ArticlePDF Available

Effect of buoyancy on the motion of long bubbles in horizontal tubes

Authors:

Abstract and Figures

As a confined long bubble translates along a horizontal liquid-filled tube, a thin film of liquid is formed on the tube wall. For negligible inertial and buoyancy effects, respectively, small Reynolds (Re) and Bond (Bo) numbers, the thickness of the liquid film depends only on the flow capillary number (Ca). However, buoyancy effects are no longer negligible as the diameter of the tube reaches millimeter length scales, which corresponds to finite values of Bo. We perform experiments and theoretical analysis for a long bubble in a horizontal tube to investigate the effect of Bond number (0.05<Bo<0.5) on the thickness of the liquid film and the bubble orientation at different capillary numbers 10−3<Ca<10−1. We investigate several features of the lubricating film around the bubble. (i) Due to the gravitational effects, the film deposited on the upper wall of the channel is thinner than the film at the bottom wall. We extend the available theory for the film thickness at the front of the bubble in a two-dimensional geometry at low capillary numbers Ca<10−3 and finite Bo to account for the effect of larger Ca. The resulting model shows very good agreement with the present experimental measurements. (ii) Due to the asymmetry in the liquid film thickness and the consequent drainage of the liquid from the top to the bottom of the tube, the bubble is inclined relative to the channel centerline and our side-view visualizations allow direct quantification of the inclination angle, which increases with both Bo and Ca. While the inclination angle at the top is smaller than that at the bottom of the tube, the average of these two values follows the predictions of a mass balance analysis in the central region of the bubble. (iii) The inclination of the bubble causes the thickness of the thin film at the back of the bubble to depend on the length of the bubble, whereas the thickness at the front of the bubble does not depend on the bubble length.
Content may be subject to copyright.
PHYSICAL REVIEW FLUIDS 2, 094304 (2017)
Effect of buoyancy on the motion of long bubbles in horizontal tubes
Omer Atasi,1,2,*Sepideh Khodaparast,2,Benoit Scheid,1,and Howard A. Stone2,§
1Transfers, Interfaces and Processes, Universite Libre de Bruxelles, Brussels 1050, Belgium
2Department of Mechanical and Aerospace Engineering, Princeton University,
Princeton, New Jersey 08544, USA
(Received 21 November 2016; published 26 September 2017)
As a confined long bubble translates along a horizontal liquid-filled tube, a thin film of
liquid is formed on the tube wall. For negligible inertial and buoyancy effects, respectively,
small Reynolds (Re) and Bond (Bo) numbers, the thickness of the liquid film depends only
on the flow capillary number (Ca). However, buoyancy effects are no longer negligible as
the diameter of the tube reaches millimeter length scales, which corresponds to finite values
of Bo. We perform experiments and theoretical analysis for a long bubble in a horizontal
tube to investigate the effect of Bond number (0.05 <Bo <0.5) on the thickness of the
liquid film and the bubble orientation at different capillary numbers 103<Ca <101.
We investigate several features of the lubricating film around the bubble. (i) Due to the
gravitational effects, the film deposited on the upper wall of the channel is thinner than the
film at the bottom wall. We extend the available theory for the film thickness at the front of
the bubble in a two-dimensional geometry at low capillary numbers Ca <103and finite
Bo to account for the effect of larger Ca. The resulting model shows very good agreement
with the present experimental measurements. (ii) Due to the asymmetry in the liquid film
thickness and the consequent drainage of the liquid from the top to the bottom of the tube,
the bubble is inclined relative to the channel centerline and our side-view visualizations
allow direct quantification of the inclination angle, which increases with both Bo and Ca.
While the inclination angle at the top is smaller than that at the bottom of the tube, the
average of these two values follows the predictions of a mass balance analysis in the central
region of the bubble. (iii) The inclination of the bubble causes the thickness of the thin film
at the back of the bubble to depend on the length of the bubble, whereas the thickness at
the front of the bubble does not depend on the bubble length.
DOI: 10.1103/PhysRevFluids.2.094304
I. INTRODUCTION
The motion of elongated bubbles in small liquid-filled confined geometries is a hydrodynamic
problem with a large variety of industrial and medical applications in lubrication, heat and mass
transfer, oil extraction, and treatment of pulmonary disorders. The problem involves translation
of a low-viscosity fluid at average speed Uinto a higher-viscosity liquid with density, viscosity,
and surface tension of ρ,μ, and γ, respectively; the density of the bubble is negligible relative to
that of the continuous liquid phase (Fig. 1). Characterization of the motion and the shape of the
bubble requires consideration of the Reynolds number Re =ρUa
μ, where ais the radius of the tube,
in order to characterize the relative magnitudes of inertial and viscous effects, the Bond number
Bo =ρga2
γ, where gdenotes the gravitational acceleration, in order to characterize buoyancy effects
relative to interfacial tension effects, and the capillary number Ca =μU
γ, in order to assess the
relative magnitudes of viscous and interfacial tension effects. In microscale geometries, the effects
*oatasi@ulb.ac.be
sepidehk@princeton.edu
bscheid@ulb.ac.be
§hastone@princeton.edu
2469-990X/2017/2(9)/094304(17) 094304-1 ©2017 American Physical Society
ATASI, KHODAPARAST, SCHEID, AND STONE
+
hfront
hfront
hback
hback
U
L
g
,
a
++
--
-
FIG. 1. Schematic of a typical elongated bubble translating at capillary number Ca 102in a horizontal
tube of radius a. The bubble is immersed in a viscous liquid of density ρand viscosity μ. Here Uand Lare the
velocity and the length of the bubble, respectively. A spherical cap is present at the front of the bubble for low
to moderate capillary numbers, followed by a transition region and a lubrication film. Because gravity gacts
perpendicular to the tube axis, the bubble is inclined and the thin film has different thicknesses at the front (h+
front
and h
front) compared to the back (h+
back and h
back). At the back of the bubble, a capillary wave is present [1].
of buoyancy and inertia are often negligible (Bo,Re 1) and the motion of elongated bubbles can
be described by a single dimensionless number, namely, the capillary number Ca. However, as the
dimensions of the geometry increase to millimeter lengths, gravitational effects have significant
impacts on the shape and dynamics of the bubbles. In this paper we aim to quantify such effects for
long bubbles in millimeter diameter tubes at moderate capillary numbers and in the absence of inertia.
It was first observed by Fairbrother and Stubbs [2] that in a circular tube filled with liquid a
bubble moves slightly faster than the bulk liquid due to the formation of a thin liquid film between
the bubble and the boundary. In other words, the deposition of a thin liquid film on the wall does not
allow the entire volume of the more viscous liquid to be pushed out of a closed geometry using a
less viscous fluid. The amount of liquid left on the tube wall is observed to increase with the velocity
of the bubble. Accurate determination of the volume of liquid left on the bounding walls of the
geometry was pioneered by the theoretical work of Bretherton [1] and experimental measurements
of Taylor [3]. Bretherton’s theory predicts the thickness of the liquid film hto increase with Ca2/3
for Ca <103in the viscocapillary regime
¯
hBr =h
a=1.34 Ca2/3.(1)
Later, Aussillous and Quéré provided a scaling analysis to extend Bretherton’s theory to a wider
range of Ca by considering the contribution of the liquid film thickness in defining the curvature at
the front of the bubble [4]:
¯
h=h
a=1.34 Ca2/3
1+3.35 Ca2/3.(2)
The empirical coefficient in Eq. (2), 3.35, was found by fitting the result of their scaling analysis to
the experimental measurements of Taylor for capillary numbers up to Ca =2[3].
More recently, several attempts have been made to investigate the gravitational effects on different
features of the shape and the dynamics of confined elongated bubbles [510]. Such effects were gen-
erally neglected in the small horizontal geometries discussed above. A summary of the most relevant
works, including studies of the motion of semi-infinite low-viscosity air fingers progressing in liquid
flows, is presented in Table I. In general, two main features have been reported in the literature.
(i) In the cross-sectional plane, buoyancy causes asymmetry in the distribution of the liquid
film around the bubble. In two-dimensional (2D) geometries, this effect simply leads to a thinner
liquid film at the top compared to the bottom of the tube. For example, Jensen et al. showed that
for Ca <103the liquid film at the front of a long finger of gas translating in viscous liquid can
094304-2
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
TABLE I. Previous research on the motion of bubbles in horizontal and inclined channels with significant gravitational effects. The Saffman-Taylor problem
refers to the motion of a low-viscosity fluid displacing a higher-viscosity fluid in a Hele-Shaw geometry. Here θis defined as the angle between the orientation of
the channel and the vertical direction, ARis the aspect ratio of the cross section of the channel defined as the ratio of the width to the height, and Lis the length of
the bubble divided by the channel radius.
Reference Problem and geometry Approach Flow parameters Focus of study
[11] Saffman-Taylor finger, 2D
(Hele-Shaw)
Theory, experiments Ca <103,0<Bo <3.3 Effect of Ca and Bo on the shape of the
front meniscus, pressure jump across
the interface, and film thickness at
the front.
[12] Saffman-Taylor finger, 2D
(Hele-Shaw)
Theory, experiments Effect of buoyancy on the shape of the
front meniscus in tilted channels.
[5] Liquid slug bounded
between two air bubbles,
2D
Theory 0 <Ca <0.05, 0 <Bo <2, 0 <180Effect of buoyancy on the shape of the
front meniscus in tilted channels.
[6] Liquid slug bounded
between two air bubbles,
2D
Numerics 0.03 <Ca <0.4, 0 <Bo <0.6, 0 <180Effect of buoyancy on the flow pattern
in the liquid plug region.
[7] Infinite finger, 3D
(rectangular cross section)
Numerics 0.002 <Ca <20, Bo =0,1,2.5, 1 <A
R<8 Effect of channel aspect ratio AR,Ca,
and Bo on the cross-sectional shape
of the bubble and pressure jump
across the front meniscus.
[8] Slug flow, 3D (circular cross
section)
Experiments 104<Ca <0.4 Effect of Ca, Bo, and Re on the
distribution of the liquid film
thickness around the bubble.
[10] Slug flow, 3D (circular cross
section)
Theory, experiments 0.03 <Ca <0.2, Bo =0.07,0.17,0.25 (i) Effect of buoyancy on the
distribution of the liquid film
thickness around the bubble. (ii)
Velocity measurement in the liquid
plug.
Present
study
Isolated elongated bubble,
3D (circular cross section)
Theory, experiments 103<Ca <0.15, 0.05 <Bo <0.42, 10 <L<35 (i) Effect of Ca and Bo on the bubble
inclination angle and the film
thickness. (ii) Effect of bubble length
on the film thickness.
094304-3
ATASI, KHODAPARAST, SCHEID, AND STONE
be predicted analytically by ¯
h±=¯
hBr
1±Bo , where the +and superscripts correspond to the top
and the bottom of the channel, respectively [11]. However, this model has not been tested against
experimental measurements performed at finite Bo. Thus, here we modify the theory proposed by
Jensen et al. to account for finite values of Ca and test its validity against experimental measurements
for long bubbles in a tube of circular cross section.
(ii) In the direction of the flow, a bubble is inclined with respect to the horizontal axis of symmetry
[see the schematic in Fig. 1(a)]. As a result, a uniform film thickness region between the front and the
back menisci, as predicted by Bretherton, no longer exists when gravitational effects are significant.
Leung et al. showed that the inclination of a bubble is caused by the drainage of the liquid in the
thin film region from the top to the bottom of the tube. Their experimental quantification of the
drainage volume followed well the predictions of a mass balance analysis proposed in their study
for Bo 0.25 [10]. The inclination angle was observed to be fairly constant along the length of
the bubble [8,10]; however, direct experimental quantification of the angle and its correlation with
the flow parameters has not been reported. In this study we quantify the angle of the inclination for
finite values of Bo and a range of Ca. Due to the inclination, the length of the bubble is expected
to be an important parameter in determining the thickness of the film along the bubble, which we
demonstrate and quantify below.
In a continuous train of bubbles separated by slugs of liquid on the orders of the tube radius and
when buoyancy effects are significant, consecutive bubbles are observed to influence one another
due to hydrodynamic interactions [6]. This effect modifies the flow pattern in front of a translating
bubble and therefore alters the shape of the front meniscus and the film thickness with respect to
that of a single translating isolated bubble. To avoid such hydrodynamic interactions, here we study
isolated independent bubbles in horizontal tubes of circular cross section.
We provide a theoretical analysis by combining the available models in the literature for different
geometrical and flow parameters, which predicts the evolution of the liquid film thickness around
an elongated air bubble in a horizontal tube of circular cross section. Also, we report experimental
data to quantify the buoyancy effects on the motion of confined elongated bubbles. In particular, we
perform bright-field side-view microscopy to directly observe the asymmetries in the flow caused
by buoyancy. This approach allows direct quantification of the inclination angle and the distribution
of the film thickness along the length of the bubble. Different Ca and Bo numbers are tested by
systematically varying the velocity of the liquid phase and tube diameter. We also identify the
influence of bubble length.
II. EXPERIMENTAL SETUP
A. Refractive index matching
Experiments were performed in glass capillaries of radii a=0.51, 0.75, 1.15, and 1.51 mm and
length L=10 cm, which were submerged in a refractive index matching box filled with glycerol.
The diameters of the tubes were measured optically using precalibrated microscope objectives. The
region of interest (ROI) was observed through the straight sidewall of the box (Fig. 2). This approach,
together with the identical refractive indices of the working fluid, tube material, and surrounding
glycerol pool (n=1.47), ensured no optical distortions of the ROI due to light refraction at the
outer and inner curved walls of the capillary [13]. Moreover, the refractive index matching design
prevented total internal reflections and thus allowed us to resolve the thin liquid film formed between
the two-phase interface and the inner tube wall.
B. Experiments
All experiments were performed at T=25 C using glycerol as the continuous liquid phase
(viscosity μ=0.84 Pa s, density ρ=1186 kg/m3, and surface tension γ=0.063 N/m) and air
as the dispersed gas phase. The glycerol used in the experiments was kept in sealed containers to
minimize the effect of humidification and the viscosity was measured before and after the tests to
094304-4
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
P)
a
FIG. 2. Schematic of the experimental setup showing a capillary placed in a box filled with refractive index
matching liquid to eliminate optical distortions.
ensure no water absorption. The glass capillary, which was fixed to the sides of the refractive index
matching box, was connected to a syringe pump using a long flexible tube at the entrance, while at
the exit another flexible tube emptied the outlet fluid at atmospheric pressure (Fig. 2). The entire
volume of all connection tubes was filled with glycerol. At the beginning of each experiment, a long
air finger was created at the inlet of the flexible tube by slightly elevating the tube inlet relative to the
outlet. The air-filled inlet was then connected to the syringe pump needle that was filled with liquid
to create a confined bubble of length L20a. The bubble was carried into the ROI as glycerol was
pumped at a constant flow rate. This procedure ensured no acceleration or deformation of the bubble
caused by the injection geometry.
C. Visualization and image processing
Experimental measurements were obtained by processing the digital shadowgraphy images. The
refractive index matching box was placed between a green collimated light source and a home-made
tube microscope equipped with long working distance objectives of 5×and 10×magnification (with
spatial resolutions of, respectively, 2.44 and 1.22 μm per pixel). Bright-field microscopy allowed
visualization and consequently detection of a sharp interface between the gas and the liquid phases
using a Canny edge detection algorithm [14]. Images of the ROI were recorded by a Nikon D5100
camera at 30 frames per second. Film thickness measurements were performed using the bright-field
images captured at the front and the back of the bubble. The maximum relative error in the film
thickness measurements reported for the smallest Ca is 14%, while the rest of the measurements
have a maximum experimental error of 5%.
Obtaining high spatial resolution in the images requires high magnification in the optical setup,
which consequently leads to a relatively small field of view (FOV). For example, using the 5×
objective in our experiments, we could visualize only about 2 mm of the tube length in the images,
which is obviously not adequate to capture the axial evolution of the shape of the long bubbles
[Fig. 3(a)]. In order to construct an image of the full length of the bubble that does not fit in the FOV,
we applied the postprocessing time-strip method to the sequence of shadowgraphy images [15].
094304-5
ATASI, KHODAPARAST, SCHEID, AND STONE
(a) (b) (c)A
A’ L
1 mm 2 s1 mm
2a
FIG. 3. Sample visualization results for an elongated bubble of length L=22 mm in a horizontal tube of
radius a=1.51 mm at Ca =0.076 and Bo =0.42. The dashed horizontal line shows the centerline of the tube.
(a) Sample image of the grayscale bright-field microscopy. The bubble translates from the left to the right.
Time-strip analysis is performed along a sample line such as section AA. The field of view in a typical image
covers less than a 2-mm length of the tube. (b) Time-strip analysis result showing evolution of section AAin
time. Note that the horizontal axis represents time. (c) The horizontal time axis in (b) is converted to the length
axis using the bubble velocity. The inclination angle is measured by fitting a straight line to the bubble interface
as shown by the red lines. Note that images similar to the one presented in (c) are only used to measure the
length and the inclination angle of the bubble.
In this approach, we followed the evolution of a specified vertical line of pixels along the tube
diameter [Fig. 3(a)] for a long period of time starting before the arrival of the bubble front until after
departure of the back of the bubble [Fig. 3(b)]. Knowing the bubble velocity, time is then converted
to the displacement of the bubble and the shape of the bubble is constructed [Fig. 3(c)]. The final
image was then processed to obtain the length and the inclination angle of the bubble. The difference
between the thickness of the liquid film at the top and at the bottom and the inclination of the bubbles
are clearly noticeable in Figs. 3(b) and 3(c). In order to obtain the inclination angle, a straight line
was fitted to the detected air-glycerol interface. The maximum experimental error in identifying
the inclination angle at the lowest Bo and Ca is 30%, while the rest of the measurements contain
a maximum experimental error of 10%. The capillary number Ca =μU
γwas calculated based on
the bubble velocity measured by tracking the nose of the bubble. The high optical resolution used
in the present study allowed determination of Ca down to values on the order of Ca =106.The
spatial resolution of the time-strip image perpendicular to the flow is fixed by the physical pixel
size and the objective magnification. On the other hand, the axial resolution of the time-strip image
along the flow depends on the image acquisition frequency and on the speed of the bubble. The
minimum and maximum axial resolutions of the time-strip images in the experiments reported here
were, respectively, 256 and 1.3 μm. Note that the low optical resolution observed at the nose and
the back of the bubble in Fig. 3(c) is due to the large velocity of the bubble at larger Ca. This effect,
however, does not influence measurement of the bubble length or the inclination angle.
The camera is leveled horizontally prior to the start of the experiments using a digital level with an
accuracy of 104rad. Furthermore, in order to ensure that the tube is horizontal, the refractive index
matching box is mounted on a stage that is translated in the axial xdirection. Two images are captured
as the box is moved between two points that are axially 5 cm apart. Comparison between these two
images ensures no inclination in the ROI within the accuracy of our measurements (104rad).
D. Experimental observation
Examples of the shape of the nose, the central region, and the back of the bubbles at different
Ca and Bo are shown in Fig. 4. Two major differences are observed when comparing the present
side-view visualizations at finite Bo numbers with the classic predictions of bubble shape with
negligible buoyancy effects. (i) Due to non-negligible buoyancy effects, the liquid film at the front
of the bubble is not uniformly distributed but instead is thinner at the top than at the bottom of the
tube. (ii) Furthermore, the back of the bubble is elevated relative to the nose. This feature was shown
094304-6
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
(a)
Ca = 0.076Ca = 0.006
Ca = 0.012 Ca = 0.063
(b)
(c)
Ca = 0.006 Ca = 0.064
a = 0.75 mma = 1.15 mma = 1.51 mm
FIG. 4. Examples of bubble shapes at different Ca and Bo: (a) Bo =0.11, (b) Bo =0.25, and (c) Bo =0.42.
Bubbles move from the left to the right. Images contain experimental visualization of the nose, the central part,
and the back of the bubble. The central region demonstrates a typical zone between the bubble nose and back,
where the liquid film at the top (bottom) acquires a linearly decreasing (increasing) thickness towards the back.
The back of the bubble exhibits a capillary wave, meaning a region of negative curvature [16]. The red dashed
lines indicate the centerline of the tube. Note that the optical magnifications used in visualization of flows at
different Bond numbers are not identical.
by Leung et al. [10] to be caused by the drainage of the liquid in the film from the top to the bottom
of the tube.
III. THEORETICAL PREDICTIONS
In this section we provide theoretical considerations to predict quantitatively both the film
thickness and the inclination angle as a function of Ca and Bo. The model presented here for the film
thickness at the front of the bubble is mainly based on the two-dimensional analysis proposed by
Jensen et al. [11] at low Ca and finite Bo in Hele-Shaw geometries. This model is extended here to
account for the effect of finite Ca using the correction that was originally suggested by Aussillous and
Quéré [4], and a physical argument is provided to show that the lateral curvature of the tube does not
affect the thin film thickness. The basic steps of Jensen et al. [11] are included in this section, wherever
necessary, to ensure the self-sufficiency of the arguments. Moreover, we utilize the mass balance
analysis proposed by Leung et al. [10] to obtain an estimate for the average inclination angle of the
bubble. In the next section, these theoretical models are compared to our experimental measurements
of the film thickness at the front and the inclination angle of the bubble, for a range of Ca and Bo (<1).
A. Thickness of the liquid film at the front of the bubble
Following the classical approach of Bretherton, we divide the front of the bubble into three
distinct regions (Fig. 5): region I, the capillary static (CS) region, where the pressure is hydrostatic;
region II, the transition region, where the surface tension force is balanced by the viscous forces;
094304-7
ATASI, KHODAPARAST, SCHEID, AND STONE
I II III
R
Rf+
Rf-
a
g
x
y
y = +
y = -Rf-
y = Rf+
y = 0
s+
s-
h+
h-
h+
h-
y = r-(x)
y = r+(x)
FIG. 5. Side-view schematic of the front of a bubble in a horizontal tube of radius a. The bubble moves
from left to right and is inclined with respect to the centerline of the tube. The +and superscripts stand for
the quantities at the top and the bottom, respectively. The bubble is divided in three regions. Region I indicates
the capillary static region, where the tip of the bubble forms a sphere of radius R. The tip of the bubble is
off-center with respect to the tube and is located at (x,y)=(0,). Region II indicates the transition region in
which the viscous forces are balanced by the capillary forces due to the variation of the curvature along the
interface. Between I and II, a matching region mris present where the interface shapes of I and II match. The
position of the interface in the matching region with respect to the tube centerline is at yR±
f. Region III
indicates the thin film region, where the axial flux (in the reference frame of the bubble) is balanced by the
lateral drainage flux of the liquid in the film.
and region III, the thin film region, where the axial flux (in the reference frame translating with the
bubble) is balanced by the lateral drainage.
1. Capillary static region
In this region, our approach follows closely that of Jensen et al. [11]. However, we extend the
original formulation by adding the condition that at higher Ca the thickness of the film is not
negligible compared to the radius of the front nose of the bubble [4]. Moreover, we consider the top
to bottom asymmetry in the interface height with respect to the tube centerline (Fig. 5).
The dimensionless film thickness at the front of the bubble ¯
h±
in a two-dimensional geometry
and at very low capillary numbers Ca 0 was determined by Jensen et al. [11]tobe
¯
h±
=hBr
1±Bo ,(3)
where hBr =1.337 Ca2/3is the classic prediction of the film thickness by Bretherton [1]atverylow
Ca and negligible buoyancy effects and ±indicates the top (+)orthebottom() relative to the
axis of symmetry. Both of these analyses were conducted by neglecting the azimuthal curvature of
the bubble.
For a two-dimensional configuration, in a Hele-Shaw cell, the pressure difference caused by
capillary forces between the liquid and the gas in the thin film region is δpf0, while in the CS
094304-8
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
region δpCS ≈−
γ
a, leading to a pressure difference between the thin film and the CS region of
δpfδpCS γ
a. In a tube of circular cross section and for thin liquid films, the pressure difference
caused by capillary forces between the liquid and the gas in the thin film region is δpf≈−
γ
adue to
the azimuthal curvature and in the CS region it is δpCS ≈−
2γ
a, which leads to a pressure difference
between the thin film and the CS region of δpfδp CS γ
a. Thus, the azimuthal curvature of the
interface does not change the pressure difference between the thin film region and the CS region in
a tube when compared to the two-dimensional case. We will therefore conduct this first part of our
analysis by not considering the azimuthal curvature.
In the CS region, where the pressure is hydrostatic, the coordinate system (x,y) has its yorigin at
the center of the channel and its xorigin at the tip of the bubble (Fig. 5). Due to buoyancy, the tip of
the bubble is elevated with respect to the center plane of the geometry and is located at y=a.
In the liquid phase, the pressure gradient is given by ∂p
∂y =−ρg, where ρis the density and gthe
gravitational acceleration. Integrating the pressure gradient with the Laplace pressure jump condition
at the interface, i.e., at y=r±(x), pgp=∓γκ±, where pgis the pressure inside the bubble, γ
the surface tension, and κthe curvature of the interface, yields p+ρgy pg=ρgr±±γκ±.The
curvature of a two-dimensional interface in a Hele-Shaw configuration is given by κ±=xxr±
[1+(xr±)2]3/2.
Since the gas pressure is constant, we can set pg=0 without loss of generality. Furthermore, we
can express the pressure in terms of the reduced pressure pr=p+ρgy such that pris independent
of yand use the scales afor rand γ
afor prto finally obtain
¯
pr=Bo¯
r±±¯
x¯
x¯
r±
[1 +(¯
x¯
r±)2]3/2,(4)
where Bo =ρga 2
γis the Bond number and an overbar denotes dimensionless variables. Note that
Eq. (4) is the same as the intermediate result of Jensen et al. [11] and is obtained in the same manner.
Next we multiply Eq. (4)by¯
x¯
r±and integrate assuming ¯
prconstant to find
¯
pr¯
r±=Bo
2¯
r±21
[1 +(¯
x¯
r±)2]1/2+c±,(5)
where c±is the integration constant. The hypothesis of constant reduced pressure in the CS region
will be verified subsequently. We evaluate this equation in the matching region, where we assume
that viscous effects become significant. In the matching region |¯
x¯
r±|1 and ¯
r±
R±
f
a, where
R±
fis the distance between the interface and the cell centerline (+) top and () bottom. Substituting
these expressions in Eq. (5), we obtain
±¯
pr
R±
f
a=Bo
2R±
f
a2
1+c±.(6)
Setting R±
f=ain Eq. (6), we recover the formulation proposed by Jensen et al. [11]forprin
the limit of vanishing Ca and R±
fa. We can evaluate Eq. (5) at the tip of the bubble, where
¯
x¯
r±→∓and ¯
r±=
a, in which
a1, and subtract the equation corresponding to the bottom
from the top to find c+=cas in Ref. [11]. We note that is an intermediate variable, which is
introduced to make the discussion more clear, but its specific value does not enter in the derivation.
We then evaluate Eq. (6) at the top and at the bottom and do the subtraction top minus bottom to
obtain pr,
¯
prR+
f
a+R
f
a=−2+Bo
2R+
f
a2
R
f
a2.(7)
094304-9
ATASI, KHODAPARAST, SCHEID, AND STONE
For Bo 1Eq.(7) can be approximated as
¯
pr≈− 2a
R+
f+R
f
.(8)
Note that ¯
prdoes not depend on the axial coordinate, ¯
x, as we assumed before integrating Eq. (4).
Having the expression of ¯
pr, we substitute it in Eq. (4) and evaluate this equation in the matching
region, i.e., for |¯
x¯
r±|1, to find
¯
x¯
x¯
r±=∓ 2a
R+
f+R
f
BoR±
f
a.(9)
Note that the choice of coordinate system gives that ¯
x¯
x¯
r+<0 and ¯
x¯
x¯
r>0 (Fig. 5).
2. Transition region
In the transition region, we start by considering the lubrication approximation and the balance
between the pressure gradient and the viscous forces (per volume) along the xdirection. The steps
are standard, as first developed by Bretherton [1]. In dimensionless variables, we have
¯
p
˜
x±=Ca 2¯
u
˜
y±2,(10)
where ¯
u=u
Ub, with Ubthe bubble velocity; ˜
y±=∓
¯
y+1; and ˜
x±=¯
x+¯
s±, with sbeing a shift in
the axial coordinate, as defined in Fig. 5, and ¯
s=s
a.WeintegrateEq.(10) twice with ¯
u=−1at ˜
y±=0
and the no-shear-stress condition at the interface, i.e., ¯
u
˜
y±=0aty=h±, where h±=1r±.The
pressure gradient along ˜
x±is given by the curvature gradient ¯
p
˜
x±=−˜
x±˜
x±˜
x±˜
h±. Integrating the
velocity field along ˜
y±yields a third-order differential equation, where the variables have been
rescaled as ¯
¯
h±=˜
h±
Ca2/3and ¯
¯
x±=˜
x±
Ca1/3,
¯
¯
h±3d3¯
¯
h±
d¯
¯
x±3=3( ¯
¯
h±¯
¯
h±
),(11)
where ¯
¯
h±
=h±
aCa2/3. We obtain the classical form by defining H=¯
¯
h±
¯
¯
h±
and X=¯
¯
x±
¯
¯
h±
,
H3d3H
dX3=3(H1).(12)
We now express the second derivative xxr±, with the variables from the transition region:
xxr±=∓ 1
a¯
¯
h±
XXH. (13)
The curvature expressed with the variables from the transition region and from the CS region should
match in a common overlap region, in which one can rewrite Eq. (9)intheform
1
¯
¯
h±
lim
X→∞ XXH=∓ 2a
R+
f+R
f
BoR±
f
a,(14)
where a numerical solution of Eq. (12) shows that limX→∞ XXH=1.337 [1], which finally leads
to
h±
a=1.337 Ca2/3
2a
R+
f+R
f
±BoR±
f
a
.(15)
094304-10
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
This result generalizes the development of Jensen et al. [11] to the case of moderate values of Ca
and the development of Klaseboer et al. [17] to the case of finite Bo. From Eq. (15)weareableto
recover the three limiting cases presented previously in the literature.
(i) At very low capillary numbers Ca 0 and consequently very thin lubricating liquid films,
the nose of the bubble approximately fits the entire gap between the top and bottom planes (or the
tube diameter), thus R±
fa. In this limit and for Bo >0, we recover
h±
a=1.337 Ca2/3
1±Bo ,(16)
which was originally obtained by Jensen et al. [11].
(ii) For negligible buoyancy effects Bo 0, no asymmetry is present at the nose of the bubble,
thus R±
f=Rf. In this limit, if we consider moderate Ca the radius of the spherical cap of the bubble
no longer fits the radius of the tube, i.e., a= Rf, and Eq. (15) is reduced to the form presented by
Aussillous and Quéré [4] and discussed further by Klaseboer et al. [17]:
h±
Rf
=1.337 Ca2/3.(17)
(iii) Finally, for Ca 0, Bo 0, and R±
fa, we recover the expression of Bretherton [1]:
h±
a=1.337 Ca2/3.(18)
The distance between the cell or tube centerline and the interface in the matching region R±
fand
the liquid thickness CNh±
should fit the radius of the tube. This gives an additional equation as
proposed by [17]
a=CNh±
+R±
f,(19)
where CNis a constant and h±
is given by Eq. (15). Notice that combining Eqs. (15) and (19)inthe
limiting case when Bo =0 and R+
f=R
f, we recover an equation similar to that in Refs. [4,17],
h±
a=1.337 Ca2/3
1+1.337CNCa2/3,(20)
where CN=2.5 by fitting Eq. (20) to the experimental measurements of Taylor [3,4] and CN=2.79
following a theoretical analysis proposed by Klaseboer et al. [17]. However, Klaseboer et al. [17]
pointed out that the exact value of CNdepends on where Rf±is evaluated in the matching region and
can take a value up to 2.9 if Rf±is evaluated for H>106, i.e., X>10 000, and 2.34 if it is evaluated
for H=6×102, i.e., X=50. We choose CN=2.5 in this work since it has been shown to fit the
experimental measurements. It should be noted that as the film gets thicker with increasing Ca, for a
given value of the tube radius a,CNcould decrease since the liquid plus the bubble should fit in the
tube radius. However, experimental results of Taylor [3,4] show no such effect for Ca up to 2. On
the other hand, the influence of Bo on CNneeds further investigation. Equations (15) and (19)form
a system of four equations with four unknowns R+
f,R
f,h+
, and h
, which is solved numerically
using the NSolve function of Wolfram Mathematica. Note that much above a critical Bond number,
i.e., Bo Boc, the pressure in the bottom film becomes hydrostatic, i.e., ρgRfl
(h
)2, where lis
the length of the transition region and scales with lCa1/3, while h
Ca2/3, thus Boc1
1Ca2/3.
The critical value for the Bond number is Boc=1 for vanishing Ca, i.e., Rf±a, however, for
moderate Ca, i.e., Rf±= a,thevalueofBo
cincreases with Ca. Equation (15)isvalidforBo<Boc.
094304-11
ATASI, KHODAPARAST, SCHEID, AND STONE
B. Inclination angle in the thin-film region
Leung et al. [10] directly visualized the drainage flow from the top to the bottom of the channel in
the thin film region. Using a mass balance analysis in the central region of the bubble and assuming
the thin film limit, they demonstrated that the drainage was causing the thin film to be inclined at an
angle. Their main point was that, in the reference frame translating with the bubble velocity, liquid
flows in the film. The volume of liquid in the film that is drained from the top to the bottom of the
channel increases as one proceeds towards the back of the bubble, leading to less liquid flowing in
the film as the back of the bubble is reached at the top of the channel; this description rationalizes
the thinning of the film at the top. From their mass balance [Eq. (14) from Ref. [10]] we can obtain
tan α±=∓
hθ=π
23
3Ca Bo,(21)
where h(θ=π
2) is the film thickness at the center plane of the channel and is approximately given
by h(θ=π
2)h+
+h
2. It should be noted that possible deformation of bubble in the cross-sectional
plane is assumed to be negligible in order to obtain Eq. (21).
IV. EXPERIMENTAL RESULTS AND DISCUSSION
In this section we present results of our experimental measurements as a function of capillary Ca
and Bond Bo numbers for the range of 5×104<Ca <0.12 and 0.05 <Bo <0.42. The largest
value of the Reynolds number in our experiments was Re =0.023, therefore, inertial effects are
negligible. We distinguish features of the thin film at the top and bottom of the front of the bubble
as well as quantifying the inclination of the bubble due to buoyancy effects. Moreover, we present
measurements of the film thickness at the back of the bubbles, which are different from those
obtained at the front due to the inclination of the bubble. Our main goal in this section is to report
our experimental results, organized systematically using Ca and Bo, and to compare with the theory
presented in the preceding section.
A. Thickness of the liquid film at the front of the bubble
Experimental measurements of the film thickness at the front of the bubble are performed at
both the top and bottom of the tube. Results of these measurements are presented in Fig. 6for four
different values of the Bond number. Note that these experimental data points may correspond to
bubbles of different lengths. Nevertheless, the dispersion of the experimental points are within the
experimental error and the film thicknesses at the front of the bubble h±
front do not depend significantly
on the length of the bubble. Consistent with the theoretical prediction, the film thickness at the front
of the bubble is thicker at the bottom than the top, while the difference between the values of the top
and bottom film thicknesses at the front become larger as the Bo increases.
The data in Fig. 6are compared with two models presented in Sec. III, one being the extended
Bretherton without buoyancy effect as presented in Eq. (20)[4] and a second model derived here,
which was inspired by Jensen et al. accounting for buoyancy effects. The extended Bretherton
formulation (solid line in Fig. 6) always lies in between the experimental measurements at the top
and bottom of the tube. The theoretical prediction obtained by solving the system of equations (15)
and (19) (dotted lines in Fig. 6) follows the experimental measurements well. However, we note that
an underprediction of the film thickness at the bottom and an overprediction of that at the top are
present for all cases, which can be due to non-negligible lateral drainage from the top to the bottom
of the channel in the transition region.
A scaling argument supporting this idea is the following. The time taken for liquid to be convected
in the transition region is tcl
Ub, where lis the length of the transition region. The drainage time
of the liquid from top to bottom is tdπa
Ud, where Udis the drainage velocity. One can estimate the
094304-12
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
top
bottom
Eq. (20)
Eq. (15) & Eq. (19)
hfront
0
0.05
0.10
0.15
0.20
Ca
0 0.03 0.06 0.09 0.12
top
bottom
Eq. (20)
Eq. (15) & Eq. (19)
hfront
0
0.05
0.10
0.15
0.20
Ca
0 0.03 0.06 0.09 0.12
top
bottom
Eq. (20)
Eq. (15) & Eq. (19)
hfront
0
0.05
0.10
0.15
0.20
Ca
0 0.03 0.06 0.09 0.12
top
bottom
Eq. (20)
Eq. (15) & Eq. (19)
hfront
0
0.05
0.10
0.15
0.20
Ca
0 0.03 0.06 0.09 0.12
(a) (b)
(c) (d)
FIG. 6. Top (open symbols) and bottom (closed symbols) film thickness measurements at the front of the
bubble for different Bond numbers: (a) Bo =0.056, (b) Bo =0.11, (c) Bo =0.25, and (d) Bo =0.42. The thin
black solid line is the prediction given by Eq. (20) with the empirical constant CN=2.5. The dotted lines are the
predictions obtained by solving the system of equations (15)and(19) numerically. The results computed with
Eq. (15) are in good agreement with the experiments. As Bo increases the agreement becomes less satisfactory
due to the non-negligible drainage of liquid in the transition region. For the smallest Bo, the dashed and dotted
lines converge towards the solid line.
drainage velocity by balancing gravity and viscous shear along the vertical direction Udμ
h(θπ
2)2ρg,
i.e., Udρgh(θ=π
2)2
μ. Using the scales lCa1/3and h(θ=π
2)Ca2/3, we obtain the ratio
between the two time scales td
tc1
Bo Ca2/3. For sufficiently small values of Bo and Ca, the ratio of
time scales is large enough to ensure no influence of the lateral drainage, but for the upper bounds
of these parameters in our experiments the ratio decreases towards 3, meaning that drainage in the
transition region begins to be significant. It should be noted that the larger difference between the
experiments and the proposed model for the film thickness at bottom compared to that at the top,
especially at higher Bond numbers (Bo =0.25 and 0.42), may also be a consequence of enhanced
drainage for thicker films.
094304-13
ATASI, KHODAPARAST, SCHEID, AND STONE
Bo = 0.056
Bo = 0.11
Bo = 0.25
Bo = 0.42
+ [10-3 rad]
0
2
4
6
8
Ca
0 0.03 0.06 0.09 0.12
- [10-3 rad]
0
2
4
6
8
Ca
0 0.03 0.06 0.09 0.12
(a) (b)
FIG. 7. Inclination angle of the bubble at the (a) top and (b) bottom as a function of capillary number Ca
for different Bond numbers Bo. The error bars (visible when larger than the symbol size) correspond to the
standard deviation of the measurements obtained in three independent experimental tests.
B. Inclination of the bubble
Measurements of the inclination angle αof the bubble at the top α+and bottom αof the tube are
presented in Fig. 7. The inclination angle increases at higher Bo due to the intensified drainage from
the top to the bottom of the tube. We observe that the inclination angle at the top of the bubble is
smaller than that at bottom of the tube. This feature is believed to be due to the slight deformation of
the circular cross section of the bubble due to the buoyancy, which is not included in the theoretical
analysis provided here. Such deformations due to gravity have been reported by de Lozar et al. [7].
Since the scaling analysis discussed in Eq. (21) is performed considering a circular cross section all
along the bubble, we compare the theoretical prediction with the average inclination angle calculated
for the top and bottom αave =α++α
2. The result of this comparison is presented in Fig. 8, where
excellent agreement is obtained between the experimental measurements and the prediction of the
mass balance analysis (21). Note that according to both theory and the present experiments, buoyancy
effects are observable even at the lowest Bond number Bo =0.056. Moreover, the dependence of
the inclination angle on the capillary number diminishes at higher Ca. Figure 8shows that the mass
balance argument provided by Leung et al. [10] can provide a reliable prediction method for the
average inclination angle of the bubble in the presence of non-negligible buoyancy.
C. Liquid film at the back of the bubble
Unlike the film thickness at the front of the bubble, the film thickness varies with the length of
the bubble at the back due to the drainage of the liquid film and the consequent inclination of the
bubble. This effect is observed in our experimental measurements of the film thickness at the back
of the bubble h±
back. Figure 9presents the ratio of the film thickness at the back of the bubble to
that at the front ¯
h±
back/¯
h±
front versus the dimensionless length of the bubble ¯
L=L/a obtained in the
experiments. Three main features are clearly observed in Fig. 9. (i) The ratio ¯
h±
back/¯
h±
front is linearly
correlated to the dimensionless length of the bubble ¯
L=L/a. (ii) The magnitude of the slopes of
the best linear fits to experimental data are slightly larger at the bottom compared to the top of the
tube. This finding once again shows that the cross section of the bubble loses its circular shape
towards the back of the bubble. (iii) For the limit of ¯
L0 all the linear trends at both the top and
the bottom of the tube reach unity (Fig. 9).
094304-14
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
Bo = 0.056
Bo = 0.11
Bo = 0.25
Bo = 0.42
αave [10-3 rad]
0
2
4
6
Ca
0 0.03 0.06 0.09 0.12
FIG. 8. Mean inclination angle of bubbles computed by averaging the measurements at the top and the
bottom αave =(α++α)/2. Solid lines present the prediction of Eq. (21).
At low Ca and finite Bo, the film thickness at the top of the tube and at the nose of the bubble
is less than 10 μm. As this thickness decreases towards the back of the bubble, it can eventually
reach values smaller than 1 μm for long bubbles of L10 mm. Thin films of partially wetting
liquids with thicknesses below 1 μm are known to be metastable and thus may dewet the wall of the
channel [1820]. This effect causes dewetting on the top wall of the channel, which was observed
in our experiments at low capillary numbers Ca <103. Therefore, not only the mean liquid film
Bo = 0.056
Bo = 0.11
Bo = 0.25
Bo = 0.42
h±
back / h±
front
0
0.5
1.0
1.5
2.0
L
0 10203040
FIG. 9. Ratio of the back to front film thicknesses ¯
h±
back/¯
h±
front obtained in the experiments versus the
dimensionless length of the bubble ¯
L=L/a. Open and closed symbols correspond to the measurements at the
top and bottom of the channel, respectively. Solid lines represent the best linear fit to the experimental data.
Here ¯
h±
back/¯
h±
front shows a linear trend with ¯
Land approaches unity in the limit ¯
L0.
094304-15
ATASI, KHODAPARAST, SCHEID, AND STONE
thickness but also the minimum film thickness caused by the effect of buoyancy at the top rear of the
bubble, which depends on the bubble length, must be considered in studies where the film thickness
has a critical effect on the final results, e.g., critical heat transfer determination for a heated pipe.
V. CONCLUSION
In this study we investigated the effect of buoyancy on the thickness of the liquid film formed
around elongated confined bubbles, which translate in a horizontal tube of circular cross section.
We performed systematic experiments, in which the dimensionless parameters, namely, capillary Ca
and Bond Bo numbers, were studied. In general, buoyancy effects were observed even at very low
Bond numbers Bo =0.056 and were amplified at higher Bo.
At finite Bo, the film thickness at the front of the bubble is thinner at the top than at the bottom.
We extended the theory originally proposed by Jensen et al. in a two-dimensional configuration
for low capillary numbers Ca <103and finite Bo [11], to account for the film thickening effect
observed at moderate capillary numbers [4,17]. The resulting theoretical correlation quantitatively
predicts the present experimental measurements with less than 12% relative error. In general, the
theory underpredicts the effect of buoyancy on the film thickness especially at higher Ca and Bo,
which may be a consequence of the lateral drainage in the transition region.
Due to buoyancy, a lateral drainage of liquid exists from the top to the bottom, which leads to
the thinning of the liquid film towards the back of the bubble at the top. Therefore, the body of
the bubble is inclined relative the centerline of the tube so that the back of the bubble is elevated.
We quantified this inclination angle at the top and the bottom of the bubble. While the angle of
inclination remains constant along the bubble, it is slightly larger at the bottom of the bubble. We
believe that this effect is a consequence of the deformation of the bubble in the cross-sectional plane
as compared to a circular shape. If the cross-sectional deformation of the bubble is neglected, a mass
balance analysis for the liquid around the bubble in the central region by Leung et al. provides a
theoretical prediction for the inclination angle [10]. We showed that this prediction can well predict
the mean value of our experimental measurements of the inclination angles at the top and bottom
of the bubble. Additionally, our findings suggest that the role of the drainage of the liquid in the
transition region at the nose of the bubble and deformation of the bubble in the cross-sectional plane
is an avenue for future theoretical analysis. Furthermore, future theoretical studies should investigate
how the nondimensional liquid thickness in the matching region CNdepends on Bo.
ACKNOWLEDGMENTS
O.A. and B.S. thank the F.R.S-FNRS for financial support as well as the BELSPO agency under
Grant No. IAP-7/38 MicroMAST. S.K. appreciates the early mobility funding from Swiss National
Science Foundation (Grant No. P2ELP2-158896). We are grateful to the referees for their insightful
comments.
[1] F. P. Bretherton, The motion of long bubbles in tubes, J. Fluid Mech. 10,166 (1961).
[2] F. Fairbrother and A. E. Stubbs, Studies in electro-endosmosis. Part VI. The “bubble-tube” method of
measurement, J. Chem. Soc. 527 (1935).
[3] G. I. Taylor, Deposition of a viscous fluid on the wall of a tube, J. Fluid Mech. 10,161 (1961).
[4] P. Aussillous and D. Quéré, Quick deposition of a fluid on the wall of a tube, Phys. Fluids 12,2367 (2000).
[5] V. Suresh and J. B. Grotberg, The effect of gravity on liquid plug propagation in a two-dimensional
channel, Phys. Fluids 17,031507 (2005).
[6] Y. Zheng, H. Fujioka, and J. B. Grotberg, Effects of gravity, inertia, and surfactant on steady plug
propagation in a two-dimensional channel, Phys. Fluids 19,082107 (2007).
[7] A. de Lózar, A. Juel, and A. L. Hazel, The steady propagation of an air finger into a rectangular tube,
J. Fluid Mech. 614,173 (2008).
094304-16
EFFECT OF BUOYANCY ON THE MOTION OF LONG . . .
[8] Y. Han and N. Shikazono, Measurement of the liquid film thickness in micro tube slug flow, Int. J. Heat
Fluid Flow 30,842 (2009).
[9] R. Gupta, H. Hagnefelt, D. F. Fletcher, and B. S. Haynes, Proceedings of the Seventh International
Conference on Multiphase Flow, Tampa, University of Florida, Gainesville, FL, 2010.
[10] S. Y. Leung, R. Gupta, D. F. Fletcher, and B. S. Haynes, Gravitational effect on Taylor flow in horizontal
microchannels, Chem. Eng. Sci. 69,553 (2012).
[11] M. H. Jensen, A. Libchaber, P. Pelcé, and G. Zocchi, Effect of gravity on the Saffman-Taylor meniscus:
Theory and experiment, Phys. Rev. A 35,2221 (1987).
[12] E. Brener, M. Rabaud, and H. Thomé, Effect of gravity on stable Saffman-Taylor fingers, Phys.Rev.E48,
1066 (1993).
[13] R. Budwig, Refractive index matching methods for liquid flow investigations, Exp. Fluids 17,350 (1994).
[14] S. Khodaparast, N. Borhani, and J. R. Thome, Application of micro particle shadow velocimetry μPSV
to two-phase flows in microchannels, Int. J. Multiphase Flow 62,123 (2014).
[15] N. Borhani, B. Agostini, and J. R. Thome, A novel time strip flow visualisation technique for investigation
of intermittent dewetting and dryout in elongated bubble flow in a microchannel evaporator, Int. J.
Multiphase Flow 53,4809 (2010).
[16] M. D. Giavedoni and F. A. Saita, The rear meniscus of a long bubble steadily displacing a Newtonian
liquid in a capillary tube, Phys. Fluids 11,786 (1999).
[17] E. Klaseboer, R. Gupta, and R. Manica, An extended Bretherton model for long Taylor bubbles at moderate
capillary numbers, Phys. Fluids 26,032107 (2014).
[18] A. Sharma and E. Ruckenstein, Dewetting of solids by the formation of holes in macroscopic liquid films,
J. Colloid Interface Sci. 133,358 (1989).
[19] G. Callegari, A. Calvo, and J. P. Hulin, Dewetting processes in a cylindrical geometry, Eur. Phys. J. E 16,
283 (2005).
[20] A. Huerre, O. Theodoly, A. M. Leshansky, M.-P. Valignat, I. Cantat, and M.-C. Jullien, Droplets in
Microchannels: Dynamical Properties of the Lubrication Film, Phys.Rev.Lett.115,064501 (2015).
094304-17
... However, recent experimental and computational studies of isolated, confined gas bubbles propagating in small channels have revealed that buoyancy may be important even when Bo ≪ 1. Atasi et al. (2021) measured the top and bottom thicknesses of the thin lubricating liquid film surrounding long gas bubbles flowing in horizontal circular capillaries, and observed apparent film thickness asymmetries already for Bo = 2 ∕ ≈ 0.1, with being the channel diameter. Motivated by the work of Atasi et al. (2021), Moran et al. (2021) carried out three-dimensional numerical simulations of slug flow in a horizontal pipe at varying capillary and Bond numbers. ...
... However, recent experimental and computational studies of isolated, confined gas bubbles propagating in small channels have revealed that buoyancy may be important even when Bo ≪ 1. Atasi et al. (2021) measured the top and bottom thicknesses of the thin lubricating liquid film surrounding long gas bubbles flowing in horizontal circular capillaries, and observed apparent film thickness asymmetries already for Bo = 2 ∕ ≈ 0.1, with being the channel diameter. Motivated by the work of Atasi et al. (2021), Moran et al. (2021) carried out three-dimensional numerical simulations of slug flow in a horizontal pipe at varying capillary and Bond numbers. They confirmed that, even when Bo ≪ 1, the gravity force tends to drain liquid from the top film, thus inducing the previously observed film asymmetry, and reported that the effect of buoyancy was more apparent at small capillary numbers. ...
... Indeed, several assumptions made in the theoretical analysis, reasonable in the vicinity of the threshold, are expected to fail in the large-Bo regime. Notably, Atasi et al. (2017) measured the top and bottom film thickness around long bubbles in horizontal tubes and observed that the film asymmetry grows with the Bond number. Thus, in large capillaries, the thin-film thickness cannot be considered as uniform along the azimuthal direction: the lubricating film is indeed much thicker in the direction x > 0 (Zukoski 1966). ...
Article
Full-text available
In confined systems, the entrapment of a gas volume with an equivalent spherical diameter greater than the dimension of the channel can form extended bubbles that obstruct fluid circuits and compromise performance. Notably, in sealed vertical tubes, buoyant long bubbles cannot rise if the inner tube radius is below a critical value near the capillary length. This critical threshold for steady ascent is determined by geometric constraints related to the matching of the upper cap shape with the lubricating film surrounding the elongated part of the bubble. Developing strategies to overcome this threshold and release stuck bubbles is essential for applications involving narrow liquid channels. Effective strategies involve modifying the matching conditions with an external force field to facilitate bubble ascent. However, it is unclear how changes in acceleration conditions affect the motion onset of buoyancy-driven long bubbles. This study investigates the mobility of elongated bubbles in sealed tubes with an inner radius near the critical value inhibiting bubble motion in a vertical setting. Two strategies are explored to tune bubble motion, leveraging variations in axial and transversal accelerations: tube rotation around its axis and tube inclination relative to gravity. By revising the geometrical constraints of the simple vertical setting, the study predicts new thresholds based on rotational speed and tilt angle, respectively, providing forecasts for the bubble rising velocity under modified apparent gravity. Experimental measurements of motion threshold and rising velocity compare well with theoretical developments, thus suggesting practical approaches to control and tune bubble motion in confined environments.
... Subsequent works retained the assumption of inertialess flow, but made technical improvements, such as accounting for the finite thickness of the residual film (Aussillous & Quéré 2000;Klaseboer, Gupta & Manica 2014), or by accounting for higher-order viscous effects, via asymptotic expansions in terms of the capillary number, Ca = μ 1 U 2 /σ , where μ 1 , U 2 and σ denote the liquid dynamic viscosity, the gas superficial velocity and the surface tension, and/or in terms of the long-wave parameter, (Park & Homsy 1984;Jensen et al. 1987). Other works accounted for the effect of gravity (Jensen et al. 1987;Lasseux 1995;Atasi et al. 2017), a contact line at the leading front of the liquid plug (Kalliadasis & Chang 1994;Bico & Quéré 2001) or a flexible tube wall (Howell, Water & Grotberg 2000). ...
Article
We study liquid plugs in the pulmonary airways based on the two-phase axisymmetric weighted residual integral boundary-layer model of Dietze et al. ( J. Fluid Mech. , vol. 894, 2020, A17), which was originally developed to study liquid films coating the inner surface of a cylindrical tube in interaction with a core gas flow. The augmented form of this model, which was never applied beyond a proof of concept, allows for the representation of liquid pseudo-plugs. Here, we demonstrate its predictive power vs experiments and direct numerical simulations, in terms of the dynamics of plug formation and the characteristics of developed liquid plugs, such as their shape, flow field, speed and length, as well as the associated wall stresses and their spatial derivatives. In particular, we show that the augmented model allows us to establish a direct continuation path from travelling-wave solutions (TWS) to travelling-plug solutions (TPS). We then apply the model to predict mucus plugs in the conducting zone of the tracheobronchial tree, based on the lung architecture model of Weibel. We proceed by numerical continuation of travelling-state solutions in terms of the airway generation, whereby we impose the wavelength of the linearly most-amplified convective instability (CI) mode or that of the absolute instability (AI) mode. We identify the critical airway generation for liquid-plug formation (TWS/TPS transition), maximum potential for wall-stress-induced epithelial cell damage and CI/AI transition, and investigate how these phenomena are affected by the main control parameters, i.e. airway orientation vs gravity, air flow rate, mucus properties and airway size.
... 33,34 In addition, radial displacement models of Newtonian and non-Newtonian fluids are numerically evaluated by Lee et al. 35 and Ward and White. 36 Research on immiscible flows has extensively investigated the dynamics of bubble and droplet behavior under various conditions, focusing on factors such as surface tension, [37][38][39] wetting, 40 unsteady flow, 41 buoyancy, 42,43 and bubble dynamics. 44 Studies have also explored shear thinning [45][46][47] and viscoplastic fluids 48,49 as well as wall roughness effects on bubble deformation 50,51 and multiple bubble dynamics, [52][53][54] including necking, snap-off, pinch-off, coalescence, and pearling. ...
Article
Full-text available
We examine immiscible displacement flows in stationary and rotating pipes, at a fixed inclination angle in a density-unstable configuration, using a viscoplastic fluid to displace a less viscous Newtonian fluid. We employ non-intrusive experimental methods, such as camera imaging, planar laser-induced fluorescence (PLIF), and ultrasound Doppler velocimetry (UDV). We analyze the impact of key dimensionless numbers, including the imposed Reynolds numbers (Re, Re*), rotational Reynolds number ( Rer), capillary number (Ca), and viscosity ratio (M), on flow patterns, regime classifications, regime transition boundaries, interfacial instabilities, and displacement efficiency. Our experiments demonstrate distinct immiscible displacement flow patterns in stationary and rotating pipes. In stationary pipes, heavier fluids slump underneath lighter ones, resulting in lift-head and wavy interface stratified flows, driven by gravity. Decreasing M slows the interface evolution and reduces its front velocity, while increasing Re* shortens the thin layer of the interface tail. In rotating pipes, the interplay between viscous, rotational, and capillary forces generates swirling slug flows with stable, elongated, and chaotic sub-regimes. Progressively, decreasing M leads to swirling dispersed droplet flow, swirling fragmented flow, and, eventually, swirling bulk flow. The interface dynamics, such as wave formations and velocity profiles, is influenced by rotational forces and inertial effects, with Fourier analysis showing the dependence of the interfacial front velocity's dominant frequency on Re and Rer. Finally, UDV measurements reveal the existence/absence of countercurrent flows in stationary/rotating pipes, while PLIF results provide further insight into droplet formation and concentration field behavior at the pipe center plane.
... The variation in width h of this film as a function of bubble velocity has been widely studied and various empirical equations have been proposed (Chen et al., 2023), in particular for elongated bubbles (Atasi et al., 2017;Danov et al., 2021;Jones et al., 2021;Moran et al., 2021). Such work dates back to Taylor (Taylor, 1960), who showed that the ratio h/Rc (where Rc is the channel radius) asymptotes to 1/3 for Ca → 2. ...
... An analytical balance (ME104E, Mettler Toledo) was installed at the outlet of the glass capillary to confirm the flow rate. In the current experiment, Bo = ρgR 2 /σ < 0.04 and Re = ρUR/μ 1, so gravity and inertia effects are negligible (Atasi et al. 2017;Magnini et al. 2019). ...
Article
Full-text available
The motion of a long gas bubble in a confined capillary tube is ubiquitous in a wide range of engineering and biological applications. While the understanding of the deposited thin viscous film near the tube wall in Newtonian fluids is well developed, the deposition dynamics in commonly encountered non-Newtonian fluids remains much less studied. Here, we investigate the dynamics of a confined bubble moving in shear-thinning fluids with systematic experiments, varying the zero-shear-rate capillary number Ca0Ca_0 in the range of O(10^{-3}\unicode{x2013}10^2) considering the zero-shear-rate viscosity. The thickness of the deposited liquid film, the bubble speed and the bubble front/rear menisci are measured, which are further rationalized with the recent theoretical studies based on appropriate rheological models. Compared with Newtonian fluids, the film thickness decreases for both the carboxymethyl cellulose and Carbopol solutions when the shear-thinning effect dominates. We show that the film thickness follows the scaling law from Aussillous & Quéré ( Phys. Fluids , vol. 12, no. 10, 2000, pp. 2367–2371) with an effective capillary number CaeCa_e , considering the characteristic shear rate in the film as proposed by Picchi et al. ( J. Fluid Mech. , vol. 918, no. A7, 2021, pp. 1–30). CaeCa_e is calculated by the Carreau number and the power-law index from the Carreau–Yasuda rheological model. The shear-thinning effect also influences the bubble speed and delays the transition to the parabolic region in the bubble front and rear menisci. In particular, a high degree of undulations on the bubble surface results in an intricate rear viscosity distribution for the rear meniscus and the deviation between the experiments and theory may require a further investigation to resolve the axial velocity field. Our study may advance the fundamental understandings and engineering guidelines for coating processes involving thin-film flows and non-Newtonian fluids.
... An analytical balance (ME104E, Mettler Toledo) was installed at the outlet of the glass capillary to confirm the flow rate. In the current experiment, Bo = ρgR 2 /σ < 0.04 and Re = ρU R/µ ≪ 1 so gravity and inertia effects are negligible (Atasi et al. 2017;Magnini et al. 2019). ...
Preprint
Full-text available
The motion of a long gas bubble in a confined capillary tube is ubiquitous in a wide range of engineering and biological applications. While the understanding of the deposited thin viscous film near the tube wall in Newtonian fluids is well developed, the deposition dynamics in commonly encountered non-Newtonian fluids remains much less studied. Here, we investigate the dynamics of a confined bubble moving in shear-thinning fluids with systematic experiments, varying the zero-shear-rate capillary number Ca0Ca_0 in the range of O(103102)O(10^{-3}-10^2) considering the zero-shear-rate viscosity. The thickness of the deposited liquid film, the bubble speed and the bubble front/rear menisci are measured, which are further rationalized with the recent theoretical studies based on appropriate rheological models. Compared with Newtonian fluids, the film thickness decreases for both the carboxymethyl cellulose and Carbopol solutions when the shear-thinning effect dominates. We show that the film thickness follows the scaling law from \citet{aussillous2000quick} with an effective capillary number CaeCa_e, considering the characteristic shear rate in the film as proposed by \citet{picchi2021motion}. CaeCa_e is calculated by the Carreau number and the power-law index from the Carreau-Yasuda rheological model. The shear-thinning effect also influences the bubble speed and delays the transition to the parabolic region in the bubble front and rear menisci. In particular, a high degree of undulations on the bubble surface results in intricate rear viscosity distribution for the rear meniscus and the deviation between the experiments and theory may require a further investigation to resolve the axial velocity field. Our study may advance the fundamental understandings and engineering guidelines for coating processes involving thin-film flows and non-Newtonian fluids.
... The seminal work of Bretherton (1961) revealed that bubble characteristics (i.e. the film thickness, the bubble speed) can be expressed as a function of the capillary number and that the film profile can be described with a similarity solution typical of Landau-Levich-Derjaguin-Bretherton problems (see de Gennes, Brochard-Wyart & Quéré 2003;Stone 2010). The theory is supported by several experimental measurements (Fairbrother & Stubbs 1935;Taylor 1961;Schwartz, Princen & Kiss 1986;Aussillous & Quéré 2000) and, in the last decades, it has been extended to include the effect of viscous forces and weak inertia (Cox 1962;Reinelt & Saffman 1985;Aussillous & Quéré 2000;Heil 2001;de Ryck 2002;Khodaparast et al. 2015;Magnini et al. 2017), unsteady flow, (Yu et al. 2018), buoyancy, (Leung et al. 2012;Atasi et al. ...
Article
Full-text available
The motion of Taylor bubbles in capillaries is typical of many engineering and biological systems, ranging from subsurface flows to small-scale reactors. Although the hydrodynamics of elongated bubbles has been the object of several studies, the case where a solute is transported in the surrounding liquid and surface mass-transfer mechanisms act on the solid wall or the bubble–fluid interface is much less understood. To fill this gap, we investigate the transport problem around a confined Taylor bubble to access the competition between advection, diffusion and surface mass transfer in the different regions of the bubble. With this aim, we derive a one-dimensional advection–diffusion–mass-transfer equation where the transport mechanisms are described through an effective velocity, an effective diffusion coefficient and an effective Sherwood number. Our model generalises the Aris–Taylor dispersion to the case of a Taylor bubble and clarifies the impact of surface mass transfer in the advection and diffusion dominated regimes for both the front and rear menisci. The model recovers the typical Pèclet square relationship of the effective diffusion coefficient, which also depends on the film thickness. Also, when the Pèclet number balances with the Sherwood number, there exist conditions that lead to the formation of hot spots of concentration. We show that the typical shape oscillations of the bubble rear locally enhance superficial mass transfer. Finally, we study the transport problem in the uniform film region where the concentration field can be found analytically.
Article
Full-text available
Thin‐film deposition of fluids is ubiquitous in a wide range of engineering and biological applications, such as surface coating, polymer processing, and biomedical device fabrication. While the thin viscous film deposition in Newtonian fluids has been extensively investigated, the deposition dynamics in frequently encountered non‐Newtonian complex fluids remain elusive, with respect to predictive scaling laws for the film thickness. Here, we investigate the deposition of thin films of shear‐thinning viscoelastic fluids by the motion of a long bubble translating in a circular capillary tube. Considering the weakly elastic regime with a shear‐thinning viscosity, we provide a quantitative measurement of the film thickness with systematic experiments. We further harness the recently developed hydrodynamic lubrication theory to quantitatively rationalize our experimental observations considering the effective capillary number CaeCaeCa_\mathrm{e} and the effective Weissenberg number WieWieWi_\mathrm{e}, which describe the shear‐thinning and the viscoelastic effects on the film formation, respectively. The obtained scaling law agrees reasonably well with the experimentally measured film thickness for all test fluids. Our work may potentially advance the fundamental understanding of the thin‐film deposition in a confined geometry and provide valuable engineering guidance for processes that incorporate thin‐film flows and non‐Newtonian fluids.
Article
Full-text available
The motion of a gaseous Taylor bubble in a capillary tube is typical of many biological and engineering systems, such as small-scale reactors and microfluidic devices. Although the dynamics of a bubble in a Newtonian liquid has been the subject of several studies since the seminal works of Taylor (J. case where the fluid exhibits a shear-thinning behaviour is much less understood. To fill this gap, we study the dynamics of a bubble that moves in a shear-thinning fluid whose viscosity is described by the Ellis viscosity model. With this aim, we derive a lubrication model in the film region to identify the scaling laws for the bubble speed, the film thickness and the pressure drop as a function of the Ellis number and the degree of shear thinning. Our model generalizes Bretherton's results to shear-thinning fluids by identification of a universal scaling law for the effective viscosity that accounts for the interplay of the zero-shear-rate and shear-thinning effects. The film thickness follows a 2/3 scaling law with respect to the capillary number based on the proposed effective viscosity. The ratio between the bubble speed and the average velocity of the fluid ahead of the bubble is a function of the effective capillary number only. We show that some portions of the bubble are dominated by the zero-shear-rate effect discussing the extent to which the use of the power-law viscosity model can be legitimized. Finally, we study the location of the recirculating vortices ahead of the bubble.
Article
Full-text available
We study the motion of droplets in a confined, micrometric geometry, by focusing on the lubrication film between a droplet and a wall. When capillary forces dominate, the lubrication film thickness evolves nonlinearly with the capillary number due to the viscous dissipation between the meniscus and the wall. However, this film may become thin enough (tens of nanometers) that intermolecular forces come into play and affect classical scalings. Our experiments yield highly resolved topographies of the shape of the interface and allow us to bring new insights into droplet dynamics in microfluidics. We report the novel characterization of two dynamical regimes as the capillary number increases: (i) at low capillary numbers, the film thickness is constant and set by the disjoining pressure, while (ii) above a critical capillary number, the interface behavior is well described by a viscous scenario. At a high surfactant concentration, structural effects lead to the formation of patterns on the interface, which can be used to trace the interface velocity, that yield direct confirmation of the boundary condition in the viscous regime.
Article
Full-text available
We are interested in the amount of fluid left behind a drop moved inside a capillary tube. Long ago, Taylor showed that for very viscous liquids moved at small velocities, the film thickness is a monotonic increasing function of the capillary number. New data obtained with liquids of low viscosity are reported here and compared with Taylor's law. Two successive effects are observed: above a threshold in capillary number, the film is thicker than a Taylor film; at a very high speed, the deposition law becomes a decreasing function of the drop velocity. Both behaviors are analyzed thanks to scaling arguments and shown to be consequences of inertia.
Article
Full-text available
The effect of plug propagation speed and gravity on the quasisteady motion of a liquid plug in a two-dimensional liquid-lined channel oriented at an angle α with respect to gravity is studied. The problem is motivated by the transport of liquid plugs instilled into pulmonary airways in medical treatments such as surfactant replacement therapy, drug delivery, and liquid ventilation. The capillary number Ca is assumed to be small, while the Bond number Bo is arbitrary. Using matched asymptotic expansions and lubrication theory, expressions are obtained for the thickness of the trailing films left behind by the plug and the pressure drop across it as functions of Ca, Bo, α and the thickness of the precursor films. When the Bond number is small it is found that the trailing film thickness and the flow contribution to the pressure drop scale as Ca2/3 at leading order with coefficients that depend on Bo and α. The first correction to the film thickness is found to occur at O(Ca) compared to O(Ca4/3) in the Bo=0 case. Asymmetry in the liquid distribution is quantified by calculating the ratio of liquid volumes above and below the centerline of the channel, VR;. VR=1 at Bo=0, indicating a symmetric distribution, and decreases with Bo and Ca, but increases with the plug length Lp. The decrease of VR with Ca suggests that higher propagation speeds in small airways may result in less homogenous liquid distribution, which is in contrast to the expected effect in large airways. For given values of the other parameters, a maximum capillary number Cac is identified above which the plug will eventually rupture. When the Bond number becomes equal to an orientation-dependent critical value Boc, it is found that the scaling of the film thickness and pressure drop change to Ca1/2 and Ca1/6, respectively. It is shown that this scaling is valid for small increments of the Bond number over its critical value, Bo=Boc+BCa1/6, but for higher Bond numbers the asymptotic approach breaks down.
Article
When (long) bubbles are transported in tubes containing a fluid, the presence of a thin film of fluid along the tube walls causes the velocity of the bubble to be different from the average fluid velocity. Bretherton ["The motion of long bubbles in tubes," J. Fluid Mech. 10, 166 (1961)] derived a model to describe this phenomenon for pressure driven flows based on a lubrication approach coupled with surface deformation of the bubble. Bretherton found that the parameter governing the physics involved is the capillary number (Ca) which expresses the relationship between speed of the bubble, surface tension, and viscosity of the liquid. The results of Bretherton are here re-derived and analyzed in a slightly more perspicuous manner. Incorporating the condition that the bubble-film combination should fit inside the tube results in an expression very similar to the one found empirically by Aussillous and Quéré ["Quick deposition of a fluid on the wall of a tube," Phys. Fluids 12, 2367 (2000)] of the Taylor ["Deposition of a viscous fluid on the wall of a tube," J. Fluid Mech. 10, 161 (1961)] experimental data. Our expression is valid for Ca values up to Ca = 2.0, but approaches Bretherton's result for low values of Ca. The analysis is done in terms of the pressure buildup which originates from the interplay between surface tension and lubrication due to the thin layer of fluid near the tube wall.
Article
Micro particle shadow velocimetry (μμPSV) is performed in the present study for simultaneous velocity measurement and interface tracking in both liquid-liquid and gas-liquid two-phase flows through circular microchannels of 500 μμm diameter. The back-lit illumination using a non-coherent LED light source, combined with full refractive index matching of the liquid phases, the tube wall material and the channel exterior medium, allowed velocimetry to be done both within and around the liquid droplets, and even close to the interfaces and boundaries. Moreover, post-processing methods are proposed and implemented in order to resolve motion of isolated gas bubbles and immiscible liquid droplets in laminar flows quantitatively. In particular, simultaneous measurements of local instantaneous phase velocities and flow rates, liquid film dynamics and its thickness, shape and volume of the dispersed phase, and development length in front and at the back of the bubbles are obtained using one single sequence of grey scale shadowgraphy images. Such results are valuable for validation of corresponding numerical simulation codes. It is believed that this approach significantly reduces the size and cost of the experimental setup while increasing the accuracy and reliability of the measurements.
Article
In this work the interfacial shapes and the flow occurring at the trailing meniscus of a long bubble is numerically analyzed. The technique employed solves the complete set of governing equations simultaneously. The numerical results reported complete previous descriptions of the creeping flow regime; the influence of the inertia forces on the free surface shapes, interfacial undulations, and flow patterns is also analyzed.
Article
Liquid plugs may form in pulmonary airways during the process of liquid instillation or removal in many clinical treatments. Studies have shown that the effectiveness of these treatments may depend on how liquids distribute in the lung. Better understanding of the fundamental fluid mechanics of liquid plug transport will facilitate treatment strategies. In this paper, we develop a numerical model of steady plug propagation driven by gravity and pressure in a two-dimensional liquid-lined channel oriented at an angle alpha with respect to gravity. We investigate the effects of gravity through the Bond number, Bo, and alpha the plug propagation speed through the capillary number, Ca, or the Reynolds number, Re; the plug length LP, and the surfactant concentration C0. Without gravity, i.e., Bo=0, the plug is symmetric, and there are two regimes for the flow: two wall layers and two trapped vortices in the core. There is no flow interaction between the upper and lower half plug domains. When Bo!=0 and alpha!=0, pi, fluid is found to flow from the upper precursor film, through the core and into the lower trailing film. Then the number of vortices can be zero, one, or two, depending on the flow parameters. The vortices have stagnation points on the interface when C0=0, however when the surfactant is present (C0>0), the vortices detach from the interface and create saddle points inside the core. The front meniscus develops a capillary surface wave extending into the precursor film. This is where the film is thinnest and thus the wall shear stress is highest, as high as ~100 dyn/cm2 in adult airways, which indicates a significant risk of pulmonary airway epithelial cell damage. Adding surfactant can decrease the peak magnitude of the shear stress, thus reducing the risk of cell damage. The prebifurcation asymmetry of the plug is quantified by the volume ratio, Vr, defined as the ratio of the liquid above to that below the center line of the channel. Vr is found to increase with LP, Ca, Re, and C0, while it decreases with Bo. The total mass left behind in the trailing films increases with Bo for any alpha at alpha>2pi/5, Ca and alpha for any value of Bo>0.
Article
The effect of the gravitational force has generally been ignored in studies of the hydrodynamics, mass and heat transfer characteristics of Taylor flow in microchannels. In this study, the effects of gravity on the bubble behavior and heat transfer rates for horizontal Taylor flow in millimeter-sized channels were studied experimentally. Significant bubble asymmetry was observed for nitrogen-ethylene glycol Taylor flow in three circular channels of different internal diameters (d=1.12, 1.69, 2.12 mm giving Bond numbers (Bo) of 0.287, 0.653 and 1.028). A parabolic drainage velocity profile inside the liquid film surrounding the bubble was noted through performing Micro-Particle Image Velocimetry (μPIV) measurements in that region. A simplified model similar to Nusselt's solution for laminar film condensation on a horizontal tube was derived to describe the drainage velocity profile in the liquid film at a cross-section away from the bubble ends and was able to explain the observed tilt of the bubble. Performing a non-dimensional analysis of the model, the ratio of the average drainage to bubble velocities was found to be a function of the Bond and Capillary (Ca) numbers. The gravitational effect on the bubble behavior vanishes for small value of Bo and/or Ca or as Ca→∞Ca→∞. Heat transfer experiments in a 2.00 mm tube (Bo=0.915) showed that gravitational effects had no significant effect on heat transfer in the Taylor flow regime for the conditions studied.