Fluids 08 00091
Fluids 08 00091
Review
Computational Fluid Dynamics Modelling of Two-Phase
Bubble Columns: A Comprehensive Review
Giorgio Besagni , Nicolò Varallo * and Riccardo Mereu
Abstract: Bubble columns are used in many different industrial applications, and their design and
characterisation have always been very complex. In recent years, the use of Computational Fluid
Dynamics (CFD) has become very popular in the field of multiphase flows, with the final goal of
developing a predictive tool that can track the complex dynamic phenomena occurring in these types
of reactors. For this reason, we present a detailed literature review on the numerical simulation of
two-phase bubble columns. First, after a brief introduction to bubble column technology and flow
regimes, we discuss the state-of-the-art modelling approaches, presenting the models describing the
momentum exchange between the phases (i.e., drag, lift, turbulent dispersion, wall lubrication, and
virtual mass forces), Bubble-Induced Turbulence (BIT), and bubble coalescence and breakup, along
with an overview of the Population Balance Model (PBM). Second, we present different numerical
studies from the literature highlighting different model settings, performance levels, and limitations.
In addition, we provide the errors between numerical predictions and experimental results concerning
global (gas holdup) and local (void fraction and liquid velocity) flow properties. Finally, we outline
the major issues to be solved in future studies.
Keywords: bubble columns; two-phase flow Computational Fluid Dynamics; momentum exchange;
bubble induced turbulence; population balance model
Fluid Dynamics (CFD) has spurred substantial research efforts into determining numerical
models that can obtain reasonably accurate predictions with limited computational time,
thereby overcoming the limitations of traditional empirical methods.
The rest of this paper is organised as follows. In Section 2, we discuss the flow
regimes occurring in a vertical pipe, and more in detail in large-diameter bubble columns.
In Section 3, we present numerical approaches to simulating two-phase bubble columns.
In particular, within the context of the Eulerian–Eulerian approach, an overview of the
population balance model used to predict the bubble size distribution follows a description
of the different models used to consider the momentum exchange between the phases.
Finally, models for considering the influence of the dispersed phase on the continuous
phase turbulence are provided. In Section 4, we present different studies concerning
the numerical simulation of two-phase bubble columns and provide their model settings
and performances.
2. Flow Regimes
In multiphase reactors, the flow regime provides information about the behaviour of
the dispersed phase and its interaction with the continuous phase. In the most general case,
four flow regimes are encountered in vertical pipes when increasing the gas flow rate using
a fixed system design and phase properties (Figure 2): (1) homogeneous regimes, (2) slug
regimes, (3) heterogeneous (churn turbulent) regimes, (4) and annular flow regimes.
Figure 2. Representation of flow regimes in vertical pipes. Reprinted with permission from [4].
The slug flow regime is not observed when dealing with industrial applications owing
to Rayleigh–Taylor instabilities [5] which arise because of large diameter effects. Quantifi-
cation of the Rayleigh–Taylor instabilities at the reactor scale is obtained by comparing the
dimensionless column diameter (D ∗H ) with a critical value (D ∗H,cr ):
DH
D ∗H = p > D ∗H,cr ≈ 52 (1)
σ/g(ρ L − ρG )
In Equation (1), D H is the hydraulic diameter of the column, σ is the surface tension,
g is the acceleration due to gravity, ρ L is the liquid phase density, and ρG is the gas phase
density. A bubble column is classified as having a large diameter if D ∗H is greater than
D ∗H,cr = 52 [6], i.e., when dc ≥ 0.13–0.15 m at ambient conditions.
Under the constraint of Equation (1), the fluid dynamics of large-diameter bubble
columns is explicated in six flow regimes and interpreted by a function of two global fluid
dynamics parameters, namely, the drift flux and the gas holdup [7].
The drift flux (JT ) is defined as the volumetric flux of either component relative to a
surface moving at volumetric average velocity, and is expressed as follows:
J T = UG ( 1 − ε G ) ± U L ε G (2)
In Equation (2), ε G is the gas holdup (defined as the ratio between the volume occupied
by the gas phase and the total volume), UG is the superficial gas velocity, and UL is the
superficial liquid velocity. The sign on the right-hand side depends on the operation mode
of the bubble column, i.e., co-current mode (+) or counter-current mode (−); in batch mode,
UL = 0.
The six flow regimes emerging upon an increase in the gas flow rate are: (1) mono-
dispersed homogeneous flow regimes, (2) poly-dispersed homogeneous flow regimes,
(3) transition flow regimes without coalescence-induced structures, (4) transition flow
regimes with coalescence-induced structures, (5) pseudo-heterogeneous flow regimes,
and (6) pure-heterogeneous flow regimes (Figure 3).
First, the mono-dispersed homogeneous and the pseudo-homogeneous flow regimes
are progressively observed; a mono-dispersed bubble size distribution characterises the for-
mer, while a poly-dispersed one in which large bubbles with a negative lift force coefficient
move through the centre of the column characterises the latter. With an increase in the gas
flow rate, the number of bubbles having a negative lift force coefficient increases, and these
Fluids 2023, 8, 91 4 of 41
continue to migrate towards the centre of the column (Figure 4). Then, the transition flow
regime (i.e., the transition flow regime without coalescence-induced structures) begins to
be established by non-stable coalescence-induced structures, which causes oscillations in
the gas holdup. Subsequently, the coalescence-induced structures stabilise, characterising
the transition flow regime with coalescence-induced structures. Despite the presence of
coalescence-induced structures, they are not entirely developed. Indeed, increasing the gas
flow rate leads to a decrease in the gas holdup, thereby unveiling the imbalance between a
higher gas flow rate and the formation of coalescence-induced structures, which indicates
a transition process from the prevailing fluid dynamics.
Finally, stable coalescence-induced structures emerge, leading to the pseudo-heterogeneous
and purely heterogeneous flow regimes. The former is characterised by an equilibrium
between the increase in the flow rate (in terms of drift flux) and the formation of coalescence-
induced structures. This leads to constant holdup with respect to an increase in the drift flux,
indicating that the coalescence-induced structures are well established. On the contrary,
in the pure-heterogeneous flow regime increasing the gas flow rate increases coalescence-
induced structures, and the increase in drift flux is followed by an increase in the gas
holdup; for more details, see the paper by Besagni (2020) [7].
Figure 3. Flow regimes and regime transitions in a large-diameter bubble column. Reprinted with
permission from [7].
Figure 4. Flow pattern evolution in two-phase bubble columns. Reprinted with permission from [8].
Fluids 2023, 8, 91 5 of 41
∂ ~ I,k
(α ρ ~u ) + ∇ · (αk ρk~uk~uk ) = −αk ∇ p + ∇ · (αk τ̄k ) + αk ρk~g + M (4)
∂t k k k
where τ̄k is the stress–strain tensor for the k-th phase, defined as
2
τ̄k = αk µk ∇~uk + ∇~ukT + αk λk − µk ∇ · ~uk I (5)
3
Here, µk and λk express the shear and bulk viscosity of the k phase, respectively.
The other terms on the right-hand side of Equation (4) represent the pressure gradient and
the interfacial momentum exchange between the phases, accounting for all the interaction
forces exchanged between the phases (see Section 3.2).
~ L→ G = − M
M ~ G→ L = ~FD,L→G + ~FL,L→G + ~FTD,L→G + ~FD,W L→G + ~FV M,L→G (6)
Fluids 2023, 8, 91 6 of 41
Figure 5. Schematic representation of the interphase forces acting on a bubble. Reprinted with
permission from [13].
gd2b
Eo = (ρ L − ρG ) (9)
σ
gµ4L (ρ L − ρG )
Mo = (10)
σ3 ρ2L
For spherical bubbles without deformation and fully contaminated systems, the func-
tional form of the drag coefficient depends only on the bubble Reynolds number, as deter-
mined by Schiller and Naumann (1935) [14]. They modified the Stokes model, originally
formulated only for creeping flows (Reb << 1), by adding a correction factor to extend its
Fluids 2023, 8, 91 7 of 41
validity up to Reb = 1000. A constant value of 0.44 was considered in order to extend the
applicability of the model:
(
24
1 + 0.15Re0.687
∞ Reb b Reb ≤ 1000
CD = (11)
0.44 Reb > 1000
This model is suitable only for small bubbles, and is commonly used in bubble columns
operated in the homogeneous flow regime.
Similarly to Shiller and Naumann (1935) [14], White (1974) [15] proposed a drag
coefficient correlation highlighting the dependency of the drag coefficient only on the
bubble Reynolds number:
∞ 24 6
CD = 0.44 + + √ (12)
Reb 1 + Reb
For deformed bubbles, the drag coefficient is affected by the Eötvös and Morton
numbers. Grace et al. (1976) [16] were among the first to publish a drag law suitable for
deformed bubbles:
∞
CD = max min CD, ellipse , CD, cap , CD, sphere (13)
where (
24
Reb Reb < 0.01
CD, sphere = 24
(14)
1 + 0.15Re0.687
Reb b Reb ≥ 0.01
4 gdb ρ L − ρG
CD, ellipse = (15)
3 Ut2 ρL
8
CD, cap = (16)
3
In Equation (15), the bubble terminal velocity (Ut ) is calculated as follows:
µL
Ut = Mo −0.149 ( J − 0.857) (17)
ρ L db
0.94H 0.757
2 < H ≤ 59.3
J= (18)
3.42H 0.441 H > 59.3
and
µ L −0.14
4 −0.149
H = EoMo (19)
3 µref
In Equation (19), µref is set to 0.0009 kg/(m·s).
A model with the same functional form as Equation (13) was proposed by Ishii and
Zuber (1979) [17], who distinguished different shape regimes and found the following:
24
CD, sphere = 1 + 0.1Re0.75
b (20)
Reb
2√
CD, ellipse = Eo (21)
3
8
CD, cap = (22)
3
Fluids 2023, 8, 91 8 of 41
Except at high values of Eötvös number, this correlation agrees well with experimental
results concerning the terminal velocity of bubbles rising in quiescent liquids.
Grevskott et al. (1996) [18] experimentally studied the drag coefficient considering
the bubble size distribution in pure water by dividing the population of bubbles into
small and large bubbles. Fitting the experimental data, they proposed the following drag
coefficient correlation:
1.05.645 db ≥ 2.0 mm
∞
CD = Eo +2.385 2 (23)
8 (1 − α ) d < 2.0 mm
3 G b
A popular correlation suitable for gas–liquid flows with a range of shapes was derived
by Tomiyama et al. (1998) [19]. The proposed drag coefficient consists of three equations,
respectively corresponding to pure, slightly contaminated, and fully contaminated systems:
48 8 Eo
∞ 16
CD = max min 1 + 0.15Re0.687
b , , (24)
Reb Reb 3 Eo + 4
72 8 Eo
∞ 24 0.687
CD = max min 1 + 0.15Reb , , (25)
Reb Reb 3 Eo + 4
8 Eo
∞ 24
CD = max 1 + 0.15Re0.687
b , (26)
Reb 3 Eo + 4
This drag correlation has a wide range of applicability (10−3 < Rb < 106 and
10−2 < Eo < 103 ) and holds for spherical as well as deformed bubbles.
Zhang et al. (2006) [20] derived a simple relation that provides results comparable
to other models when the bubble diameter is in the range between 4 mm and 10 mm,
for which the rising velocity is weakly dependent on the bubble diameter:
∞ 2√
CD = Eo (27)
3
Recently, Dijkhuizen et al. (2010) [21] used a DNS front-tracking technique to derive a
correlation valid for both spherical and deformed bubbles:
q
∞ 2 ( Re ) + C2 ( Eo )
CD = CD b D (28)
where " #
16 Reb
CD ( Reb ) = 1+ (29)
8 + 0.5 Reb + 3.315Re0.5
Reb b
4Eo
CD ( Eo ) = (30)
9.5 + Eo
The proposed drag closure matches very well with experimental data for ultra-pure
liquids (water/glycerine mixture). A comparison between the different drag coefficient
correlations is presented in Figure 6.
Fluids 2023, 8, 91 9 of 41
2.5
2.0
CD [-]
1.5
1.0
0.5
Figure 6. Evolution of the drag coefficient with respect to the bubble diameter at |u~G − u~L | = 0.1 m/s
for air bubbles in water. Comparison between the correlations of Schiller and Naumann (1933) [14],
Grace et al. (1976) [16], Ishii and Zuber (1978) [17], Grevskott et al. (1996) [18] and Tomiyama
(1998) [19].
Several authors have investigated the impact of bubble–bubble interactions on the drag
coefficient. The first approach was proposed by Bridge et al. (1964) [22] and Wallis (1969) [23]:
1
h= (32)
1 − α2n
G
for which Bridge et al. (1964) [22] selected n = 1.39 and Wallis (1969) selected [23] n = 1.
Ishii and Zuber (1979) [17] investigated the drag force over a broad range of gas
volume fractions and Reynolds numbers, establishing the following relation based on a
large sample of experimental data:
1
h= √ (33)
1 − αG
Fluids 2023, 8, 91 10 of 41
Rusche and Issa (2000) [24] proposed a swarm factor correlation for oblate bubbles,
validated for a local volume fraction αG ≤ 0.45 and 27 ≤ Reb ≤ 960:
h = exp(3.64αG ) + α0.4864
G (34)
Roghair et al. (2011) [25], using a DNS front-tracking model, derived a drag-modification
term that can be applied for high Reynolds numbers and oblate bubbles in the range
αG < 0.45, 150 ≤ Rb ≤ 1200, 1 ≤ Eo ≤ 5, 4 × 10−12 ≤ Mo ≤ 2 × 10−9 :
18αG
h = 1+ (1 − α G ) (35)
Eo
Buffo et al. (2016) [26] developed a straightforward power law equation to account
for the crowding effect of bubbles based on many investigations into fluidized beds and
liquid–liquid systems: (
(1 − αG )CA αG ≤ 0.8
h= (36)
1 αG > 0.8
The authors concluded from a sensitivity analysis that a value of C A = 1.3 provides
the best agreement with experimental data.
The swarm factor correlations discussed above hinder the rising velocity of the bubbles,
thereby increasing the drag coefficient of isolated bubbles at high gas volume fractions.
They are reported in Figure 7 and compared with the Simmonet et al. (2008) [27] model;
while reasonable agreement is found at low gas volume fractions, the predictions diverge
strongly at high volume fractions. The correction factor proposed by Simmonet et al.
(2008) [27] differs from the other models in highlighting the presence of a critical value
of the gas volume fraction at 0.15. Below this value, the rising velocity of the bubbles
decreases with increasing void fraction; on the contrary, beyond the critical value the wake
acceleration effect of large bubbles is predominant, causing an increase in the bubbles’
rising velocity and a consequent reduction in the swarm factor, which decays rapidly to
zero. For local void fractions less than 0.30 and air–water pure systems, the correlation of
Simonnet et al. (2008) [27] reads as follows:
m −2/m
m αG
h = (1 − αG ) (1 − αG ) + 4.8 (37)
1 − αG
where m is an adjustable coefficient. Simonnet et al. (2008) [27] set m = 25, as this showed
good agreement with their experimental data.
McClure et al. (2014c) [28] suggested that no hindering effect should be taken into
account for bubbles larger than 7 mm, and modified the correlation of Simonnet et al.
(2008) [27] accordingly: (
min(h0 , 1) h0 > 1
h= (38)
0.8h0 h0 < 1
where h0 is the original swarm factor of Simonnet et al. (2008) [27].
Performing a new experimental study, McClure et al. (2017) [29] criticized their
previous model and observed that the drag correction factor is a function of both the bubble
size distribution and the local gas volume fraction. Thus, they proposed a new empirical
correlation valid for αG > 0.25:
h = min (1 − αG )n + b, 1
(39)
where n and b are model constants calculated with least-square fitting of experimental data.
The authors suggested n = 50, while b depends on the gas sparger design (for example,
a value of 0.20 was found in the case of a perforated plate distributor with a 0.5 mm hole
Fluids 2023, 8, 91 11 of 41
diameter, while a value of 0.08 was found in the case of a perforated plate with a 3 mm
hole diameter).
2.0
1.5
h [-]
1.0
0.5
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
G [-]
Figure 7. Comparison of published Swarm factor as a function of local void fraction for bubbles with
db of 6.5 mm. Comparison between the correlations of Bridge et al. (1964) [22], Wallis (1969) [23],
Ishii and Zuber (1979) [17], Rusche and Issa (2000) [24], Simonnet et al. (2008) [27], Roghair et al.
(2011) [25] and Buffo et al. (2016) [26]. Note the discrepancy between the correlation proposed by
Simonnet et al. (2008) [27] and the others.
Gemello et al. (2018a) [30] modified the correlation proposed by Simonnet et al.
(2008) [27] (validated for αG < 0.3) to obtain a swarm factor feasible for a wide range
of operating conditions. In particular, they noted that the decrease in the swarm factor
proposed by Simonnet et al. (2008) [27] is too strong for high volume fractions; thus,
they added a minimum asymptotic value (hmin ) towards which h should tend at high
void fractions:
" 25 #−2/25
α G
h = max (1 − αG ) (1 − αG )25 + 4.8 , hmin (40)
1 − αG
1
k b,large = (42)
max 1.90α g f b, large
22αG, small
k b,small = (1 − αG,small ) 1 + (43)
Eo + 0.4
Fluids 2023, 8, 91 12 of 41
where f b, large is the fraction of large bubbles and αG, small = αG 1 − f b,large . From this
model equation, it is possible to observe that an increasing void fraction means that the drag
correction factor increases to a maximum and then decreases. Drag correction factors that
reduce the drag force in bubble swarms are presented in Figure 8.
0.4
0.2
0.0
0.0 0.1 0.2 0.3 0.4 0.5
G [-]
Figure 8. Drag correction factors that reduce the drag force in bubble swarms for bubbles with db of
6.5 mm. Comparison between the correlations of McClure et al. (2014c) [28], McClure et al. (2017) [29],
Gemello et al. (2018) [30] and Yang et al. [31].
where CL is the lift coefficient. Experimental and numerical studies show that the coefficient
is positive for small bubbles and negative for large bubbles. Consequently, the lift force
direction depends on the bubble size and shape. For small bubbles, the lift force acts in
the direction of decreasing liquid velocity, which in the case of batch or co-current mode is
toward the pipe wall. Conversely, when large bubbles are considered, a force assimilated
to the lift force pushes the bubbles towards the centre of the column. In other words, sign
reversal of the lift force is observed for bubbles with a diameter larger than a critical value
(dcr ). Smaller bubbles (those with a diameter lower than the critical threshold) tend to move
towards the wall, whereas larger bubbles (those for which db > dcr ) are pushed towards
the vessel centre.
Tomiyama et al. (2002b) [32] studied the motion of bubbles in a glycerol–water solution,
deriving the following correlation for the lift coefficient:
min[0.288 tanh(0.121Reb ), f ( Eo⊥ )] Eo⊥ ≤ 4
CL = f ( Eo⊥ ) 4 < Eo⊥ ≤ 10 (45)
−0.27 Eo⊥ > 10
where
3 2
f ( Eo⊥ ) = 0.00105Eo⊥ − 0.0159Eo⊥ − 0.0204Eo⊥ + 0.474; (46)
Fluids 2023, 8, 91 13 of 41
here, Eo⊥ is the modified Eötvös number calculated considering the maximum horizontal
dimension (major axis) of the bubble (d⊥ ), provided by the empirical correlation for the
aspect ratio (E) proposed by Wellek et al. (1966) [33] for contaminated flows:
g ρ L − ρ g d2⊥
Eo⊥ = (47)
σ
1/3
1 p
3
d⊥ = deq = deq 1 + 0.163Eo0.757 (48)
E
where deq is the equivalent spherical diameter of the bubble.
This correlation applies to larger-scale deformable bubbles in the ellipsoidal and
spherical cap regimes. The experimental conditions under which Equation (45) was derived
were limited to the ranges of −5.5 ≤ log10 ( Mo ) ≤ −2.8 and 1.39 ≤ Eo ≤ 5.84. Air–water
systems at normal conditions usually have Mo = 2.63 × 10−11 ( log10 ( Mo ) = −10.55),
which is quite different from the conditions at which the above correlation was derived;
nevertheless, good results have been reported for this case. When coupled with that of
Wellek et al. (1966) [33] for the calculation of d⊥ , the correlation predicts that the change in
sign of the lift coefficient occurs at db = 5.8 mm.
However, Ziegenhein and Lucas (2019) [34] pointed out that the correlation of Wellek
et al. (1966) [33] is not suitable for purified air–water systems and leads to an over-
prediction of the major axis of an elliptic bubble. Besagni et al. (2016) [35] studied the size
distribution and shapes of bubbles in an annular gap bubble column and reached the same
conclusion. Correct estimation of the critical diameter is essential in CFD modelling, as it
allows the bubble population to be divided into a small group and large group, permitting
consideration of the poly-disperse nature of the flow in which bubbles of different sizes
move in different directions.
To overcome this limitation, Ziegenhein et al. (2018) [36] suggested a quadratic
fit correlation for the lift force as a function of the modified Eötvs̈ number and bubble
Reynolds number:
(
0.002Eo⊥2 − 0.1Eo + 0.5 Eo ≤ 10.5
⊥ ⊥
CL = (49)
−0.3295 Eo⊥ > 10.5
In their model, the correlation for the bubble major axis is modified based on experi-
mental observation of bubble shapes in six column configurations:
p
3
d⊥ = deq 1 + 0.65Eo0.35 (50)
The Ziegenhein et al. (2018) [36] correlations results in a critical diameter of about
5.3 mm, which is lower than that obtained with the Wellek et al. (1966) [33] relation. Figure 9
shows a comparison between the Tomiyama et al. (2002) [32] and the Ziegenhein et al.
(2018) [36] lift models.
To conclude, Behzadi et al. (2004) [37] proposed a model highlighting that the lift coef-
ficient cannot be independent of the local gas void fraction. After fitting their experimental
observations, they proposed the following correlation:
CL = 6.51 × 10−4 α−
G
1.2
(51)
It should be pointed out that this simple correlation does not consider the change
of sign of the lift coefficient (i.e., CL does not depend on Eo⊥ ) and, due to the power law
formulation, the coefficient tends to infinity for αG −→ 0. This is avoided by limiting Cl by
a finite value, which is taken as 0.25 [37].
Fluids 2023, 8, 91 14 of 41
0.4
0.3
0.2
CL [-]
0.1
0.0
−0.1
−0.2
−0.3
2 4 6 8 10 12
db [mm]
Figure 9. Comparison between Tomiyama et al. (2002b) [32] and Ziegenhein et al. (2018) [34] models
for the lift coefficient.
In analogy to the molecular diffusion, σTD is the Schmidt number (a value of σTD = 0.9
is typically used) and CTD is the turbulent dispersion coefficient, while k LG and Dt,LG are
the mixture’s turbulent kinetic energy and the liquid dispersion scalar, respectively.
Burns et al. (2004) [39] derived an explicit formulation for the turbulent dispersion
force by using Favre averaging on the drag force:
turb
∇α L ∇αG
~FTD,L→G = − 3 CTD α L (1 − α L ) CD |~uG − ~u L | µ L − (53)
4 db σTD αL αG
This model estimates the dispersion scalar of the Simonin and Viollet (1990) [38] model
with the liquid turbulent viscosity µturb
L .
An alternative formulation is the one proposed by Lopez De Bertodano et al. (2004) [40]:
1
~FTD,L→G = −ρ L ~u0 ~u0 · ∇αG (54)
St(1 + St) L L
where St is the Stokes number and ~u0L~u0L represents the normal Reynolds stresses.
Fluids 2023, 8, 91 15 of 41
where (u~G − u~L )|| is the relative velocity component parallel to the wall, ~yˆ is the unit normal
to the wall pointing into the fluid, and CW L is the wall lubrication coefficient.
Different expressions for CW L have been proposed in the literature. The relation
proposed by Antal et al. (1991) [41] states that
Cw1 Cw2
CW L = max 0, + (56)
db yw
With Cw1 and Cw2 being two non-dimensional coefficients. The Antal et al. (1991) [41]
model is applicable only if the condition yw ≤ 5db is respected; thus, it is only active on
a sufficiently fine mesh. Tomiyama et al. (1995) [42], studying bubble trajectories in a
glycerol–water solution, proposed a different function for the wall lubrication coefficient
by considering an additional dependence on the Eötvös number and pipe diameter (D):
!
db 1 1
CW L = Cw − (57)
2 y2w ( D − y w )2
0.47 Eo < 1
e−0.933Eo+0.179
1 ≤ Eo ≤ 5
Cw = (58)
0.00599Eo − 0.0187 5 < Eo ≤ 33
0.179 33 ≤ Eo
Although the Tomiyama (1995) [42] model has proven superior to the Antal et al.
(1991) [41] model, it is restricted to flows in pipe geometries.
Hosokawa et al. (2002) [43] reviewed the experimental results of Tomiyama (1995) [42]
and obtained the following functional form for CW L :
db
CW L = Cw (59)
2yw
Here, the coefficient Cw depends on the Eötvös number and the phase-relative Reynolds
number (Reb ), as indicated below:
!
7
Cw = max , 0.0217Eo (60)
Re1.9
b
The Hosokawa et al. (2002) [43] model often shows better agreement with experimental
data in both vertical and inclined bubbly flow systems.
Later, Frank et al. (2005) [44] proposed a generalized formulation for the wall lubrica-
tion coefficient:
yw
1− C d
1 WC b
CW L = Cw (Eo) · max 0, · p −1 (61)
CWD yw ·
yw
CWC db
where CWC and CWD are the cut-off and the damping coefficient, respectively. The Eötvös
number coefficient Cw is provided by Equation (58). From numerical simulations, the au-
Fluids 2023, 8, 91 16 of 41
thors found that good agreement with the experimental data was obtained for CWC = 10,
CWD = 6.8, and p = 1.7.
Figure 10 shows the influence of the above described non-drag forces on the numerical
prediction of the local void fraction profile [45].
Figure 10. Influence of Non-Drag-Forces (NDF), (with the exception of the virtual mass force) on the
local void fraction profile. Reprinted with permission from [45].
Here, D j /Dt, Dk /Dt denote the total derivative with respect to time for the indicated
phase. As reported by Tabib et al. (2008) [46], the contribution of virtual mass is negligible
for large-diameter bubble columns (D > 0.15 m). A common practice is to set a constant
value CV M = 0.5, which has been obtained experimentally for isolated spherical bubbles,
as the virtual mass coefficient.
for all sizes together. Conversely, the iMUSIG model considers several velocity groups and
solves a separate momentum equation for each [47].
An uncertainty in bubble column design and scale-up arises from a lack of under-
standing of the phenomena governing bubble size distribution (BSD), and consequently
the column fluid dynamics and interfacial area. Bubble size distributions depend on many
phenomena occurring in the bubble column, including the balance between the coalescence
and breakup rate, the pressure change, and the gas–liquid mass transfer. In particular,
the bubble size distribution mainly results from the coalescence and breakup phenomena
occurring in the bubble column.
Consequently, the dispersed phase characteristic properties vary in time and space,
meaning that a constant bubble size distribution represents a strong assumption. To account
for the changes in the particle population, a balance equation can be included in addition
to the mass, momentum, and energy balances within the contest of MUSIG and iMUSIG
models. This concept is referred to as the Population Balance Model (PBM).
A population balance model can be described by a transport equation reflecting
particles entering or leaving a control volume via several mechanisms. The bubble number
density function, called the population balance equation (PBE), reads [48]:
∂ ∂ ∂ ∂
n(~x, Vb , t) + [n(~x, Vb , t)ub (~x, Vb )] + n(~x, Vb , t) Vb (~x, Vb ) = S(~x, Vb , t) (63)
∂t ∂z ∂Vb ∂t
where n(~x, Vb , t) is the bubble density distribution function that specifies the probable
number density of bubbles at a given time (t) in the spatial range (d(~x )) about a position ~x
for a bubble volume between Vb and Vb + d(Vb ) at time t. Finally, S(~x, Vb , t) is the following
source term:
S(~x, Vb , t) = Sb + Sc + S ph + S p + Sm + Sr (64)
where Sb , Sc , S ph , S p , Sm , and Sr are the bubble source/sink terms due to breakup, coales-
cence, phase change, pressure change, mass transfer, and reaction, respectively.
The breakup source term (Sb ) reads
Z ∞
Sb = m(Vb0 ) β(Vb , Vb0 ) g(V 0 )n(Vb0 , t) − g(Vb )n(Vb , t) (65)
v
where the first term is the birth rate of bubbles of volume Vb due to the breakup of bubbles
with volumes larger than Vb , the second term is the death rate of bubbles of volume Vb due
to breakup, m(Vb0 ) is the mean number of daughter bubbles produced by the breakup of a
parent bubble with volume Vb0 , g(Vb0 ) is the breakage frequency (the fraction of particles of
volume Vb0 breaking per unit time), and β(Vb , Vb0 ) denotes the daughter distribution factor
of a particle of volume Vb breaking from a parent bubble of volume Vb0 .
The coalescence source term (Sc ) reads
Z v Z ∞
1
Γ Vb − Vb0 , Vb0 Vb − Vb0 , t Vb0 , t dVb0 Γ Vb , Vb0 n Vb0 , t dVb
Sc = n n − n(Vb , t) (66)
2 0 0
where the first term is the birth rate of bubbles of volume Vb due to the coalescence of
bubbles of volume Vb − Vb0 and Vb0 , the second term is the death rate of bubbles of volume
Vb due to coalescence with others bubbles, and Γ(Vb , Vb0 ) is the coalescence rate between
bubbles of volume Vb and Vb0 .
Neglecting the interfacial mass transfer, the reaction contribution, the phase variations,
and the changes due to pressure gradients, the coalescence and breakup phenomena are
the only items of interest. Considering a steady state, it follows that
∂ni
ub,i = Si (~x, Vb ) = Sc + Sb (67)
∂z
Fluids 2023, 8, 91 18 of 41
2/9
h i
q
2/3 [ D∗2/3 −Λ5/3 ] (1− D∗3 ) −Λ5/3
β0 (εdb,j ) −12σ/(ρc db,j ) β=
g = Kg n with R Dmax h i
Martïnez and Bazän (1999) [52,53] db,j D∗ [ D ∗2/3 −Λ5/3 ] (1− D ∗3 )2/9 −Λ5/3 d( D ∗ )
min
β0 = 8.2 [54], K g = 0.25 with D ∗ =
db,j 0
db,i , Λ = ( β0 ρ L ), β = 8.2 [54]
12σ
" !#2
√ 22/5 db,j ρ3/5 ε2/5
d5/3 19/15 7/5
9 L
Lehr et al. (2002) [55] b,j ε ρL 2σ9/5 exp − 4 ln
σ3
g = 0.5 σ7/5
exp −
ρ L d9/5
b,j ε
6/5 L3
β = √π1 f 1/15 d ρ3/5 ε2/5
bv 2 b,i 1
1+erf 32 ln 3/5
σ
db,i and db,j refer to the parent and daughter bubble diameters, respectively [m]; λε is the turbulent eddy size [m].
where db,i and db,j are the diameters of the two colliding bubbles.
The collision frequency represents the number of collisions between bubbles per unit
of time. For a turbulent bubbly flow, there are five mechanisms of collision (Figure 12):
1. Turbulence-induced collisions occur as a result of the random motion of bubbles
caused by turbulent fluctuations.
2. Viscous shear-induced collisions are generated by global liquid velocity gradients,
meaning that bubbles in a location of high liquid velocity may collide with those in a
region of low liquid velocity.
Fluids 2023, 8, 91 20 of 41
3. Eddy-capture, in which collision events are produced by the shear rate of the flow in
the turbulent eddy.
4. Bouyancy-induced collisions, in which collisions may occur because bubbles of differ-
ent size have different rising velocities.
5. Wake entrainment collisions, which may result when bubbles are accelerated by the
wake region behind a large spherical-cap bubble.
Thus, the overall rate of collision events is determined by the sum of the effects of
all the listed mechanisms. Typically, only turbulence-induced collisions are assumed to
be predominant. The others are frequently insignificant except at very high superficial
gas velocities (where, for example, wake entrainment effects cannot be ignored due to the
formation of large spherical-cap bubbles).
Bubble collisions do not always lead to coalescence, as bubbles may bounce off each
other or separate after the collision. Hence, the efficiency term (h(db , d0b )) is introduced in
the definition of the coalescence kernel. Three models for calculation of the coalescence
efficiency have been introduced:
1. Film drainage model: first proposed by Shinnar and Church (1960) [57], this is a
widely used model of coalescence efficiency. When two bubbles collide, a thin liquid
film is trapped between their surfaces and is progressively drained. If the contact
time is sufficiently high, the liquid film reaches a minimum thickness, then ruptures,
causing coalescence.
2. Energy model: first proposed by Howarth (1964) [58], this model is based on the
concept of collision energy, in which a higher collision energy indicating a higher
probability of coalescence.
3. Critical approach velocity model: in this model, collisions result in coalescence phe-
nomena if the approach velocity of the bubbles exceeds a certain critical value; other-
wise, they bump into or bounce off of each other, and do not coalesce.
Several different coalescence frequency and efficiency models have been proposed; the
ones most commonly used in bubble column simulations are presented in Table 3.
Fluids 2023, 8, 91 21 of 41
orthogonal to the unknown number density function. The functional form (for a
univariate problem with φ as internal coordinate) reads as follows:
Z ∞ N
0
f (φ)n(φ)dφ ≈ ∑ f ( L i ) wi (70)
i =1
where the weights (ωi ) and abscissas (Li ) are determined through the Product–
Difference (PD) algorithm from the lower-order moments [64]. When the weights
and abscissas are known, the source term due to coalescence and breakup can be
calculated and the transport equation for moments can be solved. Finally, starting
from the moments of the distribution, it is possible to solve the inverse problem of
reconstructing the bubble size distribution [62].
When dealing with a multivariate population balance equation, for which the product–
difference algorithm can not be applied, other extensions of QMOM are available,
such as the Conditional Quadrature Method of Moments (CQMOM) or the Direct
Quadrature Methods of Moments (DQMOM). In DQMOM, the transport equations are
directly solved for the weights and nodes of the quadrature approximation, whereas
CQMOM represents the multivariate extension of QMOM. Moreover, in CQMOM
closure is achieved by means of multivariate quadrature approximation, and the
transport equations for the moments of the distributions are solved [65].
From the above discussion, it is clear that for univariate problems (which is the
case for bubble columns when the mass transfer is not modelled) the class method and
quadrature method of moments are the only viable candidates for solving the population
balance equation.
In the class method, a high number of bubble classes (i.e., 12–18) are required to de-
scribe the evolution of the particle population accurately, which incurs high computational
costs. On the contrary, in the quadrature method of moments approach it is sufficient to
solve a limited number of moments (i.e., 4–8), yielding a reduction in the dimensionality
of the problem compared to the class method, in turn leading to lower computational
costs [62].
turbulence. This approximation considers the Reynolds stresses proportional to the mean
rates of deformation of the fluid elements. The generic component of the viscous stress
tensor τ for a Newtonian fluid results in the following expression:
!
∂U i ∂U j 2
τij = −ρuı0 u0 = µt + − ρkδij (71)
∂x j ∂xi 3
where δij is the Kronecker delta and k = 12 u02 + v02 + w02 is the turbulent kinetic energy
per unit mass. The first term on the right-hand side of Equation (71) contains the turbulent
eddy viscosity µk . For k − ε models, the turbulent viscosity is provided by
k2
µt = ρCµ (72)
ε
where Cµ is a dimensionless constant, often taken as 0.09. With RANS models, the com-
putational resources needed for reasonably accurate flow simulations are modest, making
them the mainstay of many engineering flow calculations. Multiphase turbulence RANS
models are derived directly from their single phase equivalents.
The isotropic turbulence assumption for the core of the flow in RANS models is not
strictly satisfied for bubble columns, as the velocity fluctuations in the gravity direction
are normally twice those in the other directions [67]. As a result, the turbulence generated
by the rising bubbles is mainly anisotropic. RANS models fail to exactly represent the
directional effects of the Reynolds stress field; the Reynolds Stress Model (RSM) can be
used to address these problems adequately, as it incorporates anisotropy by solving the
transport equation of each Reynolds stress component. This increases the computer storage
and run time, and as a result has found little success in the simulation of bubbly flows.
Nonetheless, RSM seems a promising tool to provide more realistic numerical prediction
of turbulence quantities that can be used as inputs in bubble coalescence and breakage
models [68,69].
A further option is the use of Large Eddy Simulation (LES), an approach that uses an
unsteady flow computation to resolve only the turbulent structures that have a length scale
larger than a certain cutoff, then describes the smaller eddies and their effects on the larger
scales by means of Sub-Grid-Scale (SGS) models. By resolving only large portions of the
turbulent motion, LES appears to be more suitable for accounting for anisotropic structures
in bubble columns; however, this comes at the cost of a much larger volume of calculations
compared to RANS methods. Another issue is selection of the proper mesh size, which
when using LES should be bounded within a certain range in order to obtain a correct
filter cutoff width; a very dense grid is commonly required. Thus, LES may be unable to
consistently represent the correct sub-grid-scale stress for various flow situation [46].
DNS does not include any turbulence modelling, and directly resolves the instanta-
neous Navier–Stokes equations while considering all turbulence scales of motion and even
the fastest fluctuations. Unfortunately, its enormous computational and memory storage
requirements make its applicability to industrial-scale reactors unfeasible.
Bubble-Induced Turbulence
Experimental observations have found that the sources of the bulk liquid turbulence
can be divided into two categories, namely, (1) Shear-Induced Turbulence (SIT) and (2)
Bubble-Induced Turbulence (BIT). The former is independent of the bubble size, whereas
the presence of gaseous bubbles generates the latter. When a bubble rises in a pool of liquid
in a turbulent flow regime, a portion of its pressure energy is converted into liquid phase
turbulence, then into internal energy at the level of the Kolmogorov scale [70].
Within the context of the RANS k − ε and k − ω models, it is common practice to
account for the influence of the dispersed phase on the liquid phase turbulence by adding
extra source terms (Sk,L , Sε,L , or Sω,L ) to the transport equations for the turbulence quantities.
A plausible approximation is that all energy lost by bubbles due to the drag force is
Fluids 2023, 8, 91 24 of 41
converted to turbulent kinetic energy (k) in the wake of the bubble [71]. Hence, the k-
source becomes
Sk,L = ~FD,L→G |~uG − ~u L | (73)
The ε-equation source term is obtained by dividing the k-equation source term by a
characteristic time scale (τp ):
Sk,L
Sε,L = CεB (74)
τp
The ε-source term can be transformed into into an equivalent ω-source term as follows:
1 ω
Sω,L = Sε,L − L Sk,L (75)
Cµ k L kL
Modelling of the time scale (τp ) proceeds largely based on dimensional analysis [71].
The time scale factors commonly used in numerical simulation of bubble columns are
presented in Table 4.
Table 4. Time scale factors commonly used in numerical simulation of bubble columns.
Unlike the model presented in Table 4, Sato and Sekoguchi (1975) [77] did not explicitly
add source terms in the turbulence transport equations, rather, they tried to incorporate the
effects of bubble-induced turbulence in the turbulent viscosity, proposing the relation
k2L
νL = Cµ + 0.6αG db |~uG − ~u L |. (76)
εL
4. Literature Survey
Tables 5–8 collect different studies regarding CFD simulations of bubble columns vali-
dated with experimental data; Tables 5–7 refer to the operating conditions, physical settings,
and numerical settings, respectively. Conversely, Table 8 focuses on model performance.
To derive a consistent comparison between different studies, the values in Table 8 have
been computed by extracting published values using a consistent procedure. The CFD
accuracy is assessed by computing the relative error (δ) between numerical (ωCFD ) and
experimental (ωEXP ) data, as follows:
ωCFD − ωExp
δ= (77)
ωExp
When local validation wis performed with N local quantities (i.e., local void fraction
profile and local liquid velocity profile), the error (δ) is computed as follows:
N ωCFD,i − ωExp,i
1
δ=
N ∑ ωExp,i
(78)
i =1
Pfleger and Becker (2001) [74] performed 3D transient simulations of a 0.288-m inner di-
ameter bubble column operating in the homogeneous flow regime using a mono-dispersed
approach. Concerning the interfacial forces, they considered only the drag force with a
constant drag coefficient CD = 0.44. The standard k − ε model was used for turbulence in
the liquid phase. Their main aim was to study the effect of BIT on the correctness of CFD
Fluids 2023, 8, 91 25 of 41
predictions. The local liquid velocity and local void fraction showed the positive impact
of the BIT model, with δ = 30.29% (δ = 51.75% without BIT) and δ = 4.21% (δ = 11.14%
without BIT), respectively. The simulation results, including BIT or not, were found to
over-predict the global gas holdup, which the authors attributed to the simplified gas
inlet modelling.
Chen et al. (2005) [78] carried out 2D axisymmetric transient simulations in a 0.19-m
inner diameter bubble column operating in the heterogeneous flow regime. The main focus
was studying the influence of different coalescence and breakup models. In particular, they
compared the coalescence efficiency of Prince and Blanch (1990) [50], Chesters (1991) [60],
and Luo (1993) [61] and the breakup kernels of Luo and Svendsen (1996) [51] and Martinez
and Bazan (1999) [52,53]. The authors found that different bubble breakup and coalescence
closures did not significantly impact the simulation results. However, the best agreement
with the experimental values was provided by the breakup model of Luo and Svendsen
(1996) [51] coupled with the coalescence efficiency of Prince and Blanch (1990) [50] concern-
ing the local void fraction (δ = 22.74%) and the breakup model of Martinez–Bazan [52,53]
coupled with the coalescence efficiency of Luo (1993) [61] regarding the local liquid velocity
(δ = 51.44%).
Ekambara et al. (2010) [79] compared several different turbulence models: standard
k − ω, standard k − ε, k − ε RNG, RSM, and LES. They considered a mono-dispersed
approach including all the non-drag forces; turbulent dispersion was considered only
for the RANS methods. Concerning the local void fraction profile, LES provided the
best agreement with the experimental data in the near-sparger region (δLES = 18%,
δRSM = 29.55%, δk−εRNG = 31.44%, δk−ε = 34% and δk−ω = 36.24%), where the flow
is more anisotropic. The same was found for the local liquid velocity (δLES = 32.08%,
δRSM = 53.59%, δk−εRNG = 57.06%, δk−ε = 46.05%, and δk−ω = 50.65%). No remarkable
differences were found in the fully-developed region, indicating that the RANS approaches
perform well when the objective is to understand the steady and time-averaged features of
the flow.
Ziegenhein et al. (2013) [45] simulated a cylindrical bubble column operating in the
homogeneous flow regime considering mono-dispersed and iMUSIG (two bubble classes
without coalescence and breakup) approaches. Moreover, they analysed the influence of
non-drag forces, with the exception of the virtual mass, which was not included in the
simulations. Considering a superficial gas velocity of 0.15 cm/s, the mono-dispersed and
poly-dispersed approaches with non-drag forces performed similarly, with δ = 22.02%
and δ = 20.89%, respectively. The poly-disperse approach neglecting the non-drag force,
despite the error being slightly lower than the other cases (δ = 18.79%), predicted an overly
strong centre-picked void fraction profile. When increasing the superficial gas velocity,
the poly-disperse models with and without non-drag forces provided similar results (at
UG = 0.5 cm/s, δwithout NDF = 32.41%, and δwith NDF = 15.56%, respectively). How-
ever, the model including the non-drag forces performed better near the walls; conversely,
neglecting the non-drag forces resulted in better predicting the void fraction in the col-
umn centre.
Ziegenhein et al. (2013) [71] carried out transient simulations considering non-drag
forces and a swarm factor, fixed poly-dispersity (modelled using the iMUSIG model), and
BIT. In particular, they compared the BIT models of Morel (1997) [72], Troshco (2001) [73],
Politano (2003) [75], Rezehak (2012) [76], and Sato (1975) [77]. Regarding the void fraction
profile at UG = 0.3 cm/s, the results provided by the mono-dispersed model without
swarm factor matched very well with the experimental data (δ = 4.19%), and the inclusion
of BIT did not improve model prediction. When increasing the superficial gas velocity to
UG = 1.3 cm/s, the mono-dispersed treatment was not able to reproduce the gradient of
the gas volume fraction near the wall, which was better reproduced by the poly-dispersed
treatment (δMONO = 14.42% and δPOLY = 6.48%; BIT of Rzehak (2013) [76]). The same
was found for the superficial liquid velocity (δMONO = 44.27% and δPOLY = 30.20%; BIT
Fluids 2023, 8, 91 26 of 41
of Rzehak (2013) [76]). Regarding the different BIT models, the Rzehak (2013) [76]) model
performed better than the others.
Xu et al. (2013) [80] simulated a large-diameter bubble column operating in the hetero-
geneous flow regime using mono-dispersed, MUSIG, and iMUSIG approaches. In the latter
two cases, coalescence and breakup were considered by applying the coalescence model
of Luo (1993) [61] and the breakup kernel of Luo and Svendesen (1996) [51]. Moreover,
they studied the influence of the lift force by adopting the correlation of Tomiyama et al.
(2002) [32]. The Shiller and Naumann (1933) [14] drag coefficient was used in the single-size
model, and the Ishii and Zuber (1979) [17] drag coefficient was used for the MUSIG and
iMUSIG simulations. None of the mono-dispersed, MUSIG, and iMUSIG models neglect-
ing the lift force were able to reproduce the radial distribution of the local gas holdup,
with δ = 21%, δ = 21.44%, and δ = 21.53%, respectively. When the lift force was included,
the best agreement with the experimental data was provided by the iMUSIG model with
the drag coefficient of Ishii and Zuber (1979) [17] (δ = 9.68%). No remarkable differences
between the models were found in predicting the local liquid velocity profile.
McCLure et al. (2013) [81], simulating a cylindrical bubble column with a mono-
disperse approach, investigated the influence of the mean bubble diameter, lift force,
and BIT. Considering the void fraction profile at a superficial gas velocity of UG = 8 cm/s
and a bubble diameter of db = 4 mm, the inclusion of the lift force resulted in a bet-
ter agreement with the experimental data (δLIFT = 15.37% and δNO LIFT = 23.31%).
The lift force slightly influenced the results when db = 6 mm (δLIFT = 20.46% and
δNO LIFT = 22.29%). Regarding the BIT, the CFD prediction of the void fraction profile was
in reasonable agreement with experimental data when the BIT was incorporated into the
model (δBIT = 20.80%, δNO BIT = 75.80%).
Masood at al. (2014) [82] studied the hydrodynamics of a square bubble column and
tested different drag closures. In particular, they considered the models of Shiller and
Naumann (1933) [14], Grace et al. (1976) [16], and Ishii and Zuber (1979) [17], with a
constant drag coefficient of CD = 0.44. Turbulence was modelled using the k − ε RNG
model, including BIT of Sato (1975) [77]. The constant drag coefficient was the worst in
predicting the global gas holdup curve (δ = 22.75%), followed by the simplistic correlation
proposed by Schiller and Naumann (1933) [14] (δ = 14.76%). The Ishii and Zuber (1979) [17]
model was found to be slightly superior to that of Grace et al. (1976) [16], with δ = 6.18%
and δ = 9.45%, respectively.
Liu et al. (2014) [83] performed 2D-axisymmetric simulations of a bubble column
operating in the heterogeneous flow regime with a MUSIG model considering coalescence
and breakup. They included only the drag (Tomiyama et al. (1998) [19] model with the
swarm factor of Ishii and Zuber (1979) [17]) and lift forces (Tomiyama et al. (2002) [32]
and Behzadi et al. (2004) [37] models); the other forces were neglected. For the turbulence,
they compared the standard k − ε and RSM models. The authors performed a sensitivity
analysis on the number of bubble classes through 10 and 20 classes for the k − ε and RSM
models with BIT. They found that increasing the number of classes did not significantly
increase agreement with the experimental data; for example, concerning the void fraction
profile and considering the k − ε turbulence model, δ = 8.66% and δ = 5.17% for 10 classes
and 20 classes, respectively. However, much more computational time was needed when
increasing the number of bubble classes to 20. Regarding the comparison between the
lift correlations, a much better agreement with the experimental void fraction profile was
achieved using the Tomiyama et al. (2002) [32] lift model. For example, considering the
RSM turbulence model, δTOMIYAMA = 9.60% and δBEHZADI = 27.70%. Such an increment
in the relative error can be explained by the fact that the model of Behzadi et al. (2004) [37]
does not predict the change in the sign of the lift coefficient, resulting in a flat profile
of the void fraction. No remarkable differences were found between the k − ε and RSM
turbulence models.
Syed at al. (2017) [84] carried out 2D-axisymmetric simulations of a bubble column
operating in the homogenous flow regime. They performed a sensitivity analysis on the Luo
Fluids 2023, 8, 91 27 of 41
(1993) [61] coalescence kernel by changing the coalescence parameter cλ ; see Table 3 from
0.1 to 1.1. Moreover, they compared the breakup kernels of Luo and Svendsen (1996) [51]
and Lehr et al. (2002) [55]. The results showed that the radial profiles of void fraction and
local liquid velocity were significantly affected by these parameters. Considering the local
void fraction profile at UG = 0.38 cm/s and the Luo and Svendsen (1996) [51] breakup
kernel, δc0 =0.1 = 10.5% and δc0 =1.1 = 23.14%; consequently, a lower coalescence parameter
(cλ ) led to better prediction of the experimental data. The same was found for the global
gas holdup and local liquid velocity profile. Comparing the two breakup models, it was
found that the Lehr et al. (2002) [55] model improved prediction of the local void fraction
as compared to the Luo and Svendsen (1996) [51] model (δcλ =0.3,LUO and SVENDSEN = 13.3%
and δcλ =0.3,LEHR = 3.65%). Moreover, when the breakup model of Lehr et al. (2002) [55] was
considered, the results regarding the local void fraction profiles were slightly influenced
by cλ ; on the contrary, the authors found that a decrease in the coalescence parameter
negatively influenced the prediction of the local liquid velocity profiles.
Gemello et al. (2018) [30] proposed an interesting modification of the Simonnet et al.
(2008) [27] swarm factor that was suitable for very high volume fractions while avoiding
stability problems encountered in the original formulation. The results obtained with the
proposed swarm model matched very well with the experimental data. Considering the gas
holdup, the relative error was δ = 2.83% (δ = 29.82% with the Simonnet et al. (2008) [27]
swarm factor). The authors obtained good prediction of the local void and liquid velocity
profiles, with a relative error at UG = 16 cm/s of δ = 4.07% and δ = 25.18%, respectively.
Zhang et al. (2020) [85] studied the influence of non-drag forces and BIT by simulating
a cylindrical bubble column operating in the heterogeneous flow regime with a MUSIG
model coupled with the coalescence kernel of Luo (1993) [61] and the breakup model of Luo
and Svendsen (1996) [51]. Considering the local void fraction profile, good agreement with
the experimental data was obtained considering all the non-drag forces and BIT (δ = 8.07%).
When the lift force was not considered, the relative error increased to δ = 42.69%, indicating
that the lift force could not be neglected. Furthermore, neglecting the turbulent dispersion
and wall lubrication forces reduced the model’s predictive capacity, increasing the relative
error with respect to the experimental data ( δ = 23.26% and δ = 12.69%, respectively). No
evident differences were found when neglecting the BIT and virtual mass force.
As can be observed from Table 8, the accuracy of CFD models has been determined
by comparing the numerical results against global gas holdup, local void fraction, and
in a few cases the local liquid velocity, without providing any information about model
performance with respect to turbulent quantities. Indeed, measuring the local liquid
velocity and turbulence in large-scale bubble columns with optical methods is complex,
and is usually limited to low gas holdups and thin geometries. To overcome this issue,
Deen et al. (2000) [86] used Particle Image Velocimetry (PIV) and Laser Doppler Velocimetry
(LDA) to study turbulent quantities in a square bubble column. Similarly, Ziegenhein et al.
(2019) [87] designed a Particle Shadowgraph Velocimetry (PSV) technique and tested it in a
large-diameter bubble column to provide data for CFD model validation.
The main conclusion that emerges from the above literature survey is that several
different methods have been applied when dealing with the numerical simulation of two-
phase bubble columns. As can be noted from Tables 2 and 3, different strategies in terms
of physical models, interfacial momentum exchange, turbulence models, and numerical
settings have been adopted, with the result that while certain models perform better than
others, no general conclusion can be reached. Moreover, available numerical studies are of-
ten limited to air–water systems, with the consequence that a model which has outstanding
performance on air–water systems may not be suitable in other cases. Consequently, fully
predictive CFD models for bubble columns are far from being developed, and more effort
must be spent in order to overcome the lack of knowledge regarding numerical simulation
of bubble columns.
Fluids 2023, 8, 91 28 of 41
Ref. Year Code D * [m] Sparger AR [-] Fluids UG [m/s] Flow Regime
Ring **
[74] 2001 Ansys CFX-4.3 0.288 8.68 Air/water 0.5 Homogeneous
d0 = 0.7 mm
Perforate plate
[78] 2005 Ansys FLUENT 0.19 5.05 Air/water 12 Heterogeneous
d0 = 3.3 mm
Homogeneous→
[46] 2008 Ansys CFX-10.0 0.60 Perforated plate *** - Air/water 1.2 → 9.6
heterogeneous
Perforated plate ***
[79] 2010 Ansys CFX-10.0 0.15 6 Air/water 0.2 Homogeneous
d0 = 1.96 mm
Perforate plate
[45] 2013 Ansys CFX-13.0 0.288 8.68 Air/water 0.15 → 1 Homogeneous
d0 = 0.7 mm
Perforate plate **
[80] 2013 Ansys FLUENT-14.0 0.44 4 Air/water 10 Heterogeneous
d0 = 0.77 mm
Perforated plate ***
[81] 2013 Ansys CFX-14.5 0.19 2.63 Air/water 8 → 12 Heterogeneous
d0 = 1 mm
[71] 2014 Ansys CFX-13.0 0.24 × 0.72 Needle ** 5.08 Air/water 0.3 → 1.3 Homogeneous
[82] 2014 Ansys CFX-14.0 0.15 × 0.15 Full opening ** 2.74 Air/water 5 → 12
Perforate plate **
[83] 2014 OpenFOAM 0.2 5 Air/water 10 Heterogeneous
d0 = 1.2 mm
[88] 2014 Ansys CFX-14.5 0.24 × 0.72 Needle ** 5.08 Air/water 13
Table 6. Cont.
[81] k−ε No Pfleger and Becker (α) - High resolution schemes Second-order implicit 3D cylindrical 58,800
NO (β)
[82] k − ε RNG No Sato - - - 3D rectangular 46,080
k − ε (α) k − ε (α)
[83] Sato PISO - - 2D axisymmetric -
RSM (β) RSM (β)
Rzehak (A)
[88] k − ω SST No - High resolution schemes second-order implicit 3D rectangular 200,000
Sato (B)
No (C)
[92] k − ε RNG k − ε RNG - SIMPLE - - 3D cylindrical 342,230
Fluids 2023, 8, 91 32 of 41
Table 7. Cont.
Table 8. Cont.
Table 8. Cont.
5. Conclusions
This paper has presented a detailed review of CFD modelling of two-phase bubble
columns. The models most commonly adopted for interfacial momentum exchange, bubble-
induced turbulence, coalescence, and breakup are described. A quantitative comparison
between the different modelling approaches is presented in Section 4, considering various
studies from the literature and computing the relative errors between the CFD predictions
and the experimental data.
On the basis of the literature review presented in this work, the following recommen-
dations are made for future studies:
1. Concerning the interfacial forces, the momentum transfer between the phases is domi-
nated by the drag force. For a proper description of the drag coefficient, the models of
Tomiyama et al. (1998) [19], Grace et al. (1976) [16], and Ishii and Zuber (1978) [17]
can be implemented; however, they should be corrected with a swarm factor. When
presenting numerical studies, a sensitivity analysis among the different models should
be performed.
Fluids 2023, 8, 91 36 of 41
The lift, wall lubrication, and turbulence dispersion forces should be added to the
model to obtain more accurate solutions. In particular, a correlation that predicts the
change in the sign of the lift coefficient should be considered.
2. Concerning the turbulence modelling, RANS models, in particular the k − ε RNG and
the k − ω SST models, provide satisfying results in terms of average quantities. The
LES turbulence model provides better results in the near-sparger region, where the
flow is more anisotropic. However, no remarkable differences compared to the RANS
methods have been highlighted in the fully developed region.
3. The modelling approach of the dispersed phase (i.e., mono-dispersed, MUSIG, iMUSIG,
PBM) should always be related to the simulated flow regime. A mono-dispersed
approximation applies at very low superficial gas velocities. Conversely, multiple-size
models that include coalescence and breakup should be considered.
4. Regarding the numerical settings, high-order resolution discretization schemes should
be used in order to prevent or reduce numerical discretization errors as much
as possible.
Finally, it can be concluded that detailed comprehension of the phenomena governing
multiphase flows remains needed, and additional effort should be spend on developing a
fully predictive a priori CFD model.
Author Contributions: N.V. conducted the literature survey, post-processed the literature data, and
wrote the paper. G.B. and R.M. were responsible for conceptualization, methodology, and revision.
All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: The authors declare no conflict of interest.
Nomenclature
BIT Bubble-Induced Turbulence
CFD Computational Fluid Dynamics
CM Class Method
DNS Direct Numerical Simulation
LDA Laser Doppler Velocimetry
LES Large Eddy Simulation
NDF Number Density Function
PBM Population Balance Model
PBE Population Balance Equation
PIV Particle Image Velocimetry
PSV Particle Shadowgraph Velocimetry
QMOM Quadrature Method of Moment
RANS Reynolds-Averaged NavierStokes
RSM Reynolds Stress Models
~u Local velocity [m/s]
CD Drag coefficient [-]
CL Lift coefficient [-]
CTD Turbulent dispersion coefficient [-]
CW L Wall lubrication coefficient [-]
CV M Virtual mass coefficient [-]
db Bubble diameter [m]
deq Equivalent bubble diameter [m]
D ∗H Non-dimensional diameter [-]
∗
Dcr Critical non-dimensional diameter [-]
DH Hydraulic diameter [m]
E Bubble aspect ratio [-]
Eo Eötvös number [m]
Fluids 2023, 8, 91 37 of 41
References
1. Kantarci, N.; Borak, F.; Ulgen, K.O. Bubble column reactors. Process Biochem. 2005, 40, 2263–2283. [CrossRef]
2. Buffo, A.; Marchisio, D.L. Modeling and simulation of turbulent polydisperse gas-liquid systems via the generalized population
balance equation. Rev. Chem. Eng. 2014, 30, 73–126. [CrossRef]
3. Khan, K.I. Fluid Dynamic Modeling of Bubble Column Reactor. Ph.D Thesis, Politecnico di Torino, Torino, Italy, 2014.
4. Wu, B.; Firouzi, M.; Mitchell, T.; Rufford, T.E.; Leonardi, C.; Towler, B. A critical review of flow maps for gas-liquid flows in
vertical pipes and annuli. Chem. Eng. J. 2017, 326, 350–377. [CrossRef]
5. Kitscha, J.; Kocamustafaogullari, G. Breakup criteria for fluid particles. Int. J. Multiph. Flow 1989, 15, 573–588. [CrossRef]
6. Brooks, C.S.; Paranjape, S.S.; Ozar, B.; Hibiki, T.; Ishii, M. Two-group drift-flux model for closure of the modified two-fluid model.
Int. J. Heat Fluid Flow 2012, 37, 196–208. [CrossRef]
7. Besagni, G. Bubble column fluid dynamics: A novel perspective for flow regimes and comprehensive experimental investigations.
Int. J. Multiph. Flow 2021, 135, 103510. [CrossRef]
8. Krepper, E.; Lucas, D.; Prasser, H.M. On the modelling of bubbly flow in vertical pipes. Nucl. Eng. Des. 2005, 235, 597–611.
[CrossRef]
9. Lapin, A.; Lübbert, A. Numerical simulation of the dynamics of two-phase gas–liquid flows in bubble columns. Chem. Eng. Sci.
1994, 49, 3661–3674. [CrossRef]
10. Buwa, V.V.; Deo, D.S.; Ranade, V.V. Eulerian–Lagrangian simulations of unsteady gas–liquid flows in bubble columns. Int. J.
Multiph. Flow 2006, 32, 864–885. [CrossRef]
Fluids 2023, 8, 91 38 of 41
11. Besbes, S.; El Hajem, M.; Aissia, H.B.; Champagne, J.; Jay, J. PIV measurements and Eulerian–Lagrangian simulations of the
unsteady gas–liquid flow in a needle sparger rectangular bubble column. Chem. Eng. Sci. 2015, 126, 560–572. [CrossRef]
12. Khan, I.; Wang, M.; Zhang, Y.; Tian, W.; Su, G.; Qiu, S. Two-phase bubbly flow simulation using CFD method: A review of models
for interfacial forces. Prog. Nucl. Energy 2020, 125, 103360. [CrossRef]
13. Suh, J.W.; Kim, J.W.; Choi, Y.S.; Kim, J.H.; Joo, W.G.; Lee, K.Y. Development of numerical Eulerian-Eulerian models for simulating
multiphase pumps. J. Pet. Sci. Eng. 2018, 162, 588–601. [CrossRef]
14. Schiller, L. A drag coefficient correlation. Zeit. Ver. Deutsch. Ing. 1933, 77, 318–320.
15. Frank. Viscous Fluid Flow; McGraw-Hill: New York, NY, USA, 1974.
16. Grace, J.R. Shapes and velocities of single drops and bubbles moving freely through immiscible liquids. Trans. Inst. Chem. Eng.
1976, 54, 167.
17. Ishii, M.; Zuber, N. Drag coefficient and relative velocity in bubbly, droplet or particulate flows. AIChE J. 1979, 25, 843–855.
[CrossRef]
18. Grevskott, S.; Sannaes, B.; Duduković, M.; Hjarbo, K.; Svendsen, H. Liquid circulation, bubble size distributions, and solids
movement in two-and three-phase bubble columns. Chem. Eng. Sci. 1996, 51, 1703–1713. [CrossRef]
19. Tomiyama, A. Struggle with computational bubble dynamics. Multiph. Sci. Technol. 1998, 10, 369–405.
20. Zhang, D.; Deen, N.; Kuipers, J. Numerical simulation of the dynamic flow behavior in a bubble column: a study of closures for
turbulence and interface forces. Chem. Eng. Sci. 2006, 61, 7593–7608. [CrossRef]
21. Dijkhuizen, W.; Roghair, I.; Annaland, M.V.S.; Kuipers, J. DNS of gas bubbles behaviour using an improved 3D front tracking
model—Drag force on isolated bubbles and comparison with experiments. Chem. Eng. Sci. 2010, 65, 1415–1426. [CrossRef]
22. Bridge, A.; Lapidus, L.; Elgin, J. The mechanics of vertical gas-liquid fluidized system I: Countercurrent flow. AIChE J. 1964,
10, 819–826. [CrossRef]
23. Wallis, G.B. One-Dimensional Two-Phase Flow; McGraw-Hill: New York, NY, USA, 1969.
24. Rusche, H.; Issa, R. The effect of voidage on the drag force on particles, droplets and bubbles in dispersed two-phase flow. In
Proceedings of the Japanese European Two-Phase Flow Meeting, Tshkuba, Japan, 25-29 September 2000.
25. Roghair, I.; Lau, Y.; Deen, N.; Slagter, H.; Baltussen, M.; Annaland, M.V.S.; Kuipers, J. On the drag force of bubbles in bubble
swarms at intermediate and high Reynolds numbers. Chem. Eng. Sci. 2011, 66, 3204–3211. [CrossRef]
26. Buffo, A.; Vanni, M.; Renze, P.; Marchisio, D.L. Empirical drag closure for polydisperse gas–liquid systems in bubbly flow regime:
Bubble swarm and micro-scale turbulence. Chem. Eng. Res. Des. 2016, 113, 284–303. [CrossRef]
27. Simonnet, M.; Gentric, C.; Olmos, E.; Midoux, N. CFD simulation of the flow field in a bubble column reactor: Importance of the
drag force formulation to describe regime transitions. Chem. Eng. Process. Process Intensif. 2008, 47, 1726–1737. [CrossRef]
28. McClure, D.D.; Norris, H.; Kavanagh, J.M.; Fletcher, D.F.; Barton, G.W. Validation of a computationally efficient computational
fluid dynamics (CFD) model for industrial bubble column bioreactors. Ind. Eng. Chem. Res. 2014, 53, 14526–14543. [CrossRef]
29. McClure, D.D.; Kavanagh, J.M.; Fletcher, D.F.; Barton, G.W. Experimental investigation into the drag volume fraction correction
term for gas-liquid bubbly flows. Chem. Eng. Sci. 2017, 170, 91–97. [CrossRef]
30. Gemello, L.; Cappello, V.; Augier, F.; Marchisio, D.; Plais, C. CFD-based scale-up of hydrodynamics and mixing in bubble
columns. Chem. Eng. Res. Des. 2018, 136, 846–858. [CrossRef]
31. Yang, G.; Zhang, H.; Luo, J.; Wang, T. Drag force of bubble swarms and numerical simulations of a bubble column with a
CFD-PBM coupled model. Chem. Eng. Sci. 2018, 192, 714–724. [CrossRef]
32. Tomiyama, A.; Tamai, H.; Zun, I.; Hosokawa, S. Transverse migration of single bubbles in simple shear flows. Chem. Eng. Sci.
2002, 57, 1849–1858. [CrossRef]
33. Wellek, R.; Agrawal, A.; Skelland, A. Shape of liquid drops moving in liquid media. AIChE J. 1966, 12, 854–862. [CrossRef]
34. Ziegenhein, T.; Lucas, D. The critical bubble diameter of the lift force in technical and environmental, buoyancy-driven bubbly
flows. Int. J. Multiph. Flow 2019, 116, 26–38. [CrossRef]
35. Besagni, G.; Inzoli, F.; De Guido, G.; Pellegrini, L.A. Experimental investigation on the influence of ethanol on bubble column
hydrodynamics. Chem. Eng. Res. Des. 2016, 112, 1–15. [CrossRef]
36. Ziegenhein, T.; Tomiyama, A.; Lucas, D. A new measuring concept to determine the lift force for distorted bubbles in low Morton
number system: Results for air/water. Int. J. Multiph. Flow 2018, 108, 11–24. [CrossRef]
37. Behzadi, A.; Issa, R.; Rusche, H. Modelling of dispersed bubble and droplet flow at high phase fractions. Chem. Eng. Sci. 2004,
59, 759–770. [CrossRef]
38. Simonin, O.; Viollet, P. Modelling of turbulent two-phase jets loaded with discrete particles. In Phenomena in Multiphase Flows;
Hemisphere Publishing Corporation: London, UK, 1990; Volume 1990, pp. 259–269.
39. Burns, A.D.; Frank, T.; Hamill, I.; Shi, J.M. The Favre averaged drag model for turbulent dispersion in Eulerian multi-phase flows.
In Proceedings of the 5th International Conference on Multiphase Flow, ICMF, Yokohama, Japan, 30 May–4 June 2004; Volunme 4,
pp. 1–17.
40. Lopez de Bertodano, M.; Moraga, F.; Drew, D.; Lahey, R., Jr. The modeling of lift and dispersion forces in two-fluid model
simulations of a bubbly jet. J. Fluids Eng. 2004, 126, 573–577. [CrossRef]
41. Antal, S.; Lahey, R., Jr.; Flaherty, J. Analysis of phase distribution in fully developed laminar bubbly two-phase flow. Int. J.
Multiph. Flow 1991, 17, 635–652. [CrossRef]
Fluids 2023, 8, 91 39 of 41
42. Tomiyama, A. Effects of Eotvos number and dimensionless liquid volumetric flux on lateral motion of a bubble in a laminar duct
flow. In Proceedings of the 2nd International Conference on Multiphase Flow, Kyoto, Japan, 3–7 April 1995; Volume 3.
43. Hosokawa, S.; Tomiyama, A.; Misaki, S.; Hamada, T. Lateral migration of single bubbles due to the presence of wall. In
Proceedings of the Fluids Engineering Division Summer Meeting, Montreal, QC, Canada, 14–18 July 2002; Volume 36150,
pp. 855–860.
44. Frank, T. Advances in computational fluid dynamics (CFD) of 3-dimensional gas-liquid multiphase flows. In Proceedings of
the NAFEMS Seminar: Simulation of Complex Flows (CFD)–Applications and Trends, Wiesbaden, Germany, 25–26 April 2005;
pp. 1–18.
45. Ziegenhein, T.; Rzehak, R.; Krepper, E.; Lucas, D. Numerical simulation of polydispersed flow in bubble columns with the
inhomogeneous multi-size-group model. Chem. Ing. Tech. 2013, 85, 1080–1091. [CrossRef]
46. Tabib, M.V.; Roy, S.A.; Joshi, J.B. CFD simulation of bubble column—An analysis of interphase forces and turbulence models.
Chem. Eng. J. 2008, 139, 589–614. [CrossRef]
47. Yeoh, G.; Tu, J. Basic theory and conceptual framework of multiphase flows. In Handbook of Multiphase Flow Science and Technology;
Springer: Berlin, Germany, 2017; pp. 1–47.
48. Liao, Y.; Lucas, D. A literature review on mechanisms and models for the coalescence process of fluid particles. Chem. Eng. Sci.
2010, 65, 2851–2864. [CrossRef]
49. Liao, Y.; Lucas, D. A literature review of theoretical models for drop and bubble breakup in turbulent dispersions. Chem. Eng. Sci.
2009, 64, 3389–3406. [CrossRef]
50. Prince, M.J.; Blanch, H.W. Bubble coalescence and break-up in air-sparged bubble columns. AIChE J. 1990, 36, 1485–1499.
[CrossRef]
51. Luo, H.; Svendsen, H.F. Theoretical model for drop and bubble breakup in turbulent dispersions. AIChE J. 1996, 42, 1225–1233.
[CrossRef]
52. Martínez-Bazán, C.; Montañés, J.L.; Lasheras, J.C. On the breakup of an air bubble injected into a fully developed 881 turbulent
flow. Part 1. Breakup frequency. J. Fluid Mech. 1999, 401, 157–182. [CrossRef]
53. Martínez-Bazán, C.; Montañés, J.L.; Lasheras, J.C. On the breakup of an air bubble injected into a fully developed 883 turbulent
flow. Part 2. Size PDF of the resulting daughter bubbles. J. Fluid Mech. 1999, 401, 183–207. [CrossRef]
54. Batchelor, G.K. The Theory of Homogeneous Turbulence; Cambridge University Press: Cambridge, UK, 1953.
55. Lehr, F.; Millies, M.; Mewes, D. Bubble-size distributions and flow fields in bubble columns. AIChE J. 2002, 48, 2426–2443.
[CrossRef]
56. Chen, J.; Li, F.; Degaleesan, S.; Gupta, P.; Al-Dahhan, M.H.; Dudukovic, M.P.; Toseland, B.A. Fluid dynamic parameters in bubble
columns with internals. Chem. Eng. Sci. 1999, 54, 2187–2197. [CrossRef]
57. Shinnar, R.; Church, J.M. Statistical theories of turbulence in predicting particle size in agitated dispersions. Ind. Eng. Chem. 1960,
52, 253–256. [CrossRef]
58. Howarth, W. Coalescence of drops in a turbulent flow field. Chem. Eng. Sci. 1964, 19, 33–38. [CrossRef]
59. Coulaloglou, C.; Tavlarides, L.L. Description of interaction processes in agitated liquid-liquid dispersions. Chem. Eng. Sci. 1977,
32, 1289–1297. [CrossRef]
60. Chesters, A. Modelling of coalescence processes in fluid-liquid dispersions: A review of current understanding. Chem. Eng. Res.
Des. 1991, 69, 259–270.
61. Luo, H. Coalescence, Breakup and Liquid Circulation in Bubble Column Reactors. Ph.D. Thesis, Department of Chemical
Engineering, The University of Trondheim, Trondheim, Norvey, 1993.
62. Sanyal, J.; Marchisio, D.L.; Fox, R.O.; Dhanasekharan, K. On the comparison between population balance models for CFD
simulation of bubble columns. Ind. Eng. Chem. Res. 2005, 44, 5063–5072. [CrossRef]
63. McGraw, R. Description of aerosol dynamics by the quadrature method of moments. Aerosol Sci. Technol. 1997, 27, 255–265.
[CrossRef]
64. Marchisio, D.L.; Vigil, R.D.; Fox, R.O. Quadrature method of moments for aggregation–breakage processes. J. Colloid Interface Sci.
2003, 258, 322–334. [CrossRef]
65. Petitti, M.; Vanni, M.; Marchisio, D.L.; Buffo, A.; Podenzani, F. Simulation of coalescence, break-up and mass transfer in a
gas–liquid stirred tank with CQMOM. Chem. Eng. J. 2013, 228, 1182–1194. [CrossRef]
66. Ma, T.; Lucas, D.; Ziegenhein, T.; Fröhlich, J.; Deen, N. Scale-Adaptive Simulation of a square cross-sectional bubble column.
Chem. Eng. Sci. 2015, 131, 101–108. [CrossRef]
67. Dhotre, M.; Deen, N.; Niceno, B.; Khan, Z.; Joshi, J. Large eddy simulation for dispersed bubbly flows: A review. Int. J. Chem.
Eng. 2013, 2013, 343276. [CrossRef]
68. Masood, R.; Rauh, C.; Delgado, A. CFD simulation of bubble column flows: An explicit algebraic Reynolds stress model approach.
Int. J. Multiph. Flow 2014, 66, 11–25. [CrossRef]
69. Parekh, J.; Rzehak, R. Euler–Euler multiphase CFD-simulation with full Reynolds stress model and anisotropic bubble-induced
turbulence. Int. J. Multiph. Flow 2018, 99, 231–245. [CrossRef]
70. Joshi, J. Computational flow modelling and design of bubble column reactors. Chem. Eng. Sci. 2001, 56, 5893–5933. [CrossRef]
Fluids 2023, 8, 91 40 of 41
71. Ziegenhein, T.; Lucas, D.; Rzehak, R.; Krepper, E. Closure relations for CFD simulation of bubble columns. In Proceedings of the
8th International Conference on Multiphase Flow—ICMF 2014, Jeju, Republic of Korea, 26–31 May 2013.
72. Morel, C. Turbulence modelling and first numerical simulations in turbulent two-phase flows. In Proceedings of the 11th
Symposium on Turbulent Shear Flows, Grenoble, France, 8–11 September 1997; Volume 3.
73. Troshko, A.; Hassan, Y. A two-equation turbulence model of turbulent bubbly flows. Int. J. Multiph. Flow 2001, 27, 1965–2000.
[CrossRef]
74. Pfleger, D.; Becker, S. Modelling and simulation of the dynamic flow behaviour in a bubble column. Chem. Eng. Sci. 2001,
56, 1737–1747. [CrossRef]
75. Politano, M.; Carrica, P.; Converti, J. A model for turbulent polydisperse two-phase flow in vertical channels. Int. J. Multiph. Flow
2003, 29, 1153–1182. [CrossRef]
76. Rzehak, R.; Krepper, E. Bubble-induced turbulence: Comparison of CFD models. Nucl. Eng. Des. 2013, 258, 57–65. [CrossRef]
77. Sato, Y.; Sekoguchi, K. Liquid velocity distribution in two-phase bubble flow. Int. J. Multiph. Flow 1975, 2, 79–95. [CrossRef]
78. Chen, P.; Sanyal, J.; Duduković, M. Numerical simulation of bubble columns flows: Effect of different breakup and coalescence
closures. Chem. Eng. Sci. 2005, 60, 1085–1101. [CrossRef]
79. Ekambara, K.; Dhotre, M. CFD simulation of bubble column. Nucl. Eng. Des. 2010, 240, 963–969. [CrossRef]
80. Xu, L.; Yuan, B.; Ni, H.; Chen, C. Numerical simulation of bubble column flows in churn-turbulent regime: Comparison of bubble
size models. Ind. Eng. Chem. Res. 2013, 52, 6794–6802. [CrossRef]
81. McClure, D.D.; Kavanagh, J.M.; Fletcher, D.F.; Barton, G.W. Development of a CFD model of bubble column bioreactors: Part
two—Comparison of experimental data and CFD predictions. Chem. Eng. Technol. 2014, 37, 131–140. [CrossRef]
82. Masood, R.; Delgado, A. Numerical investigation of the interphase forces and turbulence closure in 3D square bubble columns.
Chem. Eng. Sci. 2014, 108, 154–168. [CrossRef]
83. Liu, Y.; Hinrichsen, O. Study on CFD–PBM turbulence closures based on k–ε and Reynolds stress models for heterogeneous
bubble column flows. Comput. Fluids 2014, 105, 91–100. [CrossRef]
84. Syed, A.H.; Boulet, M.; Melchiori, T.; Lavoie, J.M. CFD simulations of an air-water bubble column: Effect of Luo coalescence
parameter and breakup kernels. Front. Chem. 2017, 5, 68. [CrossRef]
85. Zhang, X.B.; Yan, W.C.; Luo, Z.H. CFD-PBM Simulation of Bubble Columns: Sensitivity Analysis of the Nondrag Forces. Ind.
Eng. Chem. Res. 2020, 59, 18674–18682. [CrossRef]
86. Deen, N.G.; Hjertager, B.H.; Solberg, T. Comparison of PIV and LDA measurement methods applied to the gas-liquid flow in a
bubble column. In Proceedings of the 10th International Symposium on Applications of Laser Techniques to Fluid Mechanics,
Lisbon, Portugal, 10–13 July 2000; pp. 1–12.
87. Ziegenhein, T.; Lucas, D.; Besagni, G.; Inzoli, F. Experimental study of the liquid velocity and turbulence in a large-scale air-water
counter-current bubble column. Exp. Therm. Fluid Sci. 2020, 111, 109955. [CrossRef]
88. Ziegenhein, T.; Rzehak, R.; Lucas, D. Transient simulation for large scale flow in bubble columns. Chem. Eng. Sci. 2015, 122, 1–13.
[CrossRef]
89. Raimundo, P.M. Analysis and Modelization of Local Hydrodynamics in Bubble Columns. Ph.D. Thesis, Université Grenoble
Alpes, Grenoble, France, 2015.
90. Fletcher, D.F.; McClure, D.D.; Kavanagh, J.M.; Barton, G.W. CFD simulation of industrial bubble columns: Numerical challenges
and model validation successes. Appl. Math. Model. 2017, 44, 25–42. [CrossRef]
91. Saleh, S.N.; Mohammed, A.A.; Al-Jubory, F.K.; Barghi, S. CFD assesment of uniform bubbly flow in a bubble column. J. Pet. Sci.
Eng. 2018, 161, 96–107. [CrossRef]
92. Liang, X.F.; Pan, H.; Su, Y.H.; Luo, Z.H. CFD-PBM approach with modified drag model for the gas–liquid flow in a bubble
column. Chem. Eng. Res. Des. 2016, 112, 88–102. [CrossRef]
93. Degaleesan, S. Fluid Dynamic Measurements and Modeling of Liquid Mixing in Bubble Columns; Washington University in St. Louis:
St. Louis, MO, USA, 1997.
94. Menzel, T.; In der Weide, T.; Staudacher, O.; Wein, O.; Onken, U. Reynolds shear stress for modeling of bubble column reactors.
Ind. Eng. Chem. Res. 1990, 29, 988–994. [CrossRef]
95. Kulkarni, A.; Ekambara, K.; Joshi, J. On the development of flow pattern in a bubble column reactor: experiments and CFD.
Chem. Eng. Sci. 2007, 62, 1049–1072. [CrossRef]
96. bin Mohd Akbar, M.H.; Hayashi, K.; Hosokawa, S.; Tomiyama, A. Bubble tracking simulation of bubble-induced pseudoturbu-
lence. Multiph. Sci. Technol. 2012, 24, 197-222 [CrossRef]
97. McClure, D.D.; Kavanagh, J.M.; Fletcher, D.F.; Barton, G.W. Development of a CFD model of bubble column bioreactors: Part
one—A detailed experimental study. Chem. Eng. Technol. 2013, 36, 2065–2070. [CrossRef]
98. Deen, N. An Experimental and Computational Study of Fluid Dynamics in Gas-Liquid Chemical Reactors; Aalborg University Esbjerg:
Esbjerg, Denmark, 2001.
99. Rampure, M.R.; Kulkarni, A.A.; Ranade, V.V. Hydrodynamics of bubble column reactors at high gas velocity: experiments and
computational fluid dynamics (CFD) simulations. Ind. Eng. Chem. Res. 2007, 46, 8431–8447. [CrossRef]
100. Hills, J.H. Radial non-uniformity of velocity and voidage in a bubble column. Trans. Inst. Chem. Eng. 1974, 52, 1–9.
Fluids 2023, 8, 91 41 of 41
101. McClure, D.D.; Wang, C.; Kavanagh, J.M.; Fletcher, D.F.; Barton, G.W. Experimental investigation into the impact of sparger
design on bubble columns at high superficial velocities. Chem. Eng. Res. Des. 2016, 106, 205–213. [CrossRef]
102. Guan, X.; Yang, N. Bubble properties measurement in bubble columns: From homogeneous to heterogeneous regime. Chem. Eng.
Res. Des. 2017, 127, 103–112. [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.