Sei sulla pagina 1di 21

On the bubble rise velocity of a

continually released bubble chain in


still water and with crossflow
Cite as: Phys. Fluids 27, 103301 (2015); https://doi.org/10.1063/1.4932176
Submitted: 04 May 2015 . Accepted: 14 September 2015 . Published Online: 06 October 2015

Binbin Wang , and Scott A. Socolofsky

ARTICLES YOU MAY BE INTERESTED IN

Experimental studies on the shape and path of small air bubbles rising in clean
water
Physics of Fluids 14, L49 (2002); https://doi.org/10.1063/1.1485767

Structure and dynamics of the wake of bubbles and its relevance for bubble
interaction
Physics of Fluids 11, 1781 (1999); https://doi.org/10.1063/1.870043

Numerical investigation of rising bubble wake and shape variations


Physics of Fluids 21, 122102 (2009); https://doi.org/10.1063/1.3271146

Phys. Fluids 27, 103301 (2015); https://doi.org/10.1063/1.4932176 27, 103301

© 2015 AIP Publishing LLC.


PHYSICS OF FLUIDS 27, 103301 (2015)

On the bubble rise velocity of a continually released bubble


chain in still water and with crossflow
Binbin Wanga) and Scott A. Socolofsky
Department of Civil Engineering, Texas A&M University, College Station,
Texas 77843-3136, USA
(Received 4 May 2015; accepted 14 September 2015; published online 6 October 2015)

The rise velocities of in-chain bubbles continually released from a single orifice
in still water with and without crossflow are investigated in a series of laboratory
experiments for wobbling ellipsoidal bubbles with moderate Reynolds number. For
the limiting case in still water, that is, crossflow velocity = 0, the theoretical turbulent
wake model correctly predicts the in-chain bubble rise velocity. In this case, the bub-
ble rise velocities VB are enhanced compared to the terminal velocities of the isolated
bubbles V0 due to wake drafting and are scaled with flow rate Q and bubble diameter
D. Here, we also derive an updated wake model with consideration of the superposi-
tion of multiple upstream bubble wakes, which removes the nonlinear behavior of the
non-distant (i.e., local) wake model. For the cases with crossflow, the enhancement
of the in-chain bubble rise velocity can be significantly reduced, and imaging of the
experiments shows very organized paring and grouping trajectories of rising bubbles
not observed in still water under different crossflow velocities. The in-chain bubble
rise velocities in crossflow are described by two models. First, an empirical model is
used to correct the still-water equation for the crossflow effect. In addition, a semi-
theoretical model considering the turbulent wake flow and the crossflow influence
is derived and used to develop a theoretical normalization of bubble rise velocity,
crossflow velocity, and the released bubble flow rate. The theoretical model suggests
there are two different regimes of bubble-bubble interaction, with strong interaction
occurring for the non-dimensional crossflow velocity Uc+ = πUc3 D 3V0/(18g βQ2) less
than 0.06 and weaker interaction occurring for Uc+ greater than 0.06, where Uc is
the crossflow velocity, g is the acceleration of gravity, and β is the mixing length
coefficient. C 2015 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4932176]

I. INTRODUCTION
Gas bubbles rising through water occur in many environmental flows. Natural examples include
aerated flows due to breaking waves or white-capping, release of oxygen bubbles by aquatic
plants, seeps from natural gas reservoirs, and sediment leaching of gas produced by degradation of
biogenic sources, among others. Each of these examples represents a critical component controlling
the global cycling of gases, and the rise velocity of the bubbles sets the rate of and available time
scale for dissolution and may dominate the vertical advection. Rise velocity of a single bubble in a
quiescent fluid is well characterized, but for environmental flows, bubbles rarely occur in isolation.
When bubbles are continuously released or released in groups, their rise velocity can be altered by
interaction with neighboring wakes1–5 or by a net upward water flow due to the collective buoy-
ancy of the bubble swarm or plume.6–9 Ambient currents and turbulence can further disrupt these
interactions.9–11 For high void-fraction bubbly flows, integral plume models can predict the average
flow behavior.8,12 However, many of these naturally occurring bubble sources are too weak to form
a coherent plume, yet too populated to assume bubbles do not interact, and their rise velocities may

a) Author to whom correspondence should be addressed. Electronic mail: wanger168@tamu.edu

1070-6631/2015/27(10)/103301/20/$30.00 27, 103301-1 © 2015 AIP Publishing LLC


103301-2 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

differ from the single-bubble case. Here, we present laboratory experiments and a model to predict
the rise velocity of bubbles in such weak bubble streams, accounting for the effects of multiple
wake interaction and crossflow. We limit our study to point releases of ellipsoidal-shaped bubbles,
analogous to bubble streams emanating from natural gas seeps in the oceans. These results are
important to quantify the bubble rise velocity in natural systems and to bridge the gap between
isolated bubble behavior and flows in bubble plumes.
There has been considerable attention given to bubble rise from natural gas seeps on the sea-
floor because they transport a considerable mass of hydrocarbon vertically through the ocean water
column. Seep bubble plumes are diverse, with a range of bubble size distributions, flow rates, and
variability over spatial and temporal scales.13–15 Because there is a large flux of methane from
natural seeps, and because methane is a potent greenhouse gas, it is important to accurately predict
the rise velocity of natural gas bubbles from seeps to determine the potential flux to the atmosphere.
Most seep plumes are weak, with a relatively small flux of bubbles coming from a single vent or
series of vents along a crack in the seafloor. For these seeps, bubbles follow a chain-like trajec-
tory with interacting wakes. Models for the rise of single methane bubbles in the oceans tend to
under-predict the rise height from seeps.15,16 This has been attributed to errors in the mass transfer
rate stemming from clathrate hydrate formation of methane and seawater at low temperature and
high pressure,17 and to upwelling flow due to plume effects or bubble-bubble interaction which is
neglected in the models.18,19 The focus of this paper is to quantify the latter effect under conditions
similar to the deep ocean, with low turbulence energy and weak currents.
Investigations have been carried out to study the interaction of spherical bubbles while they
rise in pairs or in chain.20–22 To simplify the bubble-bubble interaction problem, bubble pairs have
been extensively studied to investigate the wake effect of the leading bubble on the dynamics of
the tailing bubble. For example, Harper,20 and Yuan and Prosperetti22 suggested the existence of an
equilibrium spacing for single spherical bubble pairs, with different drag coefficients applying to
the leading and tailing bubbles. Katz and Meneveau23 experimentally studied the rising dynamics
of small bubble pairs to investigate the effect of wake-induced fluid motion at Reynolds number
ranging from 0.2 to 35. They also found the wake-induced motion resulted in different drag coef-
ficients, dependent on the vertical spacing between two bubbles and the bubble diameter, but their
experiment showed the bubbles contacted each other instead of rising with an equilibrium spacing.
Several studies have also investigated continuous releases of spherical bubbles. Martin and
Chandler21 found an increase of air bubble rise velocity with decreasing spacing between neighbor-
ing bubbles for bubble diameters ranging from 0.32 to 0.57 mm. However, Celata et al.24 observed
no significant increase in rise velocity for various bubble spacing for their vapor bubbles in the
diameter range of 0.1–0.7 mm. Ruzicka4 suggested a modified drag coefficient for in-chain bubbles
that can be applied to all bubbles independently of their position, based on a balance of creation and
decay of the disturbances. This drag coefficient is smaller than that of an isolated bubble and ac-
counts for the local (i.e., from the nearest neighboring bubble) and distant (i.e., from other bubbles
besides the nearest neighbor) wake effects, decreasing with decreasing bubble spacing.
For ellipsoidal bubbles, significant deformation appears on the bubbles and the rising trajec-
tories are oscillatory. Terminal velocity and motion behavior of an isolated ellipsoidal bubble while
rising are significantly influenced by the initial shape deformation.25 When bubbles rise in pair, the
trailing bubble is influenced by the wake of the leading bubble, and consequently, the rise velocity
can be strongly changed, even resulting in collision, as observed by Brucker.26 Additionally, when
ellipsoidal bubbles rise in chains, their wakes develop in a quasi-steady way, and the wake insta-
bility generates vortices that interact with the bubbles within the wake region.27 The experimental
study of Sanada et al.28 on nitrogen gas bubble chain demonstrated the impact of hydrodynamics on
the bubble trajectories with various bubble release frequencies. These dynamics and their influence
on the rise velocity of the in-chain bubbles are complex and, therefore, not well understood.
Marks2 measured rise velocities for continuously released bubbles in still water with the bub-
ble diameter ranging from 0.6 mm to 10 mm (which includes the ellipsoidal range) and release
frequency ranging from 20 to 400 bubbles/min. Vertical spacing between bubbles decreased and
the wake-induced fluid velocity increased with increasing release frequency. Consequently, the rise
velocity of in-chain bubbles increased with increasing release frequency, and Marks assumed the
103301-3 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

net rise velocity could be computed as the superposition of the isolated bubble terminal velocity
and the wake induced fluid velocity in the centerline of the wake. He found that turbulent theory in
circular wakes could predict the statistics of the rise velocity of these bubbles, where the centerline
fluid velocity in the wake Vc ∝ l −2/3 and the half-width of the wake b ∝ l 1/3, with l being the vertical
spacing between two neighboring bubbles.29 This result agreed with the study of Ellingsen and
Risso27 on quasi-steady behavior of turbulent wakes behind bubbles. However, the existing studies
only considered the local wake effect, that is, the wake of the nearest leading bubble. Here, we will
add consideration of the distant-wake effect from other upstream bubbles on the rise velocity of the
in-chain bubbles in still water. We also use the same turbulent wake theory to develop a model to
describe the bubble rise velocity in crossflow condition with simplified assumptions.
In the real ocean environment, horizontal current, or crossflow, drifts the bubbles laterally, such
that the bubbles are no longer aligned in the vertical direction but have a certain angle to the bubble
chain depending on the magnitude of the ocean current and bubble rise velocity. This displacement
moves bubbles off the centerline of the upstream wakes, and as a result, the wake effect on the
rise velocity of in-chain bubbles should be re-evaluated in crossflow condition. To help fill the
knowledge gap for such flows, we conducted a series of laboratory experiments to address the basic
questions: How does crossflow affect in-chain bubble rising dynamics? Does the wake theory still
apply for such flow? How can we quantify the bubble rise velocity under the influence of various
crossflow conditions similar to the deep ocean environment?
To answer these questions, we consider quiescent and flowing conditions separately. Sec. II
presents the experimental setup and image data processing. The bubble rise in still water is dis-
cussed in Sec. III, including analysis of bubble shape and the relation between bubble rise velocity
and release frequency (ranging from 84 to 734 bubbles/min). For these conditions, we extend the
Marks model to account for the wake effect of all leading bubbles (i.e., the distant wake effect)
rather than just the closest neighboring bubble only. Sec. IV presents experiments of bubble rise
velocity under crossflow condition (ranging from 2 to 15 cm/s). The wake interaction decreases
with increasing crossflow, and we present both an empirical and a theoretical model to describe the
transition and to collapse the non-dimensional data to a predictive curve.

II. METHODS
A. Experimental setup
The experimental setup is shown schematically in Figure 1. The experiments were carried out
in the Ocean Engineering Fluid Dynamics Laboratory at Texas A&M University. The experiment
was arranged into three stages. Figure 1(a) shows the setup for stage one, which was conducted in
still water (referred to as case S hereafter). In this stage, the tank was filled with tap water (note: the
total dissolved solids content was 548 mg/l according to the 2014 water quality report from the tap
water source, Texas Commission on Environmental Quality Public Water System No. 0210017) to a
depth of about 80 cm. Air was supplied through flexible tubing and released through a straight-pipe
orifice with diameter of 4 mm mounted on an L-shaped frame from the bottom of the tank. It should
be noted that the formation of the bubbles in this study was spontaneous, and the bubble release fre-
quency and size were controlled by “natural” characteristics. Bubble detachment from the tip of the
tube is the result of balance between buoyancy and surface tension. Alternatively, the characteristics
of bubbles, such as initial size and release frequency, can be precisely controlled using dedicated
methods.28,30 The bubbles were released in the center of the tank at different flow rates to create
successive in-chain bubbles with different spaces between two consecutive bubbles. Flow rate of
the releasing bubbles was determined by displacing water from an inverted graduated cylinder and
using a stopwatch to measure the total time to collect at least 500 ml of air. A laboratory imaging
system was used to collect shadow bubble images for measuring bubble properties such as bubble
size and rise velocity. A halogen lighting system and a translucent screen were used to provide
backlight for illumination of the bubbles. A high-speed camera (Vision Research Miro M340) was
placed at one side of the tank to take bubble images with a frame rate of up to 200 fps and an
exposure time of 250 µs. This experiment serves as a re-examination of turbulent wake theory, as in
103301-4 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 1. Schematic of experimental setup (not drawn to scale). A high-speed camera was placed at one side of the tank. Field
of view (FOV) of the camera was schematically shown in the figure. (a) Case S: continuous bubble release in still water.
For each case with a certain flow rate, the images were taken at three different depths. (b) Case T: the bubble release source
was towed at different speeds that were controlled by a linear stepper motor controller. (c) Case C: horizontal current (about
12 cm/s) was generated by the recirculation pump system of the wave tank.

Marks,2 and as the limiting case of crossflow, with crossflow velocity = 0. Table I summarizes the
experimental parameters in case S.
Each experiment started by adjusting the air supply to achieve a constant flow rate. For each
experimental condition with different flow rate, three sequences of images were acquired: two
before measuring the flow rate using the graduated cylinder and a third after measuring the flow
rate. For each image sequence, 12 GB of data are acquired, limited by the internal memory of
the camera. The duration of each image capture was from 10 s to 50 s depending on the selec-
tion of camera resolution and frame rate. The statistics showed consistent results for all image
sequences.
The stage two and three experiments were designed to investigate the rise velocity of the
in-chain bubbles under crossflow conditions. Figures 1(b) and 1(c) show the sketch of the experi-
mental setup, and Tables II and III list the parameters in the experiments. In stage two, the crossflow
scenarios were simulated with a towed bubble source using various tow speeds. The approach of
towing the release source has been used to simulate crossflows for studying jet and plume dynamics
in many experiments.31 The mounting frame for the bubble release (i.e., the L-shape frame) was
bolted to a linear stepper motor controller, which provides precise tow speed. The horizontal part
of the L-shape frame was long enough such that the measurement region was not influenced by
the disturbance due to the movement of the towing system. The same measurement procedure as
in case S (i.e., stage one) was repeated with four different tow speeds, Uc = 2, 5, 10, and 15 cm/s

TABLE I. Summary of experimental parameters in still water experiment (case S).

Case S f (1/s) l (cm) Q (ml/min) D (mm) VB (cm/s) Re CD We

Q1 0.75 32.44 2.8 5.2 24.45 1354 1.21 4.41


Q2 2.43 11.63 9.1 5.1 26.21 1440 1.04 5.03
Q3 2.77 7.56 11.5 5.2 26.73 1519 1.04 5.41
Q4 4.28 7.33 13.7 5.4 27.20 1281 0.83 4.64
Q5 3.77 6.83 16.5 4.4 26.86 1480 1.00 5.29
Q6 5.89 4.98 29.4 5.1 27.95 1530 0.92 5.70
Q7 8.30 4.21 38.4 5.0 30.57 1595 0.73 6.49
Q8 8.65 4.07 48.8 5.3 29.66 1724 0.86 6.81
Q9 8.81 3.99 56.9 5.5 31.08 1859 0.81 7.69
Q10 9.37 3.55 65.8 5.6 31.81 1936 0.79 8.20
Q11 11.69 2.87 76.2 5.7 33.48 2087 0.73 9.31
103301-5 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

TABLE II. Summary of experimental parameters in crossflow experiment


(case T).

Case T Q (ml/min) D (mm) Uc (cm/s) VB (cm/s)

Q1 4.6 4.7 2 24.03


4.6 4.6 5 23.82
4.6 4.6 10 23.73
4.6 4.7 15 23.81

Q2 14.4 5.0 2 24.55


14.4 5.0 5 24.14
14.4 5.0 10 23.76
14.4 5.0 15 23.48

Q3 22.2 5.1 0 28.90


22.2 5.1 2 25.30
22.2 5.1 5 24.42
22.2 5.0 10 23.88
22.2 5.0 15 24.40

Q4 31.6 5.0 0 30.47


31.6 5.1 2 26.08
31.6 5.0 5 24.75
31.6 5.0 10 24.75
31.6 5.0 15 24.09

Q5 54.0 5.6 0 32.86


54.0 5.7 2 28.25
54.0 5.6 5 25.66
54.0 5.7 10 24.94
54.0 5.5 15 25.01

Q6 70.0 5.6 0 35.34


70.0 5.7 2 30.20
70.0 5.6 5 26.49
70.0 5.7 10 25.66
70.0 5.5 15 24.42

Q7 91.6 6.1 0 36.69


91.6 6.3 2 31.22
91.6 6.4 5 26.66
91.6 6.4 10 25.33
91.6 6.2 15 24.61

(referred to as case T hereafter). During stage three, we performed the experiments in a constant,
real crossflow, generated using a recirculation pump in a long, open channel (referred to case C
hereafter). To smooth the velocity distribution at the flume inlet, an array of bricks and a grid of
PVC (polyvinyl chloride) pipes were placed immediately after an inflow head tank. A Vectrino II

TABLE III. Summary of experimental parameters in crossflow experiment


(case C).

Case C Q (ml/min) D (mm) Uc (cm/s) VB (cm/s)

Q1 4.9 4.6 12 23.68


Q2 15.5 4.7 12 23.88
Q3 35.0 5.1 12 23.74
Q4 70.0 5.7 12 24.54
103301-6 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 2. Sample images: (a) case S, (b) case T (tow speed = 15 cm/s), (c) case C (current velocity = 12 cm/s).

profiling Acoustic Doppler Velocimeter (ADV) was used to verify the vertical profiles of the cross-
flow velocity. The measurement section was sufficiently far downstream to have a fully developed
channel flow. The flow velocity in the camera’s field of view (FOV) was 12.4 cm/s on average with
a standard deviation of 1.2 cm/s. The background turbulence had a turbulence intensity of 0.08,
defined as the ratio of root-mean-square of turbulent velocity fluctuations to the Reynolds averaged
mean velocity; the Reynolds number of background flow was about 105, which is defined as Uc H/ν.

B. Image data processing


Figure 2 shows sample images for each case. Figure 2(a) shows a zoom-in image for case S
to illustrate the shape of bubbles. The sample image shows that the bubbles experienced significant
deformation during the rising process, and the shapes of individual bubbles were very different.
Generally, the shapes of bubbles are in ellipsoidal wobbling category.32 Figures 2(b) and 2(c) show
the entire FOV of the camera to depict the pattern of in-chain bubbles under crossflow condition
(Cases T and C). The horizontal displacement of bubbles is determined by the rise velocity and by
the crossflow speed.
All images were analyzed with an in-house code using the image processing toolbox (version
8.3) of Matlab (MathWorks C ). Due to the very low void fraction in such bubbly flow, the bubbles
can be easily distinguished from the background using an edge detection and thresholding algo-
rithm. As a result, the raw images can be converted to binary images where the bubble occupied
pixels were marked as 1 and background pixels were marked as 0. An example is shown in Figure 3.
Instantaneous rise velocity was calculated by tracking the vertical displacement of the center of
the detected bubble during the time between two consecutive image frames. The instantaneous rise
velocities of all detected bubbles during the entire recorded time were then analyzed to obtain
the statistics of the bubble rise velocity. The bubble size was calculated using two approaches. In
one approach, the equivalent spherical bubble diameter is calculated from the best-fit ellipse with
the same volume following D = (ab2)1/3, where a and b are the major and minor axis lengths,
respectively (see Figure 3(a)). In the other approach, the area-equivalent circular bubble diameter
was computed to match the projected area of each bubble as seen in the binary bubble image (see
Figure 3(b)).
Figure 4(a) shows the comparison of bubble diameters calculated with two different methods
for all experiments in Tables I–III. Generally, the equivalent spherical bubble diameter based on
the ellipsoid volume correlates well with the area equivalent bubble diameter with the projected
two-dimensional bubble image. The diameter using projection area (D1) is larger than that was
determined from the ellipsoid volume (D2), and the regression shows the relation of D1 ≈ 1.08D2.
To examine the performance of the image processing, Figure 4(b) compares the flow rates esti-
mated from the two volume measures, Qimg, to the flow rate measured in the experiments using the
graduated cylinder Qmeasured. From the image processing, the flow rate was determined by counting
103301-7 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 3. Sample processed images with bubble detection result. (a) The center of detected bubbles and the fitted ellipses
are shown. The result is used to determine the equivalent spherical bubble diameter. (b) The corresponding binary image with
detection result of location and size of the bubbles. The result is used to determine the area equivalent bubble diameter.

FIG. 4. Validation of image processing results. (a) Comparison between two methods for determining the bubble size. A
typical error bar is shown to present the standard deviation of the measurement data. The full range of the standard deviation
is 0.44–1.20 mm with the mean of 0.58 mm for D1 and 0.47–1.11 mm with the mean of 0.61 mm for D2. The solid line
represents D1 = D2. The dotted line represents the best fit of the data: D 1 ≈ 1.08D2. (b) Comparison between the measured
flow rate and that determined from the image processing. The solid line represents Q img = Q measurement.
103301-8 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

bubbles to determine their release frequency and multiplying by the average volume of bubbles
based on the two volume estimates D1 and D2. Both volume estimates appear biased in Figure 4(b),
with estimates using D1 as bubble diameter overestimating release flux (slope 1.13) but estimates
based on D2 underestimating the overall flow rate (slope 0.88). Due to the cubic relation between
the flow rate and the bubble diameter, the deviation of imaging-based flow rates using the diameter
of D1 from the measured values increased rapidly as the flow rate exceeded 70 ml/min, whereas
estimates using D2 begin to converged on the measured flow rate. Therefore, in this study, we used
D = D2 as the bubble diameter in the following analysis.

III. CASE S: IN-CHAIN BUBBLE RISING IN STILL WATER


A. Bubble size and shape
Table I reports the mean bubble size D and rise velocity VB measured in the case S experiments;
VB is the vertical velocity of the bubble in the fixed camera reference frame, which includes a
slip velocity Vs between the bubble and the water superposed with the local vertical water velocity
Vw . Corresponding values of the diameter Reynolds number Re = VB D/ν and the Weber number
We = ρVB2 D/σ are also reported, where we take the kinematic viscosity of water ν as 10−6 m2/s and
the interfacial tension between air and water σ as 73 dynes/cm2; ρ is the density of water, taken as
999 kg/m3. The drag coefficient CD is computed from CD = 43 gVD2 .
B
Tsuge and Hibino3 (referred to as TH72 in the figure legends) studied the motion of air bubbles
that were continuously released from a single orifice with diameter 1.6 mm into distilled water,
which is very similar to our case S. In their experiment, release frequency varied from 180 to
900 bubbles/min, which led to equivalent spherical diameters in the range from 5 to 9 mm. They
compared the bubble shape properties for their in-chain bubbles in distilled water to that for isolated
bubbles from previous studies. In this section, we also examine the bubble shape properties for
in-chain bubbles in tap water.
Figure 5 plots the mean values of the aspect ratio of the projected ellipse, defined as the ratio
of the minor-axis length to major-axis length, (b/a)m, to the equivalent spherical bubble diameter
(Figure 5(a)) and to We (Figure 5(b)) for case S. The data from Tsuge and Hibino3 were digitized from
their figure and re-plotted here for comparison. For reference, the relationship for isolated bubbles
was included for doubly distilled water (dashed line) and contaminated water (solid line),33 which
were also obtained from their figure. Tsuge and Hibino3 demonstrated a clear decreasing trend of
aspect ratio with increasing bubble diameter over the range 4–9 mm, regardless of the bubble release

FIG. 5. (a) Relationship between mean aspect ratio of ellipsoidal bubbles, (b/a)m, and mean value of equivalent spherical
bubble diameter for different flow rates in case S. (b) Relationship between (b/a)m and Weber number (We). The data from
Tsuge and Hibino3 are also included. The solid line represents the relationship for isolated bubbles in contaminated water
and dashed line represents for isolated bubbles in doubly distilled water.
103301-9 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

frequency. Our data using tap water over a narrower diameter range (4.4–5.6 mm) cluster well with
their data in distilled water (Figure 5(a)). However, for in-chain bubbles, release frequency and flow
rate are important factors in the rising dynamics of the bubbles. The effect on the bubble shape can be
reflected in the relationship between the aspect ratio and We. Our data extend to lower We than those
in the work of Tsuge and Hibino3 and appear to follow the same trend (see Figure 5(b)).
The data in Figure 5(a) match best to the distilled water correlation, whereas the data in 5(b) are
closer to the contaminated water line. Contamination works to immobilize the bubble-water interface,
which reduces the internal circulation within the bubble and reduces the slip velocity, but without
affecting the interfacial tension.32 Since the data in Figure 5(a) ignore the rise velocity, the agreement
between our data for tap water and the data of Tsuge and Hibino3 for distilled water suggests that both
experiments had very similar interfacial tension, hence similar shape to equivalent diameter metrics.
The trend of the data in Figure 5(b) to better match the contaminated water data indicates that both
data sets had rise velocities consistent with dirty bubbles since the Weber number includes the rise
velocity. This is expected since it is very difficult to achieve a high enough purity to obtain clean
bubble rise velocities for bubble experiments.32 Moreover, even without significant change of static
interfacial tension, the electrolytes in water could generate a fore-to-aft interfacial tension gradient
on the rising bubble surface. This in turn results in the Marangoni stresses that suppress the bubble
rise velocity.34

B. Bubble rise velocity under turbulent wake effect


The trajectories of bubbles while rising in the stagnant ambient due to buoyancy vary due to
several factors.27,30 For small bubbles (mostly spherical and ellipsoidal with close axis lengths), they
usually rise in a vertical straight line. The larger ellipsoidal bubbles rise either following zig-zag in
a 2D plane or spiral in 3D due to path instability.27,28,35–37 Aybers and Tapucu39 reported a diagram
of trajectories associated with bubble size and Reynolds number. Visual examination during the
current experiment and a previous 3D study38 for similar bubbles and environment showed that our
bubble rise regime is 3D spiraling.
Air bubbles typically accelerate immediately after release from an orifice and, for elliptic
shape, approach periodic oscillation within a very short distance in still water.40 Above the accel-
eration region, rise velocity is quasi-steady, and bubbles are associated with periodic oscillation
motion. All measurements in case S are in this periodic region. Our data did not show significant
variation on the statistics of bubble size, shape, and rise velocity while measuring at different water
depths (see three fields of view in Figure 1).
For air bubbles rising in water, rise velocity of an isolated bubble (or terminal velocity) de-
pends on the bubble size, shape, and other factors including water contamination32 and bubble
deformation due to initial release condition.25 For a continuously released bubble chain in still
water, the bubble velocity is also affected by other bubbles in the chain since the trailing bubbles
lie in the wake region of the leading bubbles, resulting in increased in-chain bubble rise velocity
compared to the isolated bubble case. The enhanced bubble rise velocity has been characterized
by the bubble release frequency or spacing between two neighboring bubbles.2,23,24,40 However, the
wake behind bubbles in different bubble shape regimes may vary significantly due to various bubble
rising dynamics. For instance, the wake effect seems negligible for small, spherical bubbles, e.g.,
0.13–0.6 mm,24 0.06–1.2 mm2, probably due to the narrow, laminar wake. In the ellipsoidal shape
regime, enhancement of bubble rise velocity due to the wake effect is significant.2,3
The bubble size range in this study was maintained in the ellipsoidal shape range; therefore, we
expect to observe enhanced rise velocity due to the wake effect. Figure 6(a) clearly demonstrates
the enhancement of rise velocity with increasing flow rate. In the previous studies of Tsuge and
Hibino3 and of Marks,2 enhanced rise velocities with increasing release frequency follow individual
curves for each bubble diameter. The data of Marks2 also showed that the rise velocity collapsed on
a universal curve for all bubble sizes when written in the format of flow rate, i.e., VB = 18.1Q0.141
for flow rate exceeding 20 ml/min. This relationship is also supported by our data and the data
of Tsuge and Hibino3 (see Figure 6(b)). While the data appear to cluster around this relationship,
trends remain dependent on bubble diameter, but with lower sensitivity such that individual curves
103301-10 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 6. The relationship between bubble rise velocity and flow rate. (a) The data in this study. The error bars show the
standard deviation of measurement data. (b) The data from Tsuge and Hibino3 are also plotted. The solid line represents the
best fit line of the data of Marks,2 VB = 18.1Q 0.141.

are less obvious than plots versus release frequency. For example, the data from Tsuge and Hibino3
for bubbles smaller than 6 mm have consistently higher rise velocity than that for bubbles larger
than 6 mm.
A non-dimensional representation of rise velocity and its relationship with flow rate can be
derived using turbulent wake theory (see the Appendix following Marks),2,29
) 5/3 ) 2/3 1/3
12gQ2
( (
VB VB
∝ * 2 4 3+ ,
− (1)
V0 , π β V0 D -
V0
where V0 is the terminal velocity of isolated bubbles, g is the acceleration of gravity, and β is a
coefficient related to the ratio of mixing length to the half-width of the wake.29
The non-dimensional parameter space suggested by Equation (1) is explored in Figure 7, where
the normalized rise velocity VB/V0 is plotted versus the parameters in the right-hand-side of (1).
Figure 7(a) is plotted using release frequency f in number/s in place of Q, where Q = πD 3 f /6, as
suggested by Marks.2 In addition, the relation between non-dimensional rise velocity and flow rate
is shown in Figure 7(b), with the fit between the data and the dashed regression line having an R2
of 0.93 and 0.94 in Figures 7(a) and 7(b), respectively. This is an excellent fit for data with this
level of variability, and biases depending on bubble diameter no longer appear. This correlation is
valid for ellipsoidal bubbles only and is validated for the diameter range from 0.8 to 10 mm and
for release rates up to 350 ml/min. Other bubble shapes exhibit different wake structure and would

FIG. 7. Non-dimensional relationship between the rise velocity and (a) release frequency, and (b) flow rate. Data of Tsuge
and Hibino3 and Marks2 are digitized and adopted here. The dashed lines represent regression of the data.
103301-11 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 8. Comparison between the models with and without distant wake effect. The relationship between non-dimensional rise
velocity and flow rate is plotted following (a) Equation (1). Data from Tsuge and Hibino3 and the regression line of Marks2
are also shown here. The dotted line ((VB /V0)5/3 − (VB /V0)2/3 = 1.669(g Q 2/V0 D 3)2/3 + 1.405(g Q 2/V0 D 3)1/3, R 2 = 0.92)
shows the nonlinear feature of the data points using the model with non-distant wake effect. The solid line (slope = 2.02)
represents the regression using our data and of Tsuge and Hibino,3 R 2 = 0.89, (b) Equation (6). Despite scattering, the
nonlinear behavior of the model is reduced. The solid lines (slope = 0.55) represents the regression using our data and of
Tsuge and Hibino,3 R 2 = 0.82.

not be expected to follow this relationship. Likewise, higher release rates increase the void fraction
such that the flow transitions from in-chain bubbles to a plume, where the higher bubble-bubble
interaction would cause this equation to break down.
Recall that Equation (1) considers only the wake of the closest upstream bubble on the rise
velocity. Marks2 noted for his data that at high flow rate, the non-dimensional velocity data in the
theoretical model showed a non-linear increasing trend. This curve is shown in Figure 8(a) as the
dotted line. At higher flow rates, the spacing decreases, and it becomes more likely that the wakes of
multiple upstream bubbles affect the rise velocity of a given in-chain bubble.
Due to mathematical difficulties in describing bubble-bubble interactions and wake dynamics
for the range of bubble diameter and Reynolds number in the present study, no theory is currently
available to explain the in-chain bubble dynamics and the effect of distant wakes on rise velocity. To
simplify the complex flow in this kind of multiple shear layer structure, a superposition scheme has
been proposed and validated for flow characteristics in cylinder wakes.41,42 The experiment of Zhou
et al.42 suggested that the superposition hypothesis was able to accurately predict the mean velocity
field for complex cylinder wake and only high-order turbulent terms were affected by non-linear
interaction of the wakes. Ruzicka4 demonstrated a cumulative effect of velocity disturbance on
drag coefficient using linear superposition of the individual contributions of spherical bubbles at
small Reynolds numbers. Similarly, we consider a linear superposition of multiple wakes of leading
bubbles such that

VB = V0 + Vw,c1 + Vw,c2 + · · · + Vw,c n + · · ·, (2)

where Vw,cn is the centerline fluid velocity in the wake of the nth leading bubble above the bub-
ble of interest, expressed as Vw,c n ∼ V0[CD A/( βnl)2]1/3, where l is the spacing between bubbles.
Therefore, Equation (1) should be modified to
) 5/3 ) 2/3 N 1/3
12gQ2 +
( (
VB VB 
− ∼ n −2/3*
, (3)
, π β V0 D -
V0 V0 2 4 3
n=1

where N is the total number of superposed wakes. In an ideal successive bubble rising model, N
can be set as infinity. However, in a real situation, N varies with the total water depth (H) and the
103301-12 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

spacing (l), which is determined by flow rate and bubble size,


H 6HQ
N = integer of = integer of . (4)
l πVB D 3
N
n−2/3 can be approximately expressed as a power function as

To simplify Equation (3),
n=1

N

n−2/3 ≈ c1 N m . (5)
n=1

Consequently, Equation (3) can be further written as


) 5/3+m ) 2/3+m )m 1/3
2
( ( (
VB VB 6HQ
* 12gQ + .
− ∼ (6)
V0 V0 πV0 D 3
, π β V0 D -
2 4 3

Equation (6) is the updated rise velocity model for in-chain ellipsoidal bubbles with consider-
ation of the distant wake effect. The prediction using this model is shown in Figure 8(b). Despite
somewhat greater scattering (R2 = 0.82 compared to 0.89 in Figure 8(a)), the updated model clearly
reduces the nonlinear relationship between the model prediction and the data, yielding a linear trend
over the validation range considered. However, we notice that the new model seemed to have worse
overall correlation with data than the model without the distant wake effect. This is due to nonlinear
interactions of the turbulent wakes ignored in the superposition model and also resulting from the
numerical approximations in the above analysis. For example, the flow motion in turbulent wake of
the leading bubbles could become entirely chaotic as bubble rise sufficiently far away, which may
not have noticeable upwelling contribution to the trailing bubbles.

IV. CASES T AND C: IN-CHAIN BUBBLE RISING WITH CROSSFLOW


A. Bubble-bubble interaction in crossflow
Katz and Meneveau23 and Brucker26 both showed that the bubble-bubble interaction is a main
factor affecting rise velocity. To understand the effects of crossflow on rise velocity, we begin by
studying the bubble-bubble interactions as the crossflow is increased. Figure 9 shows a time series
of bubble rise for case T, Q4 at crossflow velocity Uc = 2 cm/s. The velocity vectors are also plotted
in the images indicating the instantaneous magnitude and direction of the bubble motion. Two
bubbles are highlighted within the red ellipse in the figure at each time step. As can be seen, the left
hand side bubble was at the lower position at t = 0 but passed the right-hand bubble and ended in a
higher position at t = 1 s. The magnitude and direction of the bubble velocity of each bubble in this
pair varied significantly during this time. The interacting pair also leaves a large gap in the bubble
chain, as evident especially beginning at 7/10 s. Bubbles overtaking each other was not observed in
the non-crossflow cases, and the associated continuous change in bubble velocity and its location in
the wake region (e.g., lateral displacement relative to upstream bubbles) may have strong influence
on the averaged bubble rise velocity under different crossflow conditions. It should be noted that the
previous study of Kok43 showed strong bubble-bubble interaction while bubbles were close to each
other. Such bubble-bubble interaction could be strong to change their relative location, which may
cause coalescence, bounce, or overtaking with different hydrodynamics and water contamination.
Hence, the bubble behavior with bubble-bubble interaction and hydrodynamics in the different
environment is affected by multiple factors and thereby complicated.
Figure 10 shows the 2D trajectory paths of bubbles rising for crossflow cases (case T, Q4 and
case C, Q3) at different crossflow velocities. At Uc = 0 cm/s, bubble paths are constrained within
a relatively narrow lateral region, with a width of approximately 2 cm. In such case, the turbulent
wake theory applies and the rise velocity of in-chain bubbles can be well predicted by the above
model. With an increase in the crossflow velocity, the lateral spacing between neighboring bubbles
is ideally determined by the crossflow velocity, which should be evenly distributed if bubble-bubble
103301-13 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 9. A time series of bubble rising for case T, Q4, and Uc = 2 cm/s.

interaction was not considered. However, Figure 10 shows clearly evidence of “grouping” or “pair-
ing” of bubble trajectories during the rising process while Uc > 0. The projected bubble trajectories
overlap and depict a “stripe-like” pattern, indicating significant interaction during their ascent. The
trailing bubbles tend to follow the path of the leading bubbles and create a “group” of trajectories,
which may be attributed to the entrainment effect in the wake region until the crossflow breaks the
effect of the entrainment and forms a new “group.” For low crossflow velocity (e.g., Uc = 2 cm/s),
more bubbles interacted with complex, group trajectories, whereas more pair trajectories of bubbles
were observed for relatively larger crossflow velocities (e.g., Uc = 10 cm/s). At Uc = 15 cm/s, the
bubble-bubble interaction was further reduced. In that case, the bubble trajectories were more inde-
pendent, with occasional bubble pairs occurring. Here, we performed a manual check of “grouping”
and “pairing” effect of bubble trajectories. At Uc = 2 cm/s, about 30 bubble were released in 10 cm
horizontal distance resulting in 6 groups with 5 bubbles within one group on average. With increase
of the crossflow velocity, the number of bubbles in a group decreases. While Uc = 5 cm/s, about
14 bubbles rose in 4 groups with averaged 3.5 bubbles per group. For Uc = 10 and 15 cm/s, the
averaged number of bubbles per group was 1.75 and 1.25, respectively.
103301-14 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 10. ((a)-(e)) Trajectories of bubble rising for case T, Q4 at different crossflow velocities. (f) Trajectories of bubble rising
for case C, Q3, represented in a same way as in (a)–(e).

B. Rise velocity in crossflow


Figure 11(a) shows the relationship between the dimensional bubble rise velocity and crossflow
velocity (tow speed for case T) at different flow rates of air bubbles. The dashed line in the figure
gives the terminal rise velocity of an isolated bubble. The data clearly show that the in-chain bubble
rise velocity decreases rapidly with increase in crossflow velocity at all flow rates. For each case, the
rise velocity appears to asymptote to the terminal velocity of the isolated bubble at high crossflows,
e.g., after Uc exceeded 10 cm/s for this dataset.
Figure 11(b) shows the rise velocity as a function of flow rate. In crossflow conditions, the
increased rise velocity of the in-chain bubbles with increasing flow rate is strongly suppressed by
increasing crossflow velocity; above Uc > 5 cm/s, the enhancement of in-chain bubble rise velocity
is less than the experimental error.
Figure 12 examines the non-dimensional relationship between rise velocity and flow rate in
crossflow condition following the same normalization format as in Figure 7(b). This model does
not include the lateral displacement of bubbles relative to the wake centerlines of upstream bubbles
due to the crossflow; hence, it is not expected to collapse the data. However, under the effect of
crossflow, this non-dimensional relation gives quite linear trends, with the slopes depending on the
strength of the crossflow, only.
To obtain a relationship between the rise velocity of in-chain bubbles and increasing flow rate
in crossflow, we propose a simple empirical method to revise the non-dimensional relationship as
shown in Figure 12 to account for the different slopes caused by the crossflow. Through empirical
analysis of the data, we find that the slopes of the best fit regression lines for each dataset follow
an exponential relation with the crossflow velocity given by s = 1.5 exp(−0.3Uc ) (see Figure 13(a)).
103301-15 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 11. (a) Relation between the bubble rise velocity and crossflow velocity at different flow rates of air bubbles for cases
T and C. (b) Relation between the rise velocity and flow rate under different crossflow velocities. Errorbars show a typical
standard deviation of the measurement data. Data of case S are also included. The terminal velocity of the isolated bubble is
shown as dashed lines in both figures.

Using this relation, the non-dimensional model can be revised as


VB/V0 − 1 ∼ s(gQ2/V04 D 3)1/3. (7)
This empirical relation is plotted in Figure 13(b), with R = 0.97. This is a high degree of
2

performance, but is biased for some of the data, as the exponential function does not decay with
crossflow velocity as fast as the slope does. Nonetheless, this is a significant improvement over
Figure 12.

C. Effect of turbulent wakes in crossflow


A correction for the in-chain bubble rise velocity can also be developed from the turbulent
wake theory by incorporating the displacement by the crossflow in a more theoretical way. Under
crossflow condition, the bubbles drift such that the bubble chain tilts at a certain angle with respect
to the vertical direction depending on the rise velocity and crossflow speed. As a result, the trailing
bubbles are located at some offset distance away from the centerline of the wakes of the leading
bubbles, leading to a reduction of the wake influence on the rise velocity. Figure 14 shows a sche-
matic representation of the wake structure under crossflow condition. We continue our use of the

FIG. 12. Non-dimensional bubble rise velocity for all cases.


103301-16 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 13. (a) The relationship between the crossflow velocity and the slopes of the best-fit lines for VB /V0 − 1 ∼
s(g Q 2/V04 D 3)1/3; the solid line represents the exponential fitting: s = 1.5exp(−0.3Uc ). (b) Empirical model for non-
dimensional bubble rise velocity with various flow rates under crossflow condition, as in Equation (7). The solid line
represents the regression for all data: VB /V0 − 1 = 1.0s(g Q 2/V04 D 3)1/3 with the goodness-of-fit R 2 = 0.97.

near and distant bubble wake model so that the bubble rise velocity in crossflow condition can be
written as
VB = V0 + Vw,r 1 + Vw,r 2 + · · · + Vw,r n + · · ·, (8)
where Vw,r n is the vertical fluid velocity in the wake of the nth leading bubble above the bubble of
interest, and the subscript “rn” indicates the offset distance between the bubble of interest and the
centerline of the nth wake, that is, r n (see Figure 14).
In classic wake theory, the wake velocity profile can be described using a Gaussian distribution
with respect to the centerline velocity,
 ( ) 2
rn 
Vw,r n = Vw,c n exp − , (9)
 bn 
where bn is the half-width of the wake.
In this study, the offset distance away from the centerline of the wake was determined using the
crossflow velocity and bubble release frequency, i.e.,

FIG. 14. Schematic representation of wake structure under crossflow condition.


103301-17 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

Uc Uc πD3
r n = nr 0 = n =n , (10)
f 6Q
where r 0 is the horizontal distance between two closest bubbles (see Figure 14). Considering
bn ∼ ( βCD Anl)1/3 (see Appendix), the wake-induced velocity by the nth leading bubble can be
written as
2/3
πUc3 D 3V02
 
Vw,r n ∼ Vw,c n exp −n
 4/3* +  . (11)
, 18g βQ VB - 
 2

In non-dimensional form, we obtain
1/3 2/3
N
πUc3 D 3V02

VB − V0  −2/3* 12gQ2 + 
∼ n exp  −n 4/3* +  . (12)
, π β V0 VB D - , 18g βQ VB - 
2 2 2 3
 2
V0 n=1


Note that in this equation, the bubble rise velocity is implicitly represented on both sides of the
equation. This is due to the fact that the lateral separation depends on the crossflow velocity and the
rise velocity of bubbles, i.e., the in-chain bubble trajectory is the vector sum of these velocities. In
the exponential part of Equation (12), increasing Uc in the numerator leads to a decrease of VB in the
denominator, which in turn yields strong decaying of the wake enhancement as the bubble drifts out
of the wake centerline. Therefore, with increase of crossflow velocity, the decaying of the bubble
rise velocity should be very rapid as the wake effect decays following a superposition of shifted
Gaussian distributions (see Figure 14).
Because the lateral shift is greater for the distant wake than for the closest bubble wake, we test
this equation by setting N = 1. Equation (12) can be written as
1/3 2/3
πUc3 D3V02

VB 12gQ2 
− 1 ∼ * 2 2 2 3+ exp −
 * + . (13)
, π β V0 VB D -  , 18g βQ VB - 
V0 2

Therefore, we consider three normalized parameters


VB
VB+ = , (14)
V0
gQ2 gQ2
Q+ = (implicit), Q+ = (explicit), (15)
π β 2V02VB2 D 3 π β 2V04 D 3
πUc3 D 3V02 πUc3 D 3V0
Uc+ = (implicit), Uc+ = (explicit). (16)
18g βQ2VB 18g βQ2
Using these parameters, Equation (13) can be further written as

VB+ − 1 ∼ Q+ exp − Uc+


 2/3
. (17)
Following this normalization scheme, Figure 15 shows the relation among non-dimensional rise
velocity, flow rate, and crossflow velocity in both explicit and implicit formats. Here, the value of
constant β = 0.18 according to Schlichting.29
Although the data collapse to a universal curve following Equation (17), the curve does not show a
straight line with a single slope but rather has two regimes separated by the point exp[−(Uc+)2/3] = 1 −
1/e2 = 0.86 (i.e., Uc+ ≈ 0.06). When the crossflow velocity is zero, exp[−(Uc+)2/3] = 1, and the non-
dimensional parameter (VB+ − 1)/(Q+)1/3 also approaches 1. As the crossflow increases, exp[−(Uc+)2/3]
becomes less than 1 and (VB+ − 1)/(Q+)1/3 rapidly decreases. However, the rate of decrease becomes
much milder for exp[−(Uc+)2/3] less than 0.86. In Section IV, we demonstrated significant bubble-
bubble interaction, with bubble grouping at low crossflow velocity and pairing continuing even at
high crossflow. The bubble-bubble interactions reduce the ideal offset of the trailing bubbles shown in
Figure 14, which compromise the direct application of Equation (17). The behavior of changing slope
in Figure 15 may be attributed to such bubble-bubble interactions and differing degrees of bubble
groupings.
103301-18 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

FIG. 15. Two-regime model shows the non-dimensional relationship among rise velocity, flow rate, and crossflow velocity.
The errorbars show the measurement error propagation in following Equation (17). (a) Implicit, i.e., model with bubble rise
velocity implicitly shown in both sides; (b) explicit, i.e., model with terminal velocity of isolated bubble. The dashed lines in
two figures show the boundary that delineates strong bubble-bubble interactions and weak interactions. The fitted linear curves
for strong and weak interactions are (a) (VB+ − 1)/(Q +)1/3 = 5.51exp[−(Uc+)2/3] − 4.64, R 2 = 0.88 and (VB+ − 1)/(Q +)1/3 =
0.18exp[−(Uc+)2/3] + 0.047, R 2 = 0.62 for implicit normalization; (b) (VB+ − 1)/(Q +)1/3 = 3.68exp[−(Uc+)2/3] − 2.97, R 2 =
0.95 and (VB+ − 1)/(Q +)1/3 = 0.18exp[−(Uc+)2/3] + 0.046, R 2 = 0.58 for explicit normalization.

The boundary at 1 − 1/e2 (dashed lines in Figure 15) delineates these two regimes and per-
formed linear regression in both regimes. The regime with sharp slope (right-hand side of the dashed
line) is referred to “strong bubble-bubble interaction.” The best-fit lines using Equation (17) in this
regime with implicit and explicit formats are (VB+ − 1)/(Q+)1/3 = 5.51 exp[−(Uc+)2/3] − 4.64 (R2 =
0.88) and (VB+ − 1)/(Q+)1/3 = 3.68 exp[−(Uc+)2/3] − 2.97 (R2 = 0.95), respectively. The regime with
mild slope (left-hand side of the dashed line) is referred to the “weak bubble-bubble interaction,”
where the best-fine lines using Equation (17) with implicit and explicit formats are (VB+ − 1)/(Q+)1/3 =
0.18 exp[−(Uc+)2/3] + 0.047 (R2 = 0.62) and (VB+ − 1)/(Q+)1/3 = 0.18 exp[−(Uc+)2/3] + 0.046 (R2 =
0.58), respectively.

V. SUMMARY AND CONCLUSIONS


The rise velocities of in-chain bubbles in still water and with crossflow were studied for
ellipsoidal wobbling bubbles with moderate Reynolds numbers. We performed an experimental
investigation of the turbulent wake effect on the enhanced rise velocities with various flow rates and
crossflow velocities.
In-chain bubbles rising in still water were used as the limiting case with the crossflow velocity
equaling zero. The distant wake effect from multiple leading bubbles was considered in the turbu-
lent wake model and an updated model was proposed. Despite limited the small contribution to the
rise velocity from the distant bubbles in the updated model, consideration of the distant wake effect
clearly reduced the nonlinear feature of the non-distant wake turbulent (i.e., local) wake model.
Under crossflow condition, the enhancement of rise velocity of in-chain bubbles was sup-
pressed with the increasing crossflow velocity. With strong crossflow velocity (>10 cm/s), the
enhancement was found very limited, which resulted in rise velocities very close values to the
terminal velocity of the isolated bubble in still water. An empirical model was proposed to predict
rise velocity of in-chain bubbles with crossflow. Turbulent wake theory was also used to explain the
dynamics of the in-chain bubble under crossflow condition. The normalization scheme suggested
using the theory can be used to describe the rise velocity, flow rate, and crossflow velocity.
During the rising process, bubbles tend to interact and create a complex dynamics. Under cross-
flow condition, the bubble-bubble interaction reduced with increasing crossflow velocity. Asso-
ciated with the turbulent wake model, we defined a clear boundary to discriminate the “strong
interaction” and “weak interaction” among in-chain bubbles. We suggested using a “two-regime
model” with the same normalized parameters to predict the rise velocities of in-chain bubbles under
crossflow condition.
103301-19 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

ACKNOWLEDGMENTS
This research is supported by a grant from BP/The Gulf of Mexico Research Initiative. The
dataset presented in this paper is available through the Gulf of Mexico Research Initiative Informa-
tion & Data Cooperative, No. R1.x137.132:0007.44

APPENDIX: NON-DISTANT TURBULENT WAKE MODEL FOR IN-CHAIN BUBBLE


RISE VELOCITY
We recall the explanation of enhancement of in-chain bubble rise velocity using the turbulent
boundary layer theory for wake flow. Assumptions are made,2
(1) the influence of wake turbulence and wake velocity on drag coefficient is neglected such that
4gD
CD = ; (A1)
3V02
(2) turbulent theory in circular wakes applied here,29 where the ratio of wake centerline fluid
velocity to solitary body moving velocity relatively to the surrounding fluids (i.e., terminal
velocity) can be written as
( ) 1/3
Vw,c CD A
∼ . (A2)
V0 β 2l 2
Half-width of the wake region,
b ∼ ( βCD Al)1/3, (A3)
where A is surface area; β is a coefficient, defined as the ratio of mixing length to half-width
of the wake; the mean spacing between neighboring bubbles, l, can be derived from the flow
rate, bubble rise velocity, and the equivalent spherical bubble diameter,
rise velocity πVB D 3
l= = ; (A4)
release frequency 6Q
(3) Bubble rise velocity is superposition of the isolated bubble terminal velocity (V0) and the fluid
velocity in the centerline of the wake of its closest neighboring bubble (Vw,c ),
VB = V0 + Vw,c . (A5)
Combing the above equations yields the non-distant turbulent wake model for in-chain bubble
rise velocity in still water,
) 5/3 ) 2/3 1/3
12gQ2
( (
VB VB
− ∼ * 2 4 3+ . (A6)
V0 V0 , π β V0 D -
1 Z. L. Liu, Y. Zheng, L. Jia, and Q. K. Zhang, “Study of bubble induced flow structure using PIV,” Chem. Eng. Sci. 60, 3537
(2005).
2 C. H. Marks, “Measurements of terminal velocity of bubbles rising in a chain,” J. Fluids Eng. 95, 17 (1973).
3 H. Tsuge and S. Hibino, “The motion of gas bubbles generating from a single orifice submerged in a liquid,” Keio Eng.

Rep. 25, 85 (1972).


4 M. C. Ruzicka, “On bubbles rising in line,” Int. J. Multiphase Flow 26, 1141 (2000).
5 T. Miyahara, S. Kaseno, and T. Takahashi, “Studies on chains of bubbles rising through quiescent liquid,” Can. J. Chem.

Eng. 62, 186 (1984).


6 S. A. Socolofsky and E. E. Adams, “Liquid volume fluxes in stratified multiphase plumes,” J. Hydraul. Eng. 129, 905 (2003).
7 S. A. Socolofsky and E. E. Adams, “Role of slip velocity in the behavior of stratified multiphase plumes,” J. Hydraul. Eng.

131, 273 (2005).


8 S. A. Socolofsky, T. Bhaumik, and D. G. Seol, “Double-plume integral models for near-field mixing in multiphase plumes,”

J. Hydraul. Eng. 134, 772 (2008).


9 S. A. Socolofsky, B. C. Crounse, and E. E. Adams, Multi-Phase Plumes in Uniform, Stratified, and Flowing Environments

(American Society of Civil Engineers, Reston, VA, 2002).


10 G. Bellani and E. A. Variano, “Slip velocity of large neutrally buoyant particles in turbulent flows,” New J. Phys. 14, 125009

(2012).
103301-20 B. Wang and S. A. Socolofsky Phys. Fluids 27, 103301 (2015)

11 P. Bagchi and S. Balachandar, “Effect of turbulence on the drag and lift of a particle,” Phys. Fluids 15, 3496 (2003).
12 T. Asaeda and J. Imberger, “Structure of bubble plumes in linearly stratified environments,” J. Fluid Mech. 249, 35 (1993).
13 I. Leifer, “Characteristics and scaling of bubble plumes from marine hydrocarbon seepage in the coal oil point seep field,”

J. Geophys. Res.: Oceans 115, C11014, doi:10.1029/2009JC005844 (2010).


14 H. Sahling, G. Bohrmann, Y. G. Artemov, A. Bahr, M. Bruning, S. A. Klapp, I. Klaucke, E. Kozlova, A. Nikolovska, T.

Pape, A. Reitz, and K. Wallmann, “Vodyanitskii mud volcano, Sorokin trough, Black Sea: Geological characterization and
quantification of gas bubble streams,” Mar. Petrol. Geol. 26, 1799 (2009).
15 M. Romer, H. Sahling, T. Pape, G. Bohrmann, and V. Spiess, “Quantification of gas bubble emissions from subma-

rine hydrocarbon seeps at the Makran continental margin (offshore Pakistan),” J. Geophys. Res.: Oceans 117, C10015,
doi:10.1029/2011jc007424 (2012).
16 D. F. McGinnis, J. Greinert, Y. Artemov, S. E. Beaubien, and A. Wuest, “Fate of rising methane bubbles in stratified waters:

How much methane reaches the atmosphere?,” J. Geophys. Res.: Oceans 111, 15, doi:10.1029/2005JC003183 (2006).
17 G. Rehder, I. Leifer, P. G. Brewer, G. Friederich, and E. T. Peltzer, “Controls on methane bubble dissolution inside and

outside the hydrate stability field from open ocean field experiments and numerical modeling,” Mar. Chem. 114, 19 (2009).
18 L. Hu, S. A. Yvon-Lewis, J. D. Kessler, and I. R. MacDonald, “Methane fluxes to the atmosphere from deepwater hydro-

carbon seeps in the northern Gulf of Mexico,” J. Geophys. Res.: Oceans 117, C01009, doi:10.1029/2011JC007208 (2012).
19 I. R. MacDonald, V. L. Asper, J. P. Chanton, O. Garcia-Pineda, L. Hu, M. Kastner, I. Leifer, T. Naehr, E. Solomon, and S.

Yvon-Lewis, “Remote sensing and sea-truth measurements of methane flux to the atmosphere (HYFLUX project),” National
Technical Information Service (NTIS) Accession Number: DE2012-1041004.
20 J. F. Harper, “On bubbles rising in line at large Reynolds numbers,” J. Fluid Mech. 41, 751 (1970).
21 W. W. Martin and G. M. Chandler, “The local measurement of the size and velocity of bubbles rising in liquids,” Appl. Sci.

Res. 38, 239 (1982).


22 H. Yuan and A. Prosperetti, “On the in-line motion of 2 spherical bubbles in a viscous-fluid,” J. Fluid Mech. 278, 325 (1994).
23 J. Katz and C. Meneveau, “Wake-induced relative motion of bubbles rising in line,” Int. J. Multiphase Flow 22, 239 (1996).
24 G. P. Celata, M. Cumo, F. D’Annibale, and A. Tomiyama, “The wake effect on bubble rising velocity in one-component

systems,” Int. J. Multiphase Flow 30, 939 (2004).


25 A. Tomiyama, G. P. Celata, S. Hosokawa, and S. Yoshida, “Terminal velocity of single bubbles in surface tension force

dominant regime,” Int. J. Multiphase Flow 28, 1497 (2002).


26 C. Brucker, “Structure and dynamics of the wake of bubbles and its relevance for bubble interaction,” Phys. Fluids 11, 1781

(1999).
27 K. Ellingsen and F. Risso, “On the rise of an ellipsoidal bubble in water: Oscillatory paths and liquid-induced velocity,” J.

Fluid Mech. 440, 235 (2001).


28 T. Sanada, M. Watanabe, T. Fukano, and A. Kariyasaki, “Behavior of a single coherent gas bubble chain and surrounding

liquid jet flow structure,” Chem. Eng. Sci. 60, 4886 (2005).
29 H. Schlichting, Boundary-Layer Theory (McGraw-Hill, 1979), pp. 729–757.
30 P. G. Sffman, “On the rise of small air bubbles in water,” J. Fluid Mech. 1, 249 (1956).
31 S. A. Socolofsky and E. E. Adams, “Multi-phase plumes in uniform and stratified crossflow,” J. Hydraul. Res. 40, 661

(2002).
32 R. Clift, J. R. Grace, and M. E. Weber, Bubbles, Drops, and Particles (Academic Press, New York, 1978).
33 H. Tsuge and S. Hibino, “The motion of single gas bubbles rising in various liquids,” Kagaku Kogaku 35, 65 (1971).
34 S. Takagi and Y. Matsumoto, “Surfactant effects on bubble motion and bubbly flows,” Annu. Rev. Fluid Mech. 43, 615

(2011).
35 M. M. Wu and M. Gharib, “Experimental studies on the shape and path of small air bubbles rising in clean water,” Phys.

Fluids 14, L49 (2002).


36 G. Mougin and J. Magnaudet, “Path instability of a rising bubble,” Phys. Rev. Lett. 88, 014502 (2002).
37 A. W. G. de Vries, A. Biesheuvel, and L. van Wijngaarden, “Notes on the path and wake of a gas bubble rising in pure

water,” Int. J. Multiphase Flow 28, 1823 (2002).


38 B. Wang and S. A. Socolofsky, “A deep-sea, high-speed stereoscopic imaging system for in situ measurement of natural

seep bubble and droplet characteristics,” Deep Sea Res., Part I 104, 134 (2015).
39 N. M. Aybers and A. Tapucu, “The motion of gas bubbles rising through stagnant liquid,” Wärme- und Stoffübertragung 2,

118 (1969).
40 P. Di Marco, W. Grassi, and G. Memoli, “Experimental study on rising velocity of nitrogen bubbles in FC-72,” Int. J. Thermal

Sci. 42, 435 (2003).


41 P. Bradshaw, R. B. Dean, and D. M. Mceligot, “Calculation of interacting turbulent shear layers: Duct flow,” J. Fluids Eng.

95, 214 (1973).


42 Y. Zhou, R. M. C. So, M. H. Liu, and H. J. Zhang, “Complex turbulent wakes generated by two and three side-by-side

cylinders,” Int. J. Heat Fluid Flow 21, 125 (2000).


43 J. B. W. Kok, “Dynamics of a pair of gas bubbles moving through liquid. Part II. Experiment,” Eur. J. Mech. B/Fluids 12,

541 (1993).
44 B. Wang and S. A. Socolofsky, Camera images of laboratory experiments of weak bubble streams in stagnant and crossflow

conditions for experiments conducted in the Fluid Dynamics Laboratory at Texas A&M University, Harte Research Institute,
Texas A&M University-Corpus Christi, 2015.

Potrebbero piacerti anche