Vous êtes sur la page 1sur 21

Bubble motion in a converging-diverging channel

Harsha Konda, Manoj Kumar Tripathi and Kirti Chandra Sahu∗


Department of Chemical Engineering,
Indian Institute of Technology Hyderabad,
Yeddumailaram 502 205, Telangana, India
(Dated: May 26, 2015)

Abstract
The dynamics of bubble motion through a two-dimensional converging-diverging channel is nu-
merically studied using a finite volume flow approach. A parametric study is conducted inves-
tigating the effects of Reynolds number, Eötvös number, and amplitude of converging-diverging
channel. We found that increasing Re and h increases the oscillation of the bubble, and pro-
motes the migration of bubble toward one of the channel wall. While travelling inside the channel
from trough to crest regions, the bubble undergoes oblate-prolate deformation periodically at early
times, which becomes chaotic at later times. These oscillations in the shape of the bubble and the
complex path travelled by the bubble can enhance mixing, which is desirable in many applications
involving small-scale flows.


ksahu@iith.ac.in

1
I. INTRODUCTION

The dynamics of a bubble moving through channels and pipes has been the subject
of numerous experimental, theoretical and numerical studies due to their relevance to a
number of applications, ranging from microscale to macroscale flows, such as mixing in
microfluidics, biological applications, chemical reactors, etc. [1–4]. In microscale flows,
surface-tension force is dominant over gravitational force, but in large-scale systems, gravity
plays a significant role. Thus, several researchers studied the motion of bubbles and drops
due to buoyancy in vertical channels and pipes [5, 6], also in open domains [7, 8]. At
smaller scales, spatially varying walls are frequently encountered, and play a vital role in the
resultant flow dynamics [9–11]. First, we discuss briefly the flow dynamics associated with
spatially varying geometries, which has received considerable attention in the recent past,
mainly, in the context of small-scale flows.
Many researchers (see e.g., [12–20]) have investigated flow through converging-diverging
channels by conducting numerical simulations, experiments and linear stability analysis.
Stone & Vanka [21] numerically studied the flow field in a wavy channel and observed en-
hanced mixing above a certain Reynolds number (Re) after which the separation bubbles
distort/divide to form multiple roll-up structures near the walls. Sahu [10] investigated the
instability behaviour of the flow through a converging-diverging pipe having unequal lengths
of the converging and diverging portions, and found that different flow characteristics can be
obtained by changing the direction of the flow. The main emphasis in most of these studies
has been on the enhancement of mixing, heat and mass transfer rates in small-scale devices
(which are associated with low flow rates) by introducing spatially varying geometries, which
in turn makes the flow unstable. A few studies [9, 22] ware also conducted on flow through
diverging pipe in order to answer the discrepancy of linear stability theory and experimen-
tal observations for Hagan-Poiseuille flow. Furthermore, flow through corrugated channels
increases residence time of the fluid (amount of time that a fluid particle spends inside the
channel or pipe) without having to increase the length of the device [23, 24]. Thus, the corru-
gations are helpful in designing compact and efficient heat exchangers for thermally sensitive
substances, such as those encountered in food processing and pharmaceutical industries [25].
A very few studies also experimentally investigated the motion of bubble inside corrugated
channels focusing on the average features of the flow, such as pressure drop and friction

2
factor, etc. Gardeck & Lebouche [26] experimentally measured wall shear stress due to
gas-liquid flow in a corrugated channel. Vlasogiannis et al. [27] studied the heat transfer
characteristics of a plate heat exchanger, and observed recirculating zones of liquid phase in
the near-wall regions, and gaseous phase in the centreline region of the pipe. For gas-liquid
flow in wavy channels having different phase-difference between the walls (i.e. Φ = 0, π/2
and π), Nilpueng & Wongwises [28] obtained correlations between pressure drop and friction
factor. They also found that for Φ = π the recirculation zone is bigger than that observed
for Φ = 0 or π/2. Slug flow was observed only for Φ = π (named as symmetrical channel in
the present study), which was not seen for other Φ values considered in their study.
In the present study, we investigate the motion of a bubble inside a converging-diverging
channel at Reynolds numbers, which are expected to yield laminar flows. This type of
design improves mixing and increases efficiency of the above-mentioned small-scale devices.
Although the migration of single bubble in horizontal channels and pipes has been well
studied [29–31], to the best of our knowledge, none of them have investigated bubble motion
in converging-diverging geometries numerically, which is the focus of the present study.
Another motivation of the present work is to investigate the Segré-Silberberg effect [32]
for bubble migration inside a converging-diverging channel. This phenomenon was first
documented by Segré & Silberberg [32], where a macroscopic rigid sphere moving inside a
pipe or between two parallel plates consisting of a fluid, migrated toward the wall. This
migration of particle may be attributed to the nonlinear effects, which arise inside a flow for
Re > 0. Although the original study was conducted for rigid sphere, this phenomenon is also
applicable to liquid-liquid or gas-liquid systems [33]. Since then the Segré-Silberberg effect
has been investigated by several researchers for various parameter regimes. Pan & Glowinski
[34] studied the motion of buoyant circular cylinders in plane-Poiseuille flow using direct-
numerical simulations. They found that with increasing Re the cylinders migrated away
from the centreline of the channel. Matas et al. [35] analytically derived the equilibrium
position of point particles in a pipe flow. Douglas-Hamilton et al. [33] investigated motion
of particles inside a solution in the context of sperm-concentration-analyzing-devices. They
were able to predict the correction factor to be applied to the measurements in the automated
semen-analysis systems to account for the Segré-Silberberg effect. In the present study, we
investigated the dynamics of bubble in a converging-diverging channel. Here, we expect that
the wavy motion of the bubble will accentuated due to the Segré-Silberberg effect, which in

3
turn could enhance mixing.
The rest of the paper is organized as follows. The details of the problem formulation are
given in section II, and numerical approach used in the present study is provided in section
III. In section IV, a parametric study is conducted, and the effects of various dimensionless
numbers on the flow dynamics are discussed. Concluding remarks are given in section V.

II. FORMULATION

We numerically investigate the dynamics of an air bubble (designated by fluid ‘B’) of


initial radius r, moving inside a converging-diverging two-dimensional channel, consisting of
another fluid (designated by fluid ‘A’) as shown in Fig. 1 using a Volume-of-Fluid (VoF)
approach. The positions of sinusoidal top and bottom walls are described by y = R +
hsin 2πx 2πx
 
λ
+ Φ , and y = −R + hsin λ
, respectively. Here, h, λ and Φ are the half-
amplitude, the wavelength of the converging-diverging portion of the channel, and phase
difference between the walls, respectively. Both the fluids are considered to be incompressible
and Newtonian. We use a Cartesian coordinate system (x, y) with its origin at the center of
the channel inlet to model the flow, wherein x and y represent the horizontal and vertical
directions, respectively. The gravitational force is assumed to be very small as compared
to the inertial force, thus the former one is neglected in the present study. This is a valid
approximation for flow inside a channel with R of the order of a millimetre. The viscosity
and density of fluids A and B are µA , ρA and µB , ρB , respectively.
The governing equations of the problem are:

∇ · u = 0, (1)
 
∂u
+ u · ∇u = −∇p + ∇ · µ(∇u + ∇uT ) + δσκn,
 
ρ (2)
∂t
∂c
+ u · ∇c = 0, (3)
∂t
where u(u, v) represents velocity field, wherein u and v are the components of velocity in the
horizontal and vertical directions, respectively; p denotes the pressure field; c is the volume
fraction of fluid A, whose values are 0 and 1 in the air and liquid phases, respectively; δ
is the Dirac delta function, whose value is one at the interface and zero in the rest of the
domain; κ = ∇ · n is the curvature, wherein n is the unit normal to the interface pointing

4
towards fluid A, and σ is the interfacial tension coefficient of the liquid-gas interface. The
surface tension force is included in Eq. (2) using continuum surface force formulation [36].
The density, ρ and viscosity, µ are calculated as volume averaged quantities as follows:

ρ = ρB (1 − c) + ρA c, (4)
µ = µB (1 − c) + µA c. (5)

The following scaling is employed in order to render the governing equations dimensionless:
re
(x, y) = r (e
x, ye) , t = u, ve), p = ρB V 2 pe, µ = µ
t, (u, v) = V (e eµ0 , ρ = ρeρB , (6)
V
where tildes designate dimensionless quantities, and V is the velocity of the fluid injected at
the channel inlet, given by V ≡ Q/2R, where Q is the volume flow rate at the inlet; the value
of Q is taken to be one in the present study. After dropping tildes from all non-dimensional
terms, the governing dimensionless equations are given by

∇ · u = 0, (7)
 
∂u 1 δ
∇ · µ(∇u + ∇uT ) +
 
ρ + u · ∇u = −∇p + n̂∇ · n̂, (8)
∂t Re Eo
∂c
+ u · ∇c = 0, (9)
∂t
where Re(≡ ρB V R/µB ) and Eo(≡ ρB gR2 /σ) denote the Reynolds number and Eötvös
number, respectively. The dimensionless viscosity and density are given by:

ρ = (1 − c) + ρr c, (10)
µ = (1 − c) + µr c, (11)

where ρr (≡ ρA /ρB ) and µr (≡ µA /µB ) are density ratio and viscosity ratio, respectively.

III. NUMERICAL METHOD AND VALIDATION

We use Gerris, an open-source finite-volume fluid flow solver [37] to simulate the dynamics
of bubble motion in a converging-diverging symmetrical channel (Φ = π). In the present
study, the density and viscosity ratios are fixed at ρr = 10−3 and µr = 10−2 , respectively,
which correspond to air-water systems. Numerically it is difficult to handle such high values
of ρr and µr , and is known to create spurious currents at the interface separating the fluids.

5
Gerris minimizes this problem by using a balanced-force continuum surface force formulation
[36] for the calculation of surface tension. In addition, due to the curvature at the walls very
large number of grids are required to resolve the boundary layer. The adaptive refinement
feature of the Gerris is very useful for this purpose. In the present study, the near-wall and
fluid regions are refined separately. For the near-wall regions, stationary uniform grid sizes
are used, whereas adaptive grid refinement based on vorticity magnitude and position of the
interface separating the fluids is implemented for the fluid region.
The present code have been validated extensively by simulating several interfacial flow
problems. The reader is also referred to Tripathi et al. [5, 7, 8], wherein extensive validations
of the present code have been presented. In addition, we have ensured that convergence is
indeed achieved upon mesh refinement, which is shown in Fig. 2 for Eo = 10 and 25. The
rest of the parameter values are ρr = 10−3 , µr = 10−2 and Re = 100. Thus, in order to
generate the rest of the results presented in this paper, the grid − 1 for which the smallest
grid sizes at the near-wall and the fluid regions are 0.031 and 0.015, respectively, is used.
The influence of various dimensionless numbers, such as Reynolds number (Re), Eötvös
number (Eo), and half-amplitude of the converging-diverging channel (h) is discussed below.

IV. RESULTS AND DISCUSSION

We begin the presentation of our results by investigating the effects of Reynolds number
for the symmetrical channel (Φ = π) by plotting the location and velocity of center-of-
gravity, and aspect ratio (Ar ≡ w/l) of the bubble in Figs. 3 and 4 for Eo = 1 and Eo = 5,
respectively. Here w and l represent the height and length of the bubble, and the rest of
the parameter values are ρr = 10−3 , µr = 10−2 . It can be seen in Fig. 3(a) that for Eo = 1,
the bubble migrates toward the bottom wall as it travels inside the channel for all values of
Reynolds number considered. Inspection of this plot also reveals that increasing Re takes
the bubble away from the centreline, which is consistent with the literature [34], however
the corrugations do not allow for an equilibrium vertical location, as expected by Segré and
Silberberg [32, 38, 39] for straight pipes.
It can be seen that the bubble moves along the centreline of the channel at the early
times and then undergoes an oscillatory motion as it migrates toward the bottom wall.
The amplitude of oscillation increases as the bubble translates in the axial direction. Close

6
inspection of Fig. 3(a) divulges that the starting time of departure of the bubble from the
centreline reduces with increasing Re. The vertical velocity component of the center-of-
gravity of the bubble shown in Figs. 3(b) also demonstrates this oscillatory behaviour. In
Fig. 3(c), the instantaneous aspect ratio of the bubble is plotted versus time for different
values of Reynolds number. It can be seen that the bubble undergoes a continuous oblate-
prolate type deformation, which is periodic at early times (t < 70) for this set of parameter
values, but the deformation is more chaotic at later times. Similar dynamics are also observed
for Eo = 5 as shown in Fig. 4. It can be seen in Fig. 4(a) that the bubble migrates toward
the bottom wall for Re = 100, whereas for Re = 120 and 140, it moves toward the top
wall. However, it is to be noted that the path chosen by the bubble is arbitrary, and may
be influenced by the initial conditions. As the value of Eo considered in Fig. 4 is higher
than that in Fig. 3, the surface tension is less dominant, and the bubble undergoes larger
deformation for the cases considered in Fig. 4 as compared to those in Fig. 3.
In order to get more insight into the dynamics of the bubble, the spatio-temporal evolution
of horizontal and vertical components of velocity are plotted in Figs. 5 and 6 for Re = 70
and Eo = 1, and Re = 120 and Eo = 5, respectively. It can be seen in Fig. 5 that the
horizontal velocity is higher near the centre of the channel as compared to the near-wall
regions. Along the centreline, the horizontal velocity is maximum near the troughs of the
converging-diverging channel. When the bubble crosses the trough regions of the converging-
diverging channel, it elongates to become prolate shaped bubble. It deforms to an oblate
shape as it reaches the crest regions of the channel. This effect is more pronounced for
Eo = 5 due to the lesser dominance of surface tension force as compared to that for Eo = 1.
The contours of vertical velocity component show the appearance of four lobes of maximum
and minimum vertical velocity inside the bubble as it migrates toward the bottom wall.
Next, we investigate the effect of half-amplitude, h of the converging-diverging symmet-
rical channel (Φ = π) in Fig. 7 for two typical sets of Reynolds and Eötvös numbers. It is
to be noted that h = 0 case corresponds to bubble motion inside a straight channel. It can
be seen that for h = 0, the bubble migrates toward the top wall (classical Segré–Silberberg
effect). It is observed that bubble starts oscillating for h ≥ 0.25, and the oscillation am-
plitude magnifies with increasing the value of h. Inspection of aspect ratio of the bubble
plotted in Figs. 7(b) and (d) reveals that the shape of the bubble remains spherical for
h = 0, but undergoes periodic oblate-prolate deformation for intermediate value of h. For

7
h ≥ 0.5 the deformation of bubble is periodic till t = 90 (approximately), but experiences a
larger deformation for t > 90.
All the results presented so far are for a symmetrical channel, i.e. when top and bottom
walls are described by y = R + hsin 2πx 2πx
 
λ
+ Φ , and y = −R + hsin λ
, with Φ = π.
Finally, we investigate the bubble migration through asymmetrical channels by setting Φ = 0
and π/2, and comparing the variation of yCG and Ar obtained for the symmetrical channel
(Φ = π) in Fig. 8 for Re = 60, Eo = 1, ρr = 10−3 and µr = 10−2 . It can be seen that
the oscillation of center-of-gravity of the bubble is more in case of asymmetrical channel
with Φ = 0 with the rest of the parameters kept the same. This oscillation is a resultant
of Segré-Silberberg effect as well as asymmetrical nature of the channel for Φ = π and π/2.
The larger deformation of the bubble shown in Fig. 8b in case of channel with Φ = π can
be attributed to the presence of bigger recirculation zones as compared to that for channels
with Φ = 0 and π/2.

V. CONCLUDING REMARKS

In this work, we investigate the dynamics of bubble motion through converging-diverging


channels with different phase difference between the walls, using an open-source finite vol-
ume flow solver, Gerris. A parametric study is conducted, and the effects of various
dimensionless parameters, such as Reynolds number, Eötvös number, and amplitude of
converging-diverging channel are investigated. We found that increasing Re and h increases
the oscillation of the bubble. For higher Eo value considered, the bubble undergoes oblate-
prolate periodic deformation at early times, which becomes chaotic at later times, as it
travels through trough to crest regions of the converging-diverging channel. We also observe
the effect of Segré-Silberberg phenomena enhances the oscillation of the bubble inside the
converging-diverging channel. We have shown that the bubble migrating inside a asymmet-
rical channel with Φ = 0 undergoes maximum oscillation. The oscillation of shape of bubble
and the complex path followed by the bubble inside the converging-diverging channel can

8
lead to more mixing, which is desirable in many applications involving small-scale flows.

[1] S. Haeberle and R. Zengerle, “Microfluidic platforms for lab-on-a-chip applications,” Lab Chip
7, 1094 (2007).
[2] H. A. Stone, A. D. Stroock, and A. Ajdari, “Engineering flows in small devices: microfluidics
towards a lab-on-a-chip,” Annu. Rev. Fluid Mech. 36, 318 (2004).
[3] T. -L. Le, J. -C. Chen, B. -C. Shen, F. -S. Hwub and H. -B. Nguyen, “Numerical investigation
of the thermocapillary actuation behavior of a droplet in a microchannel,” Int. J. Heat Mass
Trans. 83, 721 (2015).
[4] t. . B. S. Tripathi, A. Prabhakar, N. Kumar, S. G. Singh and A. Agrawal, Biomed. Microdevices
15(3), 415 (2013).
[5] M. K. Tripathi, and K. C. Sahu, G. Karapetsas, K. Sefiane, and O. K. Matar, “Non-isothermal
bubble rise: non-monotonic dependence of surface tension on temperature,” J. Fluid Mech.
763, 82 (2015).
[6] M. K. Tripathi, and K. C. Sahu, G. Karapetsas, and O. K. Matar, “Bubble rise dynamics in
a viscoplastic material,” J. Non-Newt. Fluid Mech. (In Press) (2015).
[7] M. K. Tripathi, K. C. Sahu, and R. Govindarajan, “Why a falling drop does not in general
behave like a rising bubble,” Nature Scientific Reports 4, 4771 (2014).
[8] M. K. Tripathi, K. C. Sahu, and R. Govindarajan, “Dynamics of an initially spherical bubble
rising in quiescent liquid,” Nat. Commun. 6, 6268 (2015).
[9] K. C. Sahu and R. Govindarajan, “Stability of flow through a slowly diverging pipe,” J. Fluid
Mech. 531, 325 (2005).
[10] K. C. Sahu, “A possible linear instability mechanism in small-scale pipe flows,” Proceedings
of the Sixth IUTAM Symposium on Laminar-Turbulent Transition, Bangalore, India (2004).
[11] B. M. Jose and T. Cubaud, “Droplet arrangement and coalescence in diverging/converging
microchannels,” Microfluid. Nanofluid. 12, 687 (2012).
[12] I. J. Sobey, “On flow through furrowed channels. Part 1. Calculated flow patterns,” J. Fluid
Mech. 96, 1 (1980).
[13] K. D. Stephanoff, I. J. Sobey, and B. J. Bellhouse, “On flow through furrowed channels. Part
2. Observed flow patterns,” J. Fluid Mech. 96, 27 (1980).

9
[14] S. Blancher and R. Creff, “Analysis of convective hydrodynamics instabilities in a symmetric
wavy channel,” Phys. Fluids 16, (10), 3726 (2004).
[15] S. Blancher, R. Creff, and P. L. Quere, “Effect of Tollmien Schlichting wave on convective heat
transfer in a wavy channel. Part I: Linear analysis,” Int. J. Heat Fluid Flow 19, 39 (1998).
[16] K. J. Cho, M. Kim, and H. D. Shin, “Linear stability of two-dimensional steady flow in
wavy-walled channels,” Fluid Dynamics Research 23, 349 (1998).
[17] T. A. Rush, T. A. Newell, and A. M. Jacobi, “An experimental study of flow and heat transfer
in sinusoidal wavy passages,” International Journal of Heat and Mass Transfer 42, 1541 (1999).
[18] S. Selvarajan, E. G. Tulapurkara, and V. V. Ram, “Stability characteristics of wavy walled
channel flows,” Phys. Fluids 11, 3, 579 (1999).
[19] J. Szumbarski and J. M. Floryan, “Transient disturbance growth in a corrugated channel,” J.
Fluid Mech. 568, 243 (2006).
[20] V. Duryodhan, S. Singh, and A. Agarwal, “Liquid flow through a diverging microchannel,”
Microfluid Nanofluid 14, 53 (2013).
[21] K. Stone and S. Vanka, “Numerical study of developing flow and heat transfer in a wavy
passage,” J. Fluids Eng. 121, 713 (1999).
[22] K. C. Sahu, “The instability of flow through a slowly diverging pipe with viscous heating,” J.
Fluids Eng. 133, 071201 (2011).
[23] E. Santacesaria et al., “Use of a corrugated plates heat exchanger reactor for obtaining
biodiesel with very high productivity,” Energy Fuels 23, 5206 (2009).
[24] F. Théron, Z. Anxionnaz-Minvielle and M. Cabassud, C. Gourdon and P. Tochon, “Charac-
terization of the performances of an innovative heat-exchanger/reactor,” Chem. Eng. Process.
82, 30 (2014).
[25] H. Metwally and R. Manglik, “Enhanced heat transfer due to curvature-induced lateral vor-
tices in laminar flows in sinusoidal corrugated-plate channels,” Int. J. Heat Mass Transfer 47,
2283 (2004).
[26] M. Gradeck and M. Lebouche, “Two-phase gas-liquid flow in horizontal corrugated channels,”
Int. J. Multiphase Flow 26, 435 (2000).
[27] P. Vlasogiannis, G. Karagiannis, P. Argyropoulos, and V. Bontozoglou, “Airwater two-phase
flow and heat transfer in a plate heat exchanger,” Int. J. Multiphase Flow 28, 757 (2002).
[28] K. Nilpueng and S. Wongwises, “Flow pattern and pressure drop of vertical upward gasliquid

10
flow in sinusoidal wavy channels,” Exp. Therm. Fluid Sci 30, 513 (2006).
[29] P. Andreussi, A. Paglianti, and F. S. Silva, “Dispersed bubble flow in horizontal pipes,” Chem.
Eng. Sci. 54, 1101 (1999).
[30] Y. Murai, H. Fukuda, Y. Oishi, Y. Kodamai and F. Yamamoto, “Skin friction reduction by
large air bubbles in a horizontal channel flow,” Int. J. Multiphase Flow 33, 147 (2007).
[31] L. Böhm, T. Kurita, K. Kimura, and M. Kraume, “Rising behaviour of single bubbles in
narrow rectangular channels in Newtonian and non-Newtonian liquids,” Int. J. Multiphase
Flow 65, 11 (2014).
[32] G. Segré and A. Silberberg, “Radial Particle Displacements in Poiseuille Flow of Suspensions,”
Nature 189, 209 (1961).
[33] D. H. Douglas-Hamilton, N. G. Smith, C. E. Kuster, J. P. W. Vermeiden and G. C. Althouse,
“Capillary-Loaded Particle Fluid Dynamics: Effect on Estimation of Sperm Concentration,”
J Androl 26, 115 (2005).
[34] T.-W. Pan and R. Glowinski, “Direct simulation of the motion of neutrally buoyant circular
cylinders in plane Poiseuille flow,” J. Comput. Phys 181, 260 (2002).
[35] J.-P. Matas, J. F. Morris, and E. Guazzelli, “Lateral force on a rigid sphere in large-inertia
laminar pipe flow,” J. Fluid Mech. 621, 59 (2009).
[36] J. Brackbill, D. B. Kothe, and C. Zemach, “A continuum method for modeling surface ten-
sion,” J. Comput. Phys 100, 335 (1992).
[37] S. Popinet, “Gerris: a tree-based adaptive solver for the incompressible Euler equations in
complex geometries,” J. Comput. Phys 190, 572 (2003).
[38] G. Segré and A. Silberberg, “Behaviour of macroscopic rigid spheres in Poiseuille flow. Part
2. Experimental results and interpretation,” J. Fluid Mech. 14, 136 (1962).
[39] G. Segré and A. Silberberg, “Behaviour of macroscopic rigid spheres in Poiseuille flow, Part
1. Determination of local concentration by statistical analysis of particle passages through
crossed light beams,” J. Fluid Mech. 14, 115 (1962).

11
List of figure captions

Fig. 1: Schematic diagram (not to scale) of the bubble motion inside a two-
dimensional converging-diverging channel. The top and bottom walls are governed
by y = R + hsin 2πx 2πx
 
λ
+ Φ , and y = −R + hsin λ
, wherein h, λ and Φ are the
half-amplitude, the wavelength of the corrugated channel, and phase difference between
the walls, respectively. For the schematic shown Φ = π. The distance between the walls
in the straight section and the radius of the bubble are 2R and r, respectively, such that
R/r = 4. The length of the horizontal section at the inlet is 12r. The bubble is placed at
(x, y) = (9r, 0) of the channel. The total length of the channel is considered to be 78r, and
λ = 8r.

Fig. 2: Comparison of shapes of the bubble (a,c) at x = 9.537 and (b,d) at x = 14.37
for two different gird refinements. The panels (a,b) and (c,d) correspond to Eo = 10
and Eo = 25, respectively. Grid − 1 (shown by solid line): minimum grid sizes near the
boundary and the fluid regions are 0.031 and 0.015, respectively. Grid − 2 (shown by
dashed line): minimum grid sizes near the boundary and the fluid regions are 0.015 and
0.008, respectively. The rest of the parameter values are ρr = 10−3 , µr = 10−2 and Re = 100.

Fig. 3: The variation of (a) yCG , (b) vCG , and (c) Ar with time for different values
of Re. The rest of the parameter values are ρr = 10−3 , µr = 10−2 and Eo = 1.

Fig. 4: The variation of (a) yCG , (b) vCG , and (c) Ar with time for different values
of Re. The rest of the parameter values are ρr = 10−3 , µr = 10−2 and Eo = 5.

Fig. 5: The spatio-temporal evolution of horizontal (top panel) and vertical (bottom
wall) velocity components at different times. The rest of the parameter values are
ρr = 10−3 , µr = 10−2 , Re = 70, and Eo = 1. The color-bars presented at the bottom of the
figure correspond to horizontal and vertical velocity components.

Fig. 6: The spatio-temporal evolution of horizontal (top panel) and vertical (bottom
wall) velocity components at different times. The rest of the parameter values are

12
ρr = 10−3 , µr = 10−2 , Re = 120, and Eo = 5. The color-bars presented at the bottom of
the figure correspond to horizontal and vertical velocity components.

Fig. 7: Time evolution of yCG and Ar for different values of h: (a,b) Re = 80 and
Eo = 1, (c,d) Re = 140 and Eo = 5. The rest of the parameter values are ρr = 10−3 ,
µr = 10−2 .

Fig. 8: Time evolution of yCG and Ar for channels having different values of Φ.
The parameter values are Re = 60, Eo = 1, ρr = 10−3 and µr = 10−2 .

13
FIG. 1: Schematic diagram (not to scale) of the bubble motion inside a two-dimensional converging-
diverging channel. The top and bottom walls are governed by y = R + hsin 2πx

λ + Φ , and

y = −R + hsin 2πx

λ , wherein h, λ and Φ are the half-amplitude, the wavelength of the corrugated

channel, and phase difference between the walls, respectively. For the schematic shown Φ = π.
The distance between the walls in the straight section and the radius of the bubble are 2R and
r, respectively, such that R/r = 4. The length of the horizontal section at the inlet is 12r. The
bubble is placed at (x, y) = (9r, 0) of the channel. The total length of the channel is considered to
be 78r, and λ = 8r.

14
(a) (b)

(c) (d)

FIG. 2: Comparison of shapes of the bubble (a,c) at x = 9.537 and (b,d) at x = 14.37 for two
different gird refinements. The panels (a,b) and (c,d) correspond to Eo = 10 and Eo = 25,
respectively. Grid − 1 (shown by solid line): minimum grid sizes near the boundary and the fluid
regions are 0.031 and 0.015, respectively. Grid−2 (shown by dashed line): minimum grid sizes near
the boundary and the fluid regions are 0.015 and 0.008, respectively. The rest of the parameter
values are ρr = 10−3 , µr = 10−2 and Re = 100.

15
(a) (b)
0 0.1 Re
60
70
0.05 80
-0.5
yCG vCG 0
Re
60 -0.05
-1 70
80
-0.1
0 20 40 60 80 100 120 0 20 40 60 80 100 120
t t

(c)
1.1
Re
60
1.05 70
80

Ar 1

0.95

0.9
0 20 40 60 80 100 120
t

FIG. 3: The variation of (a) yCG , (b) vCG , and (c) Ar with time for different values of Re. The
rest of the parameter values are ρr = 10−3 , µr = 10−2 and Eo = 1.

16
(a) (b)
0.1
Re Re
0.6
100 100
120 0.05 120
0.4
140 140
0
yCG0.2 vCG
0 -0.05

-0.2
-0.1

-0.4
0 20 40 60 80 100 120 0 20 40 60 80 100 120
t t

(c)
Re
1.2 100
120
1.1 140

Ar 1

0.9

0.8

0.7
0 20 40 60 80 100 120
t

FIG. 4: The variation of (a) yCG , (b) vCG , and (c) Ar with time for different values of Re. The
rest of the parameter values are ρr = 10−3 , µr = 10−2 and Eo = 5.

17
t = 20

t = 60

t = 80

t = 100

FIG. 5: The spatio-temporal evolution of horizontal (top panel) and vertical (bottom wall) velocity
components at different times. The rest of the parameter values are ρr = 10−3 , µr = 10−2 , Re = 70,
and Eo = 1. The color-bars presented at the bottom of the figure correspond to horizontal and
vertical velocity components.

18
t = 20

t = 60

t = 80

t = 100

FIG. 6: The spatio-temporal evolution of horizontal (top panel) and vertical (bottom wall) velocity
components at different times. The rest of the parameter values are ρr = 10−3 , µr = 10−2 ,
Re = 120, and Eo = 5. The color-bars presented at the bottom of the figure correspond to
horizontal and vertical velocity components.

19
(a) (b)
1.5 1.1
h h
1 0 0
0.25 1.05 0.25
0.5 0.5 0.5

yCG 0 Ar 1

-0.5
0.95
-1

-1.5 0.9
0 50 100 0 50 100
t t

(c) (d)
1
h h
0 1.2 0
0.5 0.25 0.25
0.5 0.5

yCG 0 Ar 1

-0.5
0.8

-1
0 50 100 0 50 100
t t

FIG. 7: Time evolution of yCG and Ar for different values of h: (a,b) Re = 80 and Eo = 1, (c,d)
Re = 140 and Eo = 5. The rest of the parameter values are ρr = 10−3 , µr = 10−2 .

20
(a) (b)
Asymmetric (Φ = 0 )
0 1.05
Asymmetric (Φ = π/2)
Symmetric (Φ = π)
-0.5
yCG Ar 1
-1

Asymmetric (Φ = 0)
-1.5 0.95
Asymmetric (Φ= π/2)
Symmetric (Φ = π )
-2
0 25 50 75 100 125 0 25 50 75 100 125
t t

FIG. 8: Time evolution of yCG and Ar for channels having different values of Φ. The parameter
values are Re = 60, Eo = 1, ρr = 10−3 and µr = 10−2 .

21

Vous aimerez peut-être aussi