You are on page 1of 13

Transition between regimes of a vertical channel bubbly upflow due to bubble

deformability
Sadegh Dabiri, Jiacai Lu, and Gretar Tryggvason

Citation: Physics of Fluids 25, 102110 (2013); doi: 10.1063/1.4824006


View online: https://doi.org/10.1063/1.4824006
View Table of Contents: http://aip.scitation.org/toc/phf/25/10
Published by the American Institute of Physics

Articles you may be interested in


Effect of bubble deformability in turbulent bubbly upflow in a vertical channel
Physics of Fluids 20, 040701 (2008); 10.1063/1.2911034

The effect of bubbles on the wall drag in a turbulent channel flow


Physics of Fluids 17, 095102 (2005); 10.1063/1.2033547

Multiscale considerations in direct numerical simulations of multiphase flows


Physics of Fluids 25, 031302 (2013); 10.1063/1.4793543

Investigation and modeling of bubble-bubble interaction effect in homogeneous bubbly flows


Physics of Fluids 22, 063302 (2010); 10.1063/1.3432503

Numerical study of turbulent bubbly downflows in a vertical channel


Physics of Fluids 18, 103302 (2006); 10.1063/1.2353399

Using statistical learning to close two-fluid multiphase flow equations for a simple bubbly system
Physics of Fluids 27, 092101 (2015); 10.1063/1.4930004
PHYSICS OF FLUIDS 25, 102110 (2013)

Transition between regimes of a vertical channel bubbly


upflow due to bubble deformability
Sadegh Dabiri,1,a) Jiacai Lu,2 and Gretar Tryggvason1
1
University of Notre Dame, Notre Dame, Indiana 46556, USA
2
Worcester Polytechnic Institute, Worcester, Massachusetts 01609, USA
(Received 19 July 2013; accepted 14 September 2013; published online 22 October 2013)

The effect of the deformability of viscous bubbles on the flow rate of bubbly upflow in
a vertical channel is examined using direct numerical simulations. A sharp transition
between two different flow regimes has been observed. At large bubble deformability,
characterized by large Eötvös number (Eo), the flow rate is close to the single phase
flow rate, with adjusted pressure gradient, and the bubbles are almost uniformly
distributed in the middle of the channel. On the other hand, at low Eo the bubbles
are concentrated near channel walls and flow rates are much smaller than the single
phase flow. The transition from high flow rate to low flow rate occurs rather abruptly.
It is found that the transition occurs when the less deformable bubbles enter the
viscous sublayer due to the lateral lift force on the bubbles. This leads to an increase
in the viscous dissipation near the wall which leads to a decrease in the flow rate.
C 2013 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4824006]

I. INTRODUCTION
Multiphase flow of gas-liquid mixtures in vertical channels is ubiquitous in many industrial
systems, such as in the power, chemical, food, and pharmaceutical industries.1, 2 As a result, these
types of flows have been the subject of research for a long time.3–12 The best-known early study
of bubbly flows is by Serizawa et al.13, 14 who examined the void fraction distribution and the
velocity profile in turbulent air-water bubbly flows. While there are some differences between the
results obtained by different investigators, possibly due to different bubble sizes, all show that
for nearly spherical bubbles the void fraction distribution and the velocity profile in the core of
the channel are relatively uniform and that a void fraction peak is generally found near the wall
for upflow but not for downflow. Deformable bubbles, on the other hand, migrate away from the
walls in upflow. A number of authors have developed two-fluid models of bubbly flows in vertical
channels to model this phenomenon.15–19 Analysis of a two fluid model of laminar bubbly flow
was done first by Antal et al.20 and studied later by Azpitarte and Buscaglia.21 One of the earliest
models for a turbulent case was developed by Drew and Lahey22 who solved the resulting equations
using asymptotic analysis. The models reproduced the general trends seen experimentally, including
uniform velocity and void fraction in the core, bubble rich wall-layers, and bubble depleted wall-
layers for upflow and downflow, respectively. Subsequent investigators have solved the resulting
equations numerically.23–26 The model generally reproduces the experimental results. The two-fluid
model has been extended to handle nonuniform bubble size distributions.27, 28
During the last decade, Direct Numerical Simulation (DNS) of multiphase flows has emerged
as one of the most promising approaches to understand the dynamics of such flows.29–31 For clean
surfaces, the dynamics is believed to be completely described by the Navier-Stokes equations and a
fully verified DNS therefore constitutes an essentially exact solution of a fully verified mathematical
model. Thus, such simulations, just as for DNS of single phase turbulent flows, provide a database

a) Electronic mail: sdabiri@nd.edu

1070-6631/2013/25(10)/102110/12/$30.00 25, 102110-1 


C 2013 AIP Publishing LLC
102110-2 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

that can be used to help develop closure models needed for coarse-grained simulations of the average
properties of industrial scale flows and a benchmark against which such models can be validated.
Here, we are interested at the transition between the two regimes. We consider the buoyant
motion of many bubbles in initially turbulent upflows in vertical channels. Experimental investiga-
tions have shown that the void fraction distribution in such flows depends strongly on the size of the
bubbles: small bubbles accumulate on the walls whereas larger bubbles stay away from the walls.
Such studies include the work by Liu32 who found that a transition from wall peak to core peak at a
bubble diameter of about 5-6 mm. The distribution is affected by the lift force on buoyant bubbles
in shear flow.33, 34 For a recent review of experimental work, see Guet and Ooms.35
Lu et al.29 examined the motion of nearly spherical bubbles in laminar flows. As expected, the
results showed that in upflow a fraction of the bubbles accumulated at the walls but the rest stayed
in the middle of the channel. For nearly spherical bubbles moving upward in a vertical shear flow,
the lift on the bubble drives the bubble laterally in the direction where the fluid velocity in a frame
moving with the bubble is larger.36 The lateral migration of the bubbles changes the local mixture
density and in upflow the mixture becomes heavier in the middle of the channel, as the bubbles
accumulate at the walls. If the initial void fraction is high enough, the weight of the mixture in the
middle eventually balances the pressure gradient imposed to drive the mixture upward. Once that
happens, the shear becomes zero and further migration of bubbles to the wall ceases. In downflow,
the opposite happens. Migration of bubbles from the wall to the middle decreases the density of the
mixture until the mixture buoyancy balances the pressure that drives the flow downward. Once the
mixture is in hydrostatic equilibrium, the shear is zero and the bubbles stop migrating, since there is
no lift. Thus, the void fraction in the middle of the channel is easily predicted and is equal to what
is needed for hydrostatic equilibrium. For upflow, the excess bubbles go to the wall, allowing us to
predict the void fraction of the wall-layer.
Lu and Tryggvason37 showed that similar considerations hold for turbulent upflow and that
nearly spherical bubbles in turbulent upflow behave in a way similar to what was seen for bubbles in
a laminar upflow. Deformable bubbles behave in a very different way. Since the lift on deformable
bubbles is either nearly zero, or slightly negative (compared to the lift on nearly spherical bubbles),
such bubbles stay away from the walls. This leads to a bubble free layer near the wall. Once the
driving pressure gradient has been adjusted to account for the change in mixture density due to the
addition of bubbles, relatively minor influence on the flow rate has been observed. In the present
paper, we examine the transition between two regimes of bubble distribution and flow rates.

II. PROBLEM SETUP AND NUMERICAL METHOD


To isolate the effect of viscous bubble deformability, we have carried out several simulations of
many buoyant viscous bubbles rising in an upflow in a vertical channel, where the surface tension is
the only parameter that is varied between the different simulations. The viscous bubbles are taken
to have the same viscosity as the liquid to represent gas-liquid flows at certain conditions, e.g.,
high-pressure high-temperature liquid/vapor water, where the viscosities are similar. We use the
same setup as in Lu et al.,37 which allows us to use the results obtained there as part of our data set.
Navier Stokes equations are solved on a Cartesian staggered grid for both liquid and gas phase as a
single fluid with variable properties

∂u
ρ + ρ∇ · (uu) = −∇ p + ∇ · τ + ρg + σ κnδ(x − x f )dA f , (1)
∂t f

∇ · u = 0, (2)
where ρ, u, p, τ are the local density, velocity, pressure, and the viscous stress tensor, respectively.
The last term on the right hand side represents the surface tension as a concentrated force on the
interface. Here, σ , κ, n, and δ are the surface tension coefficient, twice the mean curvature of the
interface, unit vector normal to the interface, and the Dirac delta function, respectively.
The computations are done using a front-tracking/finite-volume method originally introduced
by Unverdi and Tryggvason.38 Various improvements as well as verification studies can be found in
102110-3 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

Tryggvason et al.39 The method is based on the one-fluid formulation of the Navier-Stokes equations,
where a single set of equations is solved for both the bubbles and the ambient liquid. The one-fluid
formulation allows us to use a regular structured grid to solve for the flow and is the basis for a
number of methods currently used for flows with sharp interfaces. In our approach, the interface
between the different fluids is identified by connected marker points that are advected by the fluid
velocity, interpolated from the grid used to solve the fluid equations. Since we are interested in the
statistically steady state with a constant number of bubbles of a given size, here we do not allow
the bubbles to coalesce. In addition, the range of parameters is such that bubbles do not break
apart.
The computational domain is a rectangular channel bounded by two flat vertical walls (normal
to y-axis) and periodic in the spanwise (z) and the streamwise (x) directions. The flow is driven
upward (+x) by a specified pressure gradient. The channel has dimensions of π h × 2h × π h/2 in
x − y − z directions and is discretized by a 256 × 192 × 128 grid. Here, h is the half width of the
channel and is taken to be one in the present study. There are 21 bubbles with nominal diameter of
d/h = 0.3 (diameter of spherical bubble with the same volume). This gives a 3% volume fraction of
the bubbles in the channel. There are on average 25 grid points per diameter of the bubble. Esmaeeli
and Tryggvason40 showed that this resolution is enough for vorticity boundary layer of the bubbles
with Reynolds number of O(100) which is the case in this study. The density of the bubble is taken to
be one tenth of the liquid density and the viscosity of both fluids are taken to be equal.√The pressure
gradient is characterized by friction Reynolds number, Reτ = huτ /ν l , where u τ = τw /ρl is the
friction velocity, τw the average wall shear, and ν l is the liquid kinetic viscosity. Other dimensionless
parameters are the Archimedes number Ar = ρl2 gd 3 /μl2 = 2.43 × 104 for all cases, Eötvös number
(Eo=ρ l gd2 /σ ) which varies between 0.3 and 4.5, and a Reynolds number based on channel width
which is calculated a posteriori (Reh = 2hU/ν l ). The bubbles are initially inserted into a fully
developed single phase flow and the computations are carried out until the flow is approximately
at steady state, as monitored by the total flow rate and the wall shear. The computations were then
continued and data gathered to allow us to compute the various average quantities at steady state.

III. RESULTS
The results presented here consist of 14 simulations of bubbles of different deformability at
two different values of the pressure gradients that drive the flow upward, plus the results of two
simulations already presented in Ref. 37. These cases are summarized in Table I. It is important to
note that since the rise velocity of the bubbles varies by their location in the channel, e.g., slower
near the wall, an average rise velocity is used in Reb which is defined as the difference between the
average rise velocity of gas phase and liquid phase over the entire channel.
Figure 1 shows the volumetric flow rate in the channel, normalized by the flow rate of a single
phase flow. The pressure gradient of the single phase flow is adjusted to compensate for the higher
weight of the liquid compared to the two-phase mixture. In this manner, the average drag force for
single phase and bubbly flow are the same leading to the same values of friction Reynolds number,

TABLE I. Parameters of simulations.

Reτ Eo Reh Reb Reτ Eo Reh Reb

127 0.3 1440 121 90 0.9 880 167


127 0.9 1450 120 90 1.8 1010 155
127 1.8 1610 119 90 2.5 1420 143
127 2.25 1890 116 90 3 1800 166
127 2.5 2440 111 90 4 2280 198
127 2.75 3060 119
127 3 3380 136
127 3.5 3530 154
127 3.75 3570 150
102110-4 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

1
Reτ=127
0.8 Reτ=90

Flow rate
0.6

0.4

0.2

0
0 1 2 3 4 5
Eo

FIG. 1. The flow rate versus Eo, normalized by the flow rate in a channel with no bubbles, showing the asymptotic states at
high and low Eo and the transition around Eo = 2.5.

Reτ . The flow rate levels off at both ends of the Eo range with a sharp transition from low to high
flow rate at around Eo = 2.5. For Eo greater than 3.5, there are no bubbles sliding along the walls of
the channel and the presence of the bubbles has minor impact on the flow rate. Similarly, for lower
Eo, the flow rate levels off, since once the bubbles are essentially spherical, lowering Eo further does
not change their shape. The transition between the low flow rate and high flow rate occurs around
Eo = 2.5. Later, it will be shown that the bubbles start to separate from the wall at this Eötvös
number. The normalized flow rates of the cases with Reτ = 90 follow a trend similar to the case of
Reτ = 127 and show the same transition from low to high flow rates around Eo = 2.5. Transition
to low flow rates as the bubbles move toward the wall can be explained by the increase of shear
force on the wall as observed by Ref. 12. In contrast to the drag reduction due to micro-bubbles in
horizontal turbulent boundary layer, buoyant bubbles in vertical flow lead to increase in the drag and
slow down of the flow.
In Figure 2, we show the bubbles and the vorticity for deformed bubbles (left) and for nearly
spherical ones (right) after the flow has reached an approximate steady state. The bubbles are shown
alone in the top row and the bubbles along with the vortical structures in the bottom row. Here,
we identify the vortical structures using the lambda-2 method,41 plotting iso-surfaces of the second
eigenvalue of a tensor consisting of the sum of the squares of the symmetric and antisymmetric parts
of the velocity gradient tensor. For an overview of methods to visualize the vorticity, see Jiang et al.42
We have selected Reτ = 127, Eo = 3.0 for (a) and (c) and Eo = 0.9 for (b) and (d), representing two
very different flow states. The deformable bubbles on the left are distributed throughout the center
region of the channel and no bubble is close to the wall.
A significant fraction of the nearly spherical bubbles on the right are, however, hugging the wall.
The vorticity field is also different for these cases. For the deformable bubbles, we see longitudinal
vortices near the walls as well as vortical wakes behind the bubbles. The other case with the nearly
spherical bubbles has, however, essentially no vortical structures in the middle of the channel and
only weak vorticity behind the bubbles near the wall.
To get an impression of the movement of the bubbles, we plot their paths, projected on a plane
parallel to the flow in Figure 3, for the same two cases shown in Figure 2. To make it possible to
identify the individual paths, we have plotted them for a time interval of 100 computational units,
after the flow has reached an approximate steady state. The deformable bubbles in the left frame
move aggressively both across the channel and parallel to the walls. They do, however, never go
close to the wall, resulting in a bubble free wall region as seen in the figures above. Most of the
nearly spherical bubbles move along the wall and although there is some horizontal motion along
the walls, there is essentially no exchange of bubbles between the wall layer and the center region,
where a few bubbles move in both horizontal directions, although not quite as vigorously as in
the deformable case. One may also notice that nearly spherical bubbles travel much less than the
deformable bubbles during the same period of time.
102110-5 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

FIG. 2. The bubbles and the vortical structures in the liquid, as visualized by the lambda-2 method. In (a) and (b), the bubbles
only are plotted and in (c) and (d) we include the vorticity as well. Eo = 3.0 in (a) and (c) and Eo = 0.9 in (b) and (d).

The bubble dynamics is quantified in Figure 4, where we plot different quantities versus Eo. In
Figure 4(a), we show average velocity fluctuations of the bubbles in the vertical direction and the
wall-normal direction. The fluctuations show a strong dependency on Eo, with significant increase
as the bubbles become more deformable and move into the interior of the channel. The increase of
the bubble deformability affects the bubble velocity fluctuations in two different ways. As shown
by Bunner and Tryggvason,43 the velocity fluctuations of the bubbles rising in a homogeneous flow
increase with the void fraction. Since the motion of the bubbles in the center of the channel is close to
homogeneous flow, the migration of the bubbles to the center of the channel increases the void fraction
in the center and increases the velocity fluctuations as well. In addition, the velocity fluctuations
are smaller near the wall due to restriction of motion that is imposed by the no slip/impermeability
condition on the walls. Therefore, motion of the bubbles away from the wall increases their velocity
fluctuations. Figure 4(b) shows the bubble deformation averaged over all the√bubbles in the channel
and over time. The deformation of each bubble is measured using χ = Imax /Imin , the square
root of the ratio of the largest and smallest eigenvalues of the second moment of inertia tensor,
given by

1
Ii j = (xi − xio )(x j − x jo )d V. (3)
vol vol
Here, vol is the volume of the bubble and xio and xjo are the coordinates of the bubble’s centroid
in the i and j directions. It is shown that for small deviation from spherical shape, χ matches the
aspect ratio of the bubble.44 The volume integrals are calculated as surface integrals over the bubble
interface using the divergence theorem. As expected, the deformation increases monotonically with
Eo. The deformation of bubbles increases almost linearly with Eo up to Eo = 3.75.
102110-6 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

(a) (b)
90 90

80 80

70 70

60 60

50 50
x

x
40 40

30 30

20 20

10 10

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
y y
FIG. 3. The paths of the bubbles projected on a plane normal to the spanwise direction, shown for a time interval of 100
computational units, after the flow has reached an approximate steady state. Eo = 3 in frame (a) and Eo = 0.9 in frame (b).
Initial and final locations of bubbles are marked by circles.

Two correlations for the deformation of a single bubble rising in a liquid are used for comparison.
Since the bubbles deformation is different near the wall because of shear, the average deformation of
the bubbles in the middle of the channel is calculated and plotted versus the average Weber number
of the bubbles in the middle of the channel. The results are presented in Figure 5 and compared
with the correlations, one by Loth45 which for Re>100 is given as 1/χ = 1 − 0.75 tanh(0.165W e)
and the other by Legendre46 as 1/χ = 1 − 64 9
W e for very small Morton number (Mo = gμl4 /ρl σ 3 ).
Note that these correlations have been obtained for the deformation of a single bubble rising in a
quiescent fluid. The dispersion in the calculated deformation in Figure 5 is due to the wide range of
Morton number in the calculations (4.57 × 10−11 < Mo < 1.54 × 10−7 ).
The dynamics of the bubbles changes greatly as Eo changes, as seen in Figures 2–4. These
differences affect the liquid flow in a profound way. In Figure 6, we show the average motion of

3 2
Bubbles’ velocity fluctuation

Streamwise (a) 1.8


(b)
Wall-normal
2
Deformation

1.6

1.4
1
1.2

0 1
0 1 2 3 4 5 0 1 2 3 4 5
Eo Eo

FIG. 4. Bubble dynamics vs. Eötvös number for the cases with Reτ = 127: (a) fluctuation in the stream-wise and wall-normal
velocity of bubbles normalized by uτ ; (b) the average bubble deformation measured as described in the text.
102110-7 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

1.8

1.6

χ
1.4

1.2

1
0 1 2 3 4 5
We

FIG. 5. Aspect ratio of bubbles in the center of the channel compared with correlations by Loth45 (solid line) and Legendre46
(dashed line).

the mixture, computed over the same time period used to calculate the average bubble dynamics in
Figure 4. The various quantities are averaged over planes parallel to the walls and plotted versus
the wall-normal coordinate. These quantities are averaged over left and right half of domain as
well. In Figure 6(a), the average velocity is plotted for three values of Eo and in Fig. 5(b) we plot
the average void fraction. Both the average velocity and the void fraction change significantly as
the Eo is increased. For low Eo, we see a strong wall peak in the void fraction and the velocity
is greatly reduced. The void fraction is well predicted by the model introduced in Ref. 29 where
the void fraction in the channel core is assumed to be in hydrostatic equilibrium and the excess
bubbles are pushed to the wall. As Eo increases, the wall-peaked void fraction profile changes to a
core-peaked profile and the velocity remains essentially unchanged when compared with a single

20 0.15
(a) single phase (b) Eo = 3.75
2.5
15 0.9
Void fraction

0.1
+

10
u

0.05
5

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
y y
1 2
(c) (d)
0.5 1.5
Reynolds stress

2
ωx

0 1

-0.5 0.5

-1 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
y y

FIG. 6. The average flow at approximately steady state for three values of Eo at Reτ = 127. The various quantities are
averaged over planes parallel to the walls and plotted versus the wall-normal coordinate. (a) The average velocity normalized
by uτ ; (b) the void fraction; (c) the cross velocity component of the Reynolds stress normalized by wall friction; (d) the
streamwise vorticity squared.
102110-8 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

3 1.5
Single phase
Two phase, Eo=4.5
Two phase, Eo=3
Two phase, Eo=0.9
2 1
u′+

v′+
1 0.5

(a) (b)
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
y y

FIG. 7. Velocity fluctuation profiles across the channel at Reτ = 127 normalized by uτ : (a) streamwise velocity;
(b) wall-normal velocity.

phase flow (with the appropriate adjustment of the pressure gradient to account for the reduction in
mixture density as the bubbles are added). Additional differences in the flow are visible in frames
(c) where the cross velocity component of the Reynolds stresses is shown and in frame (d) which
shows the streamwise vorticity. In the cases with deformable bubbles, Reynolds stress has a linear
profile except near the walls, similar to a single phase turbulent flow, whereas, in the case with nearly
spherical bubbles the Reynolds stress is almost zero except in a layer near the walls. Similarly, the
streamwise vorticity becomes confined to the near-wall region. These observations suggest that the
flow has been laminarized when bubbles are nondeformable.
Velocity fluctuations in both streamwise and wall-normal direction are shown in Figure 7. In
the case of Eo = 0.9, the velocity fluctuations are generally reduced across the channel. There is
the exception of the near wall region where the stirring motion of the liquid cased by rise of the
bubbles near the wall increases the velocity fluctuations especially in the wall-normal direction. Both
velocity fluctuations are increased by the motion of deformable bubbles at high Eo especially in the
center of the channel where the streamwise fluctuations are nearly doubled. This suggests that the
increase of the fluctuation in the velocity of the bubbles (Figure 4(a)) is a result of bubble agitation.
Figure 8 shows
 the average distance  of the bubbles from the closest wall which is defined
as Yb  = h1N i=1,N min(ybi , 2h − ybi ) with  representing the time average. As opposed to the
abrupt change in the flow rate near the critical Eo, the average wall distance increases gradually as
the bubbles become more deformable. One may also observe in Figure 6(b) that the peak of the void
fraction profile gradually moves away from the wall. Predictions of a physical model for spherical
bubbles in the limit of small Eo are also shown in dashed lines. Averaging the conservative form of
Eq. (1) over the x − z plane and over time and assuming a statistically steady state flow, one can

0.8
Reτ=127
Reτ=90
0.6 Model
Wall distance

0.4

0.2

0
0 1 2 3 4 5
Eo

FIG. 8. The average distance of bubbles from the nearest channel wall. The dashed lines represent the prediction of the
physical model for spherical bubbles (small Eo) for each Reτ .
102110-9 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

write the following form of the average momentum equation in x-direction:

dρu  v  ∂ p dτx y
=− + + ρg + F f . (4)
dy ∂x dy

Here, the symbol () represents the average over x-z planes and over time. The term F f represents
the average of the surface tension forces resulting from the last term in Eq. (1). If we consider the
flow with nearly spherical bubbles, as it can be seen in Figure 6, the Reynolds and viscous stresses
are nearly zero in the center of the channel. In addition, since the bubbles in the center of the
channel are rising in a uniform flow with a uniform void fraction distribution, i.e., a homogeneous
bubbly flow in the center, the average surface tension forces should be zero there. Therefore, we
end up with only two nonzero terms for momentum balance in the center of the channel. We
define β = −∂ p/∂ x + ρave g as the total pressure gradient, consisting of the sum of the imposed
pressure gradient and the hydrostatic pressure gradient due to the weight of the mixture with ρave
being the average density over the entire domain. We also define ρ  = ρ − ρave . This leads to

β + ρ  (y)g = 0 (5)

for the center of the channel. Rewriting this in terms of void fraction leads to a simple expression
for the average void fraction in the center of the channel
β
c = ave + , (6)
g ρ
which gives the values of void fractions at the center of the channel to be c = 0.02 for Reτ = 90
and c = 0.01 for Reτ = 127. We define the wall void fraction as the void fraction in a layer of
d
thickness d near the wall, i.e., w = d1 0 (y)dy. The average wall distance of the bubbles in the
wall layer is d/2 and the average wall distance of the bubbles in the center is (h + d)/2, because the
void fraction in the center of the channel is uniform if the bubbles are nearly spherical.37 From these,
we can calculate the average wall distance of the bubbles in the channel in the limit of very small
Eötvös number. Then, the void fraction near the wall can be calculated by conserving total volume
of bubbles. Now, by assuming a uniform distribution of bubbles in the center of the channel37 and
an average wall distance of a bubble radius for the bubbles in the wall layer, one can predict the
average wall distance of the nearly spherical bubbles. The numerical values are 0.3833 for Reτ = 90
and 0.2666 for Reτ = 127. This prediction matches well with the results of the simulation for small
Eo as shown in Figure 8. As the deformability of bubbles increases, bubbles start to move away
from the walls due to change in the direction of the lift force on them. Therefore, the average wall
distance starts to increase. This increase in the average wall distance coincides with the transition to
high flow rate (Figure 1).

20
5
+ +5.9
ny
15 .44l
u =2
+
+

10
u

u+=y+

5 Single phase flow


Bubbly flow (Eo=3)
Law of the wall
0 0 1 2
10 10 10
y+

FIG. 9. The details of the velocity profile near the wall for Reτ = 127 and Eo = 3 compared to the single phase flow at the
same friction Reynolds number and law of the wall.
102110-10 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

1.5 0.06

Void fraction in viscous sublayer


Single phase (a) (b)
Viscous dissipation Eo=3 0.05
Eo=2.5
2
1 Eo=1.8 0.04

1 0.03

0.5 0 0.02
0 0.05 0.1
0.01

0 0
0 0.2 0.4 0.6 0.8 1 0 1 2 3 4 5
y Eo

FIG. 10. (a) The distribution of the viscous dissipation across the channel for Reτ = 127 normalized by τw2 /μ and the details
near the wall; (b) The void fraction inside the viscous sublayer (y+ ≤ 10) for Reτ = 127.

The sharp change in the flow rate in the channel can be explained by considering the viscous
dissipation in the channel and near the walls. First, the velocity profiles for single phase flow and
one of the cases of two phase flow with Reτ = 127 and Eo = 3 are plotted along with the law
of the wall in Figure 9 showing the viscous sublayer. The average viscous dissipation across the
channel and the detail near the wall are shown in Figure 10(a). The viscous dissipation for nearly
spherical bubbles has a very sharp peak near the wall. This is due to the generation of high levels of
strain rates, e.g., streamwise vorticity (refer to Figure 6(d)), between the bubbles and the wall as the
bubbles slide on the wall. As the bubbles move away from the wall, this peak shrinks and becomes
closer to the single-phase flow. Therefore, the presence of bubbles in the viscous sublayer would
increase the viscous dissipation and leads to lower flow rate. Presence of the bubbles in the viscous
sublayer is quantified by plotting the average void fraction in the viscous sublayer (0 ≤ y+ ≤ 10)
in Figure 10(b). The void fraction becomes very small (less than 0.4%) at Eo = 2.75 and smaller
at higher Eo. Tran-Cong et al.47 have made a similar observation that air bubbles of size 4 mm and
above do not stay in the viscous sublayer in a turbulent boundary layer on a vertical wall. This air
bubble size in water corresponds to Eo = 2.18 which is close to what is found here. Based on these
observations, as the bubbles become more spherical they enter the viscous sublayer near the wall
and the peak of viscous dissipation rate increases near the wall, resulting in a decrease in the flow
rate in the channel.
It is worth mentioning that the effect of gas-to-liquid viscosity ratio is not studied here. The
viscous bubble has a different boundary condition on the interface compared to a bubble with
a viscosity much smaller than the liquid. This may affect flow characteristics such as vorticity
production on the interface, the wake structure, the bubble path instability, and the viscous dissipation
in the flow.

IV. CONCLUSION
For bubbly upflow in a vertical channel, it is well known that spherical or nearly spherical
bubbles hug the walls but deformable bubbles remain in the center of the channel. Here, we show
that the transition from low flow rate channel with wall-hugging nearly spherical bubbles to high
flow rate channels with deformable bubbles is sharp. It is shown that above the critical Eo (near 2.5
for the parameters in present study), the deformability of the bubbles has minor impact on the flow
rate (at very high Eo, we would expect the bubbles to break into smaller bubbles so the assumption
of a high Eo bubbles would not hold). It is shown that for Eo below the critical value the flow rate
is approximately constant and much lower than for the deformable bubbles. In the transition region,
between the low flow rate at low Eo and the high flow rate at high Eo, the flow rate depends strongly
on the bubble deformability. Examination of the viscous dissipation profile across the channel shows
a sudden increase of the peak value of dissipation at the wall when the bubbles enter the viscous
sublayer which results in a decrease in the flow rate.
102110-11 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

ACKNOWLEDGMENTS
This study was supported in part by the Consortium for Advanced Simulation of Light Water
Reactors (CASL) and NSF Grant No. CBET-1132410.
1 S. Furusaki, L.-S. Fan, and J. Garside, The Expanding World of Chemical Engineering, 2nd ed. (Taylor and Francis, New
York, 2001).
2 W.-D. Deckwer, Bubble Column Reactors (John Wiley and Sons Ltd, Chichester, 1992).
3 S. K. Wang, S. J. Lee, O. C. Jones, Jr., and R. T. Lahey, Jr., “3-D turbulence structure and phase distribution in bubbly

two-phase flows,” Int. J. Multiphase Flow 13, 327–343 (1987).


4 T. J. Liu and S. G. Bankoff, “Structure of air-water bubbly flow in a vertical pipe–I. Liquid mean velocity and turbulence

measurements,” Int. J. Heat Mass Transfer 36, 1049–1060 (1993).


5 V. Nakoryakov, O. Kashinsky, V. Randin, and L. Timkin, “Gas-liquid bubbly flow in vertical pipes,” J. Fluids Eng. 118,

377–382 (1996).
6 T. J. Liu, “Investigation of the wall shear stress in vertical bubbly flow under different bubble size conditions,” Int. J.

Multiphase Flow 23, 1085–1109 (1997).


7 O. N. Kashinsky and V. V. Randin, “Downward bubbly gas–liquid flow in a vertical pipe,” Int. J. Multiphase Flow 25,

109–138 (1999).
8 S. So, H. Morikita, S. Takagi, and Y. Matsumoto, “Laser doppler velocimetry measurement of turbulent bubbly channel

flow,” Exp. Fluids 33, 135–142 (2002).


9 S. Guet, G. Ooms, R. Oliemans, and R. F. Mudde, “Bubble size effect on low liquid input drift-flux parameters,” Chem.

Eng. Sci. 59, 3315–3329 (2004).


10 A. Matos, E. de Rosa, and F. Franca, “The phase distribution of upward co-current bubbly flows in a vertical square

channel” J. Braz. Soc. Mech. Sci. Eng. 26, 308–316 (2004).


11 S. Mendez-Diaz, R. Zenit, S. Chiva, J. Muñoz-Cobo, and S. Martinez-Martinez, “A criterion for the transition from wall

to core peak gas volume fraction distributions in bubbly flows,” Int. J. Multiphase Flow 43, 56–61 (2012).
12 M. N. Descamps, R. V. A. Oliemans, G. Ooms, and R. F. Mudde, “Air-water flow in a vertical pipe: Experimental study of

air bubbles in the vicinity of the wall,” Exp. Fluids 45, 357–370 (2008).
13 A. Serizawa, I. Kataoka, and I. Michiyoshi, “Turbulence structure of air-water bubbly flow–II. Local properties,” Int. J.

Multiphase Flow 2, 235–246 (1975).


14 A. Serizawa, I. Kataoka, and I. Michiyoshi, “Turbulence structure of air-water bubbly flow–III. Transport properties,” Int.

J. Multiphase Flow 2, 247–259 (1975).


15 M. Ishii, Thermo-Fluid Dynamic Theory of Two-Phase Flows (Eyrolles, Paris, 1975).
16 I. Kataoka and A. Serizawa, “Basic equations of turbulence in gas-liquid two-phase flow,” Int. J. Multiphase Flow 15,

843–855 (1989), http://www.sciencedirect.com/science/article/pii/0301932289900451.


17 D. Z. Zhang and A. Prosperetti, “Ensemble phase-averaged equations for bubbly flows,” Phys. Fluids 6, 2956–2970

(1994).
18 D. A. Drew and S. L. Passman, Theory of Multicomponent Fluids (Springer-Verlag, New York, 1999).
19 A. Prosperetti and G. Tryggvason, Computational Methods for Multiphase Flow (Cambridge University Press, New York,

2007).
20 S. P. Antal, R. T. Lahey, and J. E. Flaherty, “Analysis of phase distribution in fully developed laminar bubbly two-phase

flows,” Int. J. Multiphase Flow 17, 635–652 (1991).


21 O. E. Azpitarte and G. C. Buscaglia, “Analytical and numerical evaluation of two-fluid model solutions for laminar fully

developed bubbly two-phase flows,” Chem. Eng. Sci. 58, 3765–3776 (2003).
22 D. A. Drew and R. T. Lahey, Jr. “Analytical modeling of multiphase flow,” Particulate Two-Phase Flow, edited by M. C.

Roco (Butterworth-Heinemann, 1993), pp. 509–566.


23 M. Lopez De Bertodano, R. T. Lahey, Jr., and O. C. Jones, “Development of a k– model for bubbly two-phase flow,” J.

Fluids Eng. 116(1), 128–134 (1994).


24 M. Lopez De Bertodano, R. T. Lahey, Jr., and O. C. Jones, “Phase distribution in bubbly two-phase flow in vertical ducts”

Int. J. Multiphase Flow 20, 805–818 (1994).


25 T. C. Kuo, C. Pan, and C. C. Chieng, “Eulerian–Lagrangian computations on phase distribution of two-phase bubbly flows,”

Int. J. Numer. Methods Fluids 24, 579–593 (1997).


26 S. Guet, G. Ooms, and R. Oliemans, “Simplified two-fluid model for gas-lift efficiency predictions,” AIChE J. 51, 1885–

1896 (2005).
27 I. Celik and A. Gel, “A new approach in modeling phase distribution in fully developed bubbly pipe flow,” Flow, Turbul.

Combust. 68, 289–311 (2002).


28 M. Politano, P. Carrica, and J. Converti, “A model for turbulent polydisperse two-phase flow in vertical channel,” Int. J.

Multiphase Flow 29, 1153–1182 (2003).


29 J. Lu, S. Biswas, and G. Tryggvason, “A DNS study of laminar bubbly flows in a vertical channel,” Int. J. Multiphase Flow

32, 643–660 (2006).


30 S. Dabiri, W. A. Sirignano, and D. D. Joseph, “Interaction between a cavitation bubble and shear flow,” J. Fluid Mech.

651, 93–116 (2010).


31 M. Gorokhovski and M. Herrmann, “Modeling primary atomization,” Annu. Rev. Fluid Mech. 40, 343–366 (2008).
32 T. J. Liu, “Bubble size and entrance length effects on void development in a vertical channel,” Int. J. Multiphase Flow 19,

99–113 (1993).
33 A. Tomiyama, H. Tamai, I. Zun, and S. Hosokawa, “Transverse migration of single bubbles in simple shear flows,” Chem.

Eng. Sci. 57, 1849–1858 (2002).


102110-12 Dabiri, Lu, and Tryggvason Phys. Fluids 25, 102110 (2013)

34 A. Kariyasaki, “Behavior of a single gas bubble in a liquid flow with a linear velocity profile,” in Proceedings of the 1987
ASME-JSME Thermal Engineering Joint Conference, edited by P. Marto and I. Tanasawa (ASME Technical Publishing
Office, New York, 1987), Vol. 5, p. 261.
35 S. Guet and G. Ooms, “Fluid mechanical aspects of the gas-lift technique,” Annu. Rev. Fluid Mech. 38, 225–249 (2006).
36 E. Ervin and G. Tryggvason, “The rise of bubbles in a vertical shear flow,” J. Fluids Eng. 119, 443 (1997).
37 L. Lu and G. Tryggvason, “Effect of bubble deformability in turbulent bubbly upflow in a vertical channel,” Phys. Fluids

20, 040701 (2008).


38 S. O. Unverdi and G. Tryggvason, “A front tracking method for viscous incompressible flows,” J. Comput. Phys. 100,

25–37 (1992).
39 G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han, S. Nas, and Y. J. Jan, “A front tracking

method for the computations of multiphase flow,” J. Comput. Phys. 169, 708–759 (2001).
40 A. Esmaeeli and G. Tryggvason, “A direct numerical simulation study of the buoyant rise of bubbles at O(100) Reynolds

number,” Phys. Fluids 17, 093303 (2005).


41 J. Jeong and F. Hussain, “On the identification of a vortex,” J. Fluid Mech. 285, 69–94 (1995).
42 M. Jiang, R. Machiraju, and D. Thompson, “Detection and visualization of vortices,” The Visualization Handbook

(Academic Press, Orlando, 2005), pp. 295–309.


43 B. Bunner and G. Tryggvason, “Dynamics of homogeneous bubbly flows. Part 2. Velocity fluctuations,” J. Fluid Mech.

466, 53–84 (2002).


44 B. Bunner and G. Tryggvason, “Effect of bubble deformation on the properties of bubbly flows,” J. Fluid Mech. 495,

77–118 (2003).
45 E. Loth, “Quasi-steady shape and drag of deformable bubbles and drops,” Int. J. Multiphase Flow 34, 523–546 (2008).
46 D. Legendre, R. Zenit, and J. R. Velez-Cordero, “On the deformation of gas bubbles in liquids,” Phys. Fluids 24, 043303

(2012).
47 S. Tran-Cong, J. L. Marie, and R. J. Perkins, “Bubble migration in a turbulent boundary layer,” Int. J. Multiphase Flow 34,

786–807 (2008).

You might also like