of Sparger for a Bubble Column Reactor. Chemical Engineering & Technology, 2010. 33(6): p. 1015-1022.
15. Su, X., Gas holdup in a gas-liquid-fiber semi-batch bubble column. phd thesis. Iowa State University, 2005.
16. K. Saravanan, R.V., Chandramohan., Gas holdup in Multiple Impeller Agitated Vessels. Modern Applied Science, 2009. 3(2).
17. Besagni, G. and F. Inzoli, Comprehensive experimental investigation of counter-current bubble column hydrodynamics: Holdup, flow regime transition, bubble size
distributions and local flow properties. Chemical Engineering Science, 2016. 146: p. 259-290.
18. Thorat, B.N., et al., Effect of Sparger Design and Height to Diameter Ratio on Fractional Gas Hold-up in Bubble Columns. Chemical Engineering Research and Design, 1998.
76(7): p. 823-834.
19. Majumder, S.K., G. Kundu, and D. Mukherjee, Bubble size distribution and gas–liquid interfacial area in a modified downflow bubble column. Chemical Engineering
Journal, 2006. 122(1): p. 1-10.
20. Besagni, G. and F. Inzoli, Bubble size distributions and shapes in annular gap bubble column. Experimental Thermal and Fluid Science, 2016. 74: p. 27-48.
21. Aloufi, F.M., An investigation of gas void fraction and transition conditions for two-phase flow in an annular gap bubble column. Loughborough University, 2011.
22. Besagni, G., et al., Experimental investigation on the influence of ethanol on bubble column hydrodynamics. Chemical Engineering Research and Design, 2016. 112: p. 1-15.
23. Yamashita, F., Y. Mori, and S. Fujita, sizes and size distributions of bubbles in a bubble column, comparison between the two point electric probe method and the
photographic method. Journal of Chemical Engineering of Japan, 1979. 12(1): p. 5-9.

24. Polli, M., et al., Bubble size distribution in the sparger region of bubble columns. Chemical Engineering Science, 2002. 57(1): p. 197-205.
25. Hormozi, M.A.a.F., Experimental Determination of Bubble Size in Solution of Surfactants of the Bubble Column. J Adv Chem Eng 2014. 4(101).
26. Ishii, M. and N. Zuber, Drag coefficient and relative velocity in bubbly, droplet or particulate flows. AIChE Journal, 1979. 25(5): p. 843-855.
27. Hosokawa, S., Tomiyama, A., Misaki, S., Hamada, T., Lateral Migration of Single Bubbles Due to the Presence of Wall. Proceedings of the ASME Joint U.S.-European Fluids
Engineering Division Conference, Montreal, Canada., 2002.
28. Burns, A.D., Frank, T., Hamill, I., Shi, J.-M., The Favre averaged drag model for turbulence dispersion in Eulerian multi-phase flows. Proceedings of the 5th International
Conference on Multiphase Flow, ICMF2004, Yokohama, Japan., 2004.
29. Sato, Y., M. Sadatomi, and K. Sekoguchi, Momentum and heat transfer in two-phase bubble flow—I. Theory. International Journal of Multiphase Flow, 1981. 7(2): p. 167-
177.
30. Sato, Y. and K. Sekoguchi, Liquid velocity distribution in two-phase bubble flow. International Journal of Multiphase Flow, 1975. 2(1): p. 79-95.
31. Politano, M.S., P.M. Carrica, and J. Converti, A model for turbulent polydisperse two-phase flow in vertical channels. International Journal of Multiphase Flow, 2003. 29(7):
p. 1153-1182.
32. Morel, C., Turbulence Modeling and First Numerical Simulations in Turbulent Two-phase Flows. CEA Grenoble, France (1997).
33. Ziegenhein, T., Tomiyama, A. & Lucas, D.,, A new measuring concept to determine the lift force for distorted. . International Journal of Multiphase Flow, submitted., 2017. .
34. Tomiyama, A., Struggle with computational bubble dynamics,ICMF'98, 3rd Int. Conf. Multiphase Flow, Lyon, France, June 8-12, 1998: p. pp. 1-18.

35. Wellek, R.M., A.K. Agrawal, and A.H.P. Skelland, Shape of liquid drops moving in liquid media. AIChE Journal, 1966. 12(5): p. 854-862.
36. S., L., Application of the MUSIG model to bubbly flows. AEAT-1096, AEA Technology, June 1996.
37. Krepper, E., et al., The inhomogeneous MUSIG model for the simulation of polydispersed flows. Nuclear Engineering and Design, 2008. 238(7): p. 1690-1702.
38. Luo, S.M., and Svendsen, H., Theoretical Model for Drop and Bubble Breakup in Turbulent Dispersions. AIChE Journal 42: p. 1225 -1233.
39. Prince, M.a.B., H., Bubble Coalescence and Break-Up in Air-Sparged Bubble Columns. AIChE Journal. 36: p. 1485-1499.

28

You might also like