You are on page 1of 18

FLOW AND DRAG PHENOMENA OF SPHEROID BUBBLES IN

SHEAR-THINNING POWER-LAW LIQUIDS

Anjani Ravi Kiran Gollakota, Nanda Kishore*


Department of Chemical Engineering, Indian Institute of Technology Guwahati, Assam, India 781039

Effects of the Reynolds number, aspect ratio of non-spherical bubble, and power-law index of the shearthinning liquids on the flow and drag behavior of non-spherical bubbles are elucidated using a
computational fluid dynamics based numerical solver, ANSYS Fluent 14. The solution methodology is
extensively benchmarked via detailed domain and grid independence study and by comparing present
results of spherical bubbles by their literature counterparts. Further extensive new results obtained over
wide range of pertinent conditions as: Reynolds number Re: 1 200; bubble aspect ratio e: 0.52.5 and
power-law index, n: 0.21. The size of the recirculation wake decreases with the decreasing power-law
index, and/or with the decreasing bubble aspect ratio and/or with the decreasing Reynolds number. For
bubbles of aspect ratio e >1, a crossover Reynolds number is observed with respect to the power-law
index, i.e., below the crossover Reynolds number, the drag coefficient of bubble increases with the
decreasing power-law index; whereas, above the crossover Reynolds number, a reverse trend is observed.
Finally, based on the present numerical results, a simple predictive correlation is proposed for total drag
coefficients of non-spherical bubbles rising in Newtonian and power law liquids, which can be used in
new applications.

Keywords: non-spherical bubbles, aspect ratio, drag coefficient, power-law fluid, Reynolds number

1. INTRODUCTION

The rise of bubbles in non-Newtonian fluids is a common phenomenon in many liquid-gas


contacting equipments of chemical, biochemical and processing industries. Some examples
include foaming, sewage sludge treatment, fermentation, polymer devolatalization, nonNewtonian bubble columns, composite processing, etc. The liquid which flows past a bubble
exert forces on the bubble surface, which arise due to pressure non-uniformity, and hence the
_____________________
*Corresponding author, email: nkishore@iitg.ernet.in ; mail2nkishore@gmail.com

bubble undergoes change in its orientation and shape. Furthermore, depending on the size of the
bubble, nature of liquid, and the magnitude of external force acting on them, they attain various
shapes such as ellipsoidal, spheroidal, spherical-cap, ellipsoidal-cap, dimpled, skirts with or
without open wake, wobbling and disk-like, etc. Despite this fact, many gas-liquid contacting
equipments are designed based on the assumption of spherical bubble rising in Newtonian fluids
in a creeping flow regime. On the other hand, many polymeric liquids in chemical and
biochemical industries obey non-Newtonian rheological behavior including power-law,
viscoplastic and viscoelastic fluids (Chhabra and Richardson, 2008). Furthermore, according to
many experimental studies on deformation of bubbles and drops, it is observed that the spheroid
shape is a very good approximation for non-spherical objects (Clift et al., 1978). Thus not only
from real-life applications point of view but also from theoretical stand point, detailed study on
the flow and drag behavior of non-spherical bubbles in non-Newtonian fluids is a prerequisite for
the reliable design of liquid-gas contacting equipments. This information on non-spherical bubble
rise velocity (or drag coefficient) can be conveniently expressed as function of the Reynolds
number, aspect ratio of the bubble and the rheological constants of non-Newtonian fluids.
Therefore, in this work, a predictive correlation for the drag coefficients of spheroid bubbles in
power-law liquids is developed over a wide range of Reynolds number, bubble aspect ratio and
power-law index through numerical investigation.

2. LITERATURE REVIEW
The literature pertaining to the shape and rise velocity of bubbles in viscous fluids is vast and
most of the studies of bubble shapes and their rise behavior are reviewed by Clift et al., (1978),
Soo (1999), and Chhabra (2006); hence only a few of them are presented herein. Davies and
Taylor (1950) experimentally studied the mechanics of the large bubbles rising through extended
liquids and described the shape and rise velocity of air bubbles (of volume between 1.5 cm3 to
200 cm3) in nitrobenzene or water. They observed that the greater part of the upper surface of the
bubble is spherical due to the reason that the pressure over the front of the bubble is same as that
in hydrodynamic flow around a sphere. Further, Peebles and Garber (1953) categorized bubble
shapes into four categories depending on the value of the prevailing Reynolds number: spherical
bubble for Re < 400, oblate spheroids of varying geometric proportion for 400 < Re < 1100,

oblate spheroids of constant geometric proportion for 1100 < Re < 5000 and mushroom shape
with spherical cap for Re > 5000. Astarita and Apuzza (1965) experimentally studied the motion
of gas bubbles in non-Newtonian fluids. The authors observed a discontinuity which is a resultant
of the transition from Stokes regime to Hadamard regime (a change from the rigid to mobile
surface). Karthik et al. (2009) investigated the discontinuity concept of Astarita and Apuzzo
(1965) and Leal and Skoog (1971) in detail and reported that the discontinuity mainly represents
a change over from the rigid interface to the mobile interface regime. Aybers and Tapucu (1969)
mentioned a critical weber number value of 0.62 at which the transition of bubble shape occurs
from spherical to non-spherical. The work of Calderbank et al. (1970) confirmed the findings of
the Astarita and Apuzzo (1965) with additional information regarding the bubble eccentricity.
Margaritis et al. (1999) experimentally studied the bubble rise velocities and drag coefficients of
bubbles in different non-Newtonian polysaccharide solutions and reported that at low values of
the Reynolds number Re < 1.0, the drag curve is similar to that of Hadamard (1911) and
Rybczynski (1911) models, while at Re > 60 the Cd value is constant i.e., 0.95. Dewsbury et al.
(1999) analysed the parameters affecting the motion of bubbles and revealed that light solid
particles behave similar to gas bubbles for a wide range of Reynolds numbers.

Numerous amount of literature pertaining to the prediction of total drag coefficients of rigid
spheres and spherical bubbles is available on the basis of theoretical /numerical studies. A group
of researchers made efforts in order to predict the total drag coefficient theoretically among
which Hadamard (1911) and Rybczynski (1911) provided theoretical framework to predict the
drag coefficient of bubble and drops in Newtonian liquids at Re 0. Moore (1959, 1963, 1965)
provided a significant theoretical contribution on the motion and deformation of the spherical and
non-spherical gas bubbles rising in irrotational viscous fluids and modified the Levichs drag
coefficient by incorporating a deformation parameter in their correlation. Kawase and MooYoung (1985) analytically presented the drag coefficients of bubbles with mobile and immobile
moving interface rising in shear thinning fluids. Later on, Kurose and Komori (1999) performed
three dimensional numerical simulations and summarized that with the increase in the fluid shear
rate, the drag of bubbles increased. Li et al. (2003) numerically investigated that the shape of the
bubble has an effect on the flow, and observed wake at the rear end of inviscid spheroidal bubble
for e = 0.2 and Re > 10 in Newtonian fluids. Dhole et al. (2007) numerically studied the drag on

spherical bubbles rising in power law fluids at intermediate Reynolds numbers. They found that
the drag coefficient increases in the case of shear-thickening fluids whereas, the trend is reverse
in the case of shear-thinning fluids for Re > 5. Kishore et al. (2007) conducted numerical studies
to predict the drag coefficients and Sherwood numbers of swarms of spherical bubbles rising in
power law fluid at moderate Reynolds and Schmidt numbers. They observed that for a fixed
value of the Reynolds number, the drag coefficient decreases with the decreasing volume fraction
of dispersed phase and/or with the decreasing power-law index. Further, Hayashi and Toiyama
(2009) formulated a drag correlation in relation to Astarita and Apuzzo (1965) and Aybers and
Tapucu (1969) for the rise of fluid particles in stagnant liquids at high Reynolds number ranging
between 0.083 Re 200, and Eotvos number in the range of 0.13 o 30. Bozzano and
Dente (2009) reported the variation in the drag coefficient with respect to deformation for a
single gas bubble moving freely in the quiescent continuous phase under the influence of gravity.
Karthik et al. (2009) concluded from their studies that at low shear rates, for a given pressure
gradient the residence time for a non-Newtonian fluid is higher than the Newtonian flow. Thus,
the motion of a single gas bubble in a pool of non-Newtonian liquid is rheologically complex but
pragmatically important topic. Further, on the basis of the preceding discussion it is safe to
conclude that there is a lack of knowledge about the motion of non-spherical bubbles in nonNewtonian fluids. Therefore, the present work is aimed to numerically investigate the flow and
drag characteristics of spheroid bubbles in shear thinning type power-law fluids.

3. PROBLEM STATEMENT AND MATHEMATICAL FORMULATION


The relative motion between bubble and non-Newtonian fluid is numerically established by
locating a non-spherical bubble at an upstream distance Lu from the inlet and at a downstream
distance Ld from the outlet of the tube as shown in Fig. 1 and allowing a power-law liquid to flow
through the tube (of length L = Lu + Ld) at velocity of V. The sphere volume equivalent diameter
of the non-spherical bubble (oblate and/or prolate) is designated as deq. The length of the tube is
considered to be very long so that the entry and exit effects are negligible. Furthermore, in the
present study, unconfined flow condition is established by considering the large values of Lu and

Ld distance; and assuming moving wall boundary condition along the tube wall. The flow is
further assumed to be steady and axisymmetric in the present range of conditions.

Fig. 1. Schematic representation of computational domain for the flow past oblate and prolate bubble

The governing equations for this fluid flow problem are the conservation of the mass and the
momentum as written below:
Continuity equation:
(1)
Momentum equation:
(2)
Extra stress tensor for an incompressible fluid is defined as:
(3)
Viscosity equation for a power-law fluid can be written as:

(4)

I2 is the second variant of the rate of strain tensor which is related to the derivatives of velocity
components and available in any standard text book. The boundary conditions for this problem
are given as follows.

At the inlet, the uniform flow condition is prescribed i.e.,


(5)

Along the surface of bubble, it is assumed that the bubble is shear-free; i.e.,

xy = 0

(6)

Along the imaginary wall, moving wall condition is used so that there is no wall effect, i.e.,
(7)

Along the central axis of the tube, symmetric conditions is used i.e.,
(8)

The fully converged results, obtained by the use of a commercial CFD solver, are post processed
for the streamlines, surface pressure coefficients, and the drag coefficients as a function of
pertinent dimensionless numbers. The non-dimensional numbers appear in this work are:
Reynolds number:

Re

V 2n deqn

(9)

Aspect ratio of the bubble:

eb

(10)

Total drag coefficient:

Cd

2 Fd
V 2 Ap

(11)

4. NUMERICAL METHODOLOGY
The governing conservation equations of mass and momentum along with the boundary

conditions are solved using commercial software based on computational fluid dynamics,
ANSYS Fluent 14, along with the mesh generating software, Gambit. The detailed numerical
methodology is presented elsewhere (Reddy and Kishore 2012); hence only salient features are
presented herein. The momentum equations are discretized using QUICK scheme and SIMPLE
algorithm is chosen. The tolerance for the residuals of continuity and momentum equations are
-8

held at 10 . Once the fully converged steady velocity and pressure fields are obtained, same are
further used to delineate effects of the pertinent dimensionless numbers on the streamlines,
surface pressure coefficients, and the drag coefficients.

4.1.

Domain and Grid Independence

It is mandatory in numerical studies to cross check the numerical artifacts such as the domain and
grid independence. Further, the domain effects are known to be severe at small Reynolds
numbers and for highly shear-thinning fluids, whereas, the grid effects are more prominent at
high Reynolds numbers. A detailed domain independence study is carried out at Re = 0.1 for
extreme values of bubble aspect ratio and power-law indices (Table 1).
Table 1. Domain independence study at Re = 0.1

e = 0.5
Domain

Lu

Ld

Dt/2

n = 0.2

e = 2.5

n = 1.0

n = 0.2

n = 1.0

Cd
D-1

20

40

30

108.11

99.632

610.06

365.08

D-2

40

80

60

109.68

102.99

610.43

352.74

D-3

60

120

90

109.70

104.76

610.19

352.60

D-4

80

160

120

109.67

105.76

610.40

352.31

Although domains D-2, D-3 and D-4 reproduce almost identical results, a moderately large
domain D-3 is chosen for all other computations. For this optimum domain, thorough grid
independence is carried out at Re = 200 for either extreme values of the aspect ratio in highly

shear-thinning fluids of n = 0.2 and presented in Table 2. From this table, it can be seen that all
grids produced almost identical results with the exception of Grid 1. However, finest possible
grids, i.e., Grid 6 (for e < 1) and Grid 7 (for e > 1) are chosen for all other computations.

Table 2. Grid independence study at Re = 200 and n = 0.2

Grid

Lu

Ld

Inlet

Outlet

Bubble

Imaginary
Wall

Cd

(a) e = 0.5, Re = 200


Grid 1

120

240

50

50

240

200

0.0618

Grid 2

120

240

50

50

240

300

0.0602

Grid 3

120

240

50

50

360

200

0.0600

Grid 4

120

240

75

75

240

200

0.0591

Grid 5

120

240

75

75

240

200

0.0600

Grid 6

180

300

125

125

360

200

0.0610

(b) e = 2.5, Re = 200


Grid 1

120

240

50

50

240

300

0.3782

Grid 2

120

240

50

50

240

500

0.3610

Grid 3

120

240

50

50

300

500

0.3612

Grid 4

120

240

50

50

400

500

0.3620

Grid 5

120

240

75

75

400

500

0.3612

Grid 6

120

240

125

125

400

500

0.3624

Grid 7

210

330

75

75

400

500

0.3648

5. RESULTS AND DISCUSSION


5.1.

Validation

Before presenting the new results it is mandatory to validate the accuracy and reliability of the
present numerical solver. Thus, the solver used in the present study is validated extensively for

spherical bubbles in Newtonian fluids. Table 3 compare the present Cd values of unconfined
spherical bubbles in Newtonian liquids; and are found to be in excellent agreement with the
literature values. Table 4 shows a comparison between the present drag coefficients of
unconfined spherical bubbles in shear thinning fluids (n < 1) with those of Dhole et al. (2007) and
here too agreement is satisfactory.

Table.3. Comparison of Cd of unconfined spherical bubbles in Newtonian fluids (n = 1)

Re

Brabston and Keller


(1975)

Ryskin and Leal


(1984)

Mei et
al.
(1994)

Saboni et
al.
(2010)

Kishore
et al.
(2007)

Present

17.59

17.50

17.714

17.344

4.33

4.208

10

2.35

2.43

2.48

2.49

2.392

2.395

20

1.36

1.41

1.43

1.43

1.397

1.379

50

0.67

0.68

0.69

0.666

0.656

100

0.38

0.38

0.39

0.374

0.364

200

0.19

0.22

0.20

0.22

0.205

0.198

Table. 4. Comparison of Cd of a spherical bubble in shear-thinning fluids (n < 1)


Re

n = 0.6

n = 0.8

Dhole et al. (2007)

Present

Dhole et al. (2007)

Present

--

20.759

--

19.022

4.07

4.287

4.053

4.197

10

2.12

2.255

2.343

2.301

20

1.33

1.207

1.252

1.281

50

0.50

0.526

0.571

0.583

100

0.26

0.277

0.310

0.315

200

0.14

0.145

0.170

0.168

Hence these comparisons encourage the solver to extend the studies on the motion of the nonspherical bubbles in power-law fluids. Therefore, in order to delineate the flow drag and
characteristics due to the flow of power-law fluid past unconfined non-spherical bubbles, the
following range of parameters are considered: Re = 1, 5, 10, 20, 50, 100, 200, e = 0.5, 1, 1.5, 2.5
and n = 0.2, 0.4, 0.6, 0.8, 1.0.

5.2.

Streamline Patterns

In general, for spherical bubbles both in Newtonian and power-law liquids, there is no
recirculation wake formation for any value of the Reynolds number because of imposed shear
free boundary condition on the surface of the bubble. However, in the present case of spheroid
bubbles, the recirculation wake formation is possible in the rear of the bubble because of the
radius of curvature of the prolate bubbles as shown in Figs. 2 and 3. Fig. 2 shows the streamline
patterns around spheroid bubbles of aspect ratio e = 0.5, 1.5, 2.5 in shear thinning fluids of
power-law index n = 0.6 (Fig. 2 (A-C)) and n = 0.4 (Fig. 2 (D-F)) at Re = 20 and Re = 200. As
the flow is assumed to be axisymmetric in this work, the simulations were carried out in half
domain. Thus the flow patterns are also shown in half domain only, i.e., above the horizontal
central line (upper half) flow patterns are shown at Re = 20 and below this central line (lower
half), patterns are shown at Re = 200. By this presentation, one can compare two flow patterns for
two different values of the Reynolds number for any combinations of the bubble aspect ratio and
the power-law index. From this figure, it is observed that for oblate bubbles of aspect ratio e =
0.5 (Fig. 2(A, D)) the streamline patterns resembles symmetric both fore and aft the bubble for
both values of the Reynolds number and the power law index without any recirculation wake. On
the other hand, in the case of prolate bubbles with aspect ratio e = 1.5 (Fig. 2 (B, E)), a tiny
recirculation wake is observed at both values of the Reynolds number and power-law index.
However, for the case of prolate bubbles of aspect ratio, e = 2.5 (Fig. 2(C, F)), a recirculation
wake of substantially large size forms in the rear half of the bubble for both values of the powerlaw index and the Reynolds number of 200. Similarly, Fig. 3 represents the flow patterns around
spheroidal bubbles of aspect ratio ranging between 0.5 < e < 2.5 at Re = 10 (upper half) and Re =
100 (lower half) for the power law indices of n = 0.2 and n = 0.8; and qualitatively similar

observations can be made as in the case of Fig. 2. However, by comparing the streamline profiles
in Figs. 2 and 3, it can be observed that the size of the recirculation decreases with the decreasing
power-law index, and/or with the decreasing bubble aspect ratio, and/or with the decreasing
Reynolds numbers.

Fig. 2. Streamlines around spheroid bubbles at Re = 20 (upper half) and Re = 200 (lower half) in
power law fluids: (A-C) n = 0.6, (D-F) n = 0.4

Fig. 3. Streamlines around spheroid bubbles at Re = 10 (upper half) and Re = 100 (lower half) in
power law fluids: (A-C) n = 0.8 (D-F) n = 0.2

5.3.

Surface Pressure Coefficient

Fig. 4. Pressure coefficient along the surface of spheroid bubbles in power law fluids at Re = 50

Fig. 4 shows the distribution of pressure coefficient along the surface of spheroid bubbles of
aspect ratio e = 0.5, 1.5 and 2.5 in shear thinning power-law fluids at Re = 50. In particular, Fig. 4
elucidates the combined effects of the aspect ratio and the power law index on the surface
pressure coefficients of the non-spherical bubbles for a given value of the Reynolds number. In
general, the pressure coefficient is maximum at the front stagnation point and it decreases along
the curved surface up to almost equator of curved object (depending on the value of Reynolds
number, aspect ratio and power-law index); and finally starts increasing from equator to the point
of recirculation where the slope of curve changes and rises to some finite value at rear point. It is
known that the pressure recovery will be high for the bubbles or deformed particles when
compared to solid spherical particles due to the low pressure drag. Furthermore, in the present
case, regardless the value of the aspect ratio, the pressure recovery in the rear half of the bubble
decreases as the value of the power law index is decreased. Furthermore, for all values of the
bubble aspect ratio, it is observed that the value of the pressure coefficient at the equator of
bubble is decreasing as the values of the power-law index and/or aspect ratio of bubbles

increases. For a given value of the power law index, as the value of the Reynolds number is
increased the extent of pressure recovery is also increased. In summary, regardless the values of
the aspect ratio and the power law index, the pressure recovery improved at the rear end of the
bubble as the value of the Reynolds number is increased. Further for a fixed value of the
Reynolds number, the pressure recovery increases with the decreasing bubble aspect ratio and/or
with the decreasing power-law index.

5.4.

Drag Coefficient

Fig. 5 shows the influence of the power-law index and the Reynolds number on the total drag
coefficient of spheroid bubbles of different aspect ratio. In general, for any combination of
Reynolds number and the power law index, the value of drag coefficient of bubble is less than
their counterparts of solid particles because of surface mobility. For a fixed combination of the
Reynolds number and the power-law index, the total drag coefficient increases with the
increasing aspect ratio of the spheroid bubbles. The characteristic drag curve i.e., the decreasing

Fig. 5. Total drag coefficient of unconfined non-spherical bubbles in Newtonian and shear-thinning fluids

drag coefficient with the increasing Reynolds number has been recovered for all values of the
power-law index and the bubble aspect ratio. For oblate bubbles of aspect ratio e = 0.5, regardless
the values of the Reynolds number, the drag coefficient decreases with the decreasing power-law
index shown in Figure 5A. However, for bubbles of aspect ratio e 1, there is a cross over
Reynolds number which is strong function of the power-law index and the bubble aspect ratio.
Below this crossover Reynolds number, the drag coefficient increases with the decreasing powerlaw index whereas the reverse trend is observed for Reynolds number greater than the crossover
Reynolds number. Furthermore, the value of this crossover Reynolds number is found to increase
with the increasing bubble aspect ratio.

Fig. 6. Parity plot of simulations and correlated drag coefficient values.

Finally, from the engineering applications standpoint, it is useful to develop a correlation based
on the present numerical results which can be used to estimate the drag coefficient of spheroid
bubbles while they rise in shear thinning fluids. The following form is found to be satisfactory to
correlate the present numerical results in the range of 1 Re 200; 0.5 e 2.5 and 0.2 n 1:
(12)

The above correlation reproduces the present numerical results (176 points) with an average error
of 7.32% which rises to maximum of 48.29% for extreme cases of the pertinent dimensionless
parameters. Figure 6 shows the parity plot of the correlation Cd values by comparing them with
numerical Cd values and it can be seen that most of the points are falling along the diagonal line.

6. CONCLUSIONS
The hydrodynamics of steady shear-thinning power-law fluid flow over unconfined spheroid
bubbles of aspect ratio ranging between 0.5 < e < 2.5, the Reynolds number ranging between 1 <
Re < 200 and power law index ranging between 0.2 < n < 1 is numerically investigated. For
oblate bubbles of e < 1, the streamline contours exhibit fore and aft symmetry irrespective of the
Reynolds number and power law index. On the other hand, for prolate spheroid bubbles of (e >
1), recirculation wakes observed even for small Reynolds numbers. The size of the recirculation
wake increases with the increasing bubble aspect ratio and/or increasing Reynolds number and/or
with the increasing power-law index. The pressure recovery is found increase with the decreasing
aspect ratio and/or with the decreasing power-law index. Cd of an oblate bubble e < 1 is found to
decrease with the increasing Re and/or with the decreasing n. However, for prolate bubbles, a
cross over Reynolds number is found below which the Cd increases with the decreasing powerlaw index and above which the reverse trend is observed. This crossover Reynolds number
slightly increases with the increasing aspect ratio of the spheroidal bubble. Finally on the basis of
present numerical results, a simple correlation is developed which can be used to estimate Cd of
spheroid bubbles in power-law fluids in real life liquid-gas contacting systems provided the
liquid phase obeys power-law non-Newtonian behavior.

SYMBOLS
Ap

Project area, m2

Cd

Drag coefficient

deq

Sphere volume equivalent diameter, m

Aspect ratio

Fd

Drag force, kgm/s2

Length of the tube, m

Power law consistency index, Pasn

Power law behavior index

Pressure, Pa

Re

Reynolds number

Velocity, m/s

Greek symbols

Strain tensor, s-1

Apparent viscosity, Pas

Density, kg/m3

Extra stress tensor, Pa

Subscripts
d

Downstream

Upstream

REFERENCES
Astarita, G., Apuzzo, G., 1965. Motion of gas bubbles in non-Newtonian liquids. AIChE J.,
11(5), 815-820. doi: 10.1002/aic.690110514
Aybers, N. M., Tapucu, A., 1969. The motion of gas bubbles rising through stagnant liquid.
Wrme - und Stoffbertragung, 2(2), 118-128. doi: 10.1007/BF01089056
Bozzano, G., Dente, M., 2009. Single bubble and drop motion modeling. 9th International
conference on chemical and Process Engineering. Rome, 10-13 May 2009,53-60.
Brabston, D. C., Keller, H. B., 1975. Viscous flows past spherical gas bubbles. J. Fluid Mech.,
69(01), 179-189. doi: 10.1017/S0022112075001371
Chhabra R.P., Richardson J.F., 2008. Non-Newtonian Flow and Applied Rheology. 2nd edition,
Elsevier, USA.
Chhabra R.P., 2006. Bubbles, Drops and Particles in Non-Newtonian fluids. 2nd edition, CRC
Press, USA.
Clift R., Grace J.R., Weber M.E., 1978. Bubbles, Drops and Particles. 1st edition, Academic
Press, New York.

Davies, R. M., Taylor, G., 1950. The mechanics of large bubbles rising through extended liquids
and through liquids in tubes. Proc. Royal Soc. London, Series A. Math. Phy. Sci., 200(1062),
375-390. doi: 10.1098/rspa.1950.0023
Dewsbury, K., Karamanev, D., Margaritis, A., 1999. Hydrodynamic characteristics of free rise of
light solid particles and gas bubbles in non-Newtonian liquids. Chem. Eng. Sci., 54(21),
4825-4830. doi: 10.1016/S0009-2509(99)00200-6
Dhole, S. D., Chhabra, R. P., Eswaran, V., 2007. Drag of a Spherical Bubble Rising in Power
Law Fluids at Intermediate Reynolds Numbers. Ind. Eng. Chem. Res., 46(3), 939-946. doi:
10.1021/ie0610086
Hadamard, J., 1911. Mouvement permanent lent dune sphere liquide et visqueuse dans un
liquide visqueux. CR Acad. Sci., 152(25), 1735-1738.
Hayashi, K., Tomiyama, A., 2009. A drag correlation of fluid particles rising through stagnant
liquids in vertical pipes at intermediate Reynolds numbers. Chem. Eng. Sci., 64(12), 30193028. doi: 10.1016/j.ces.2009.03.013
Kawase, Y., Moo-Young, M., 1985. Approximate solutions for drag coefficient of bubbles
moving

in

shear-thinning

elastic

fluids.

Rheol

.Acta,

24(2),

202-206.

doi:

10.1007/Bf01333248
Kishore, N., Chhabra, R., Eswaran, V., 2007. Drag on a single fluid sphere translating in powerlaw liquids at moderate Reynolds numbers. Chem. Eng. Sci., 62(9), 2422-2434. doi:
10.1016/j.ces.2007.01.057
Kishore, N., Chhabra, R.P., Eswaran, V., 2008. Bubble swarms in power-law liquids at moderate
Reynolds numbers: Drag and mass transfer. Chem. Eng. Res. Des., 86(1), 39-53. doi:
10.1016/j.cherd.2007.10.009
Kurose, R., Komori, S., 1999. Drag and lift forces on a rotating sphere in a linear shear flow. J.
Fluid Mech., 384, 183-206. doi: 10.1017/S0022112099004164
Leal, L. G., Skoog, J., Acrivos, A., 1971. On the motion of gas bubbles in a viscoelastic liquid.
Can. J. Chem. Eng., 49(5), 569-575. doi: 10.1002/cjce.5450490504
Li, W. Z., Yan, Y. Y., Smith, J. M., 2003. A numerical study of the interfacial transport
characteristics outside spheroidal bubbles and solids. Int. J. Multiphase Flow, 29(3), 435-460.
doi: 10.1016/S0301-9322(02)00163-5

Margaritis, A., Bokkel, D. W., Karamanev, D. G., 1999. Bubble rise velocities and drag
coefficients in non-Newtonian polysaccharide solutions. Biotechnol. Bioeng., 64(3), 257-266.
doi: 10.1002/(SICI)1097-0290(19990805)64:3<257::AID-BIT1>3.0.CO;2-F
Mei, R., Klausner, J. F., Lawrence, C. J., 1994. A note on the history force on a spherical bubble
at finite Reynolds number. Phy. Fluids, 6(1), 418-420. doi: 10.1063/1.868039
Moore, D. W., 1959. The rise of a gas bubble in a viscous liquid. J. Fluid Mech., 6(01), 113-130.
doi: 10.1017/S0022112059000520
Moore, D. W., 1963. The boundary layer on a spherical gas bubble. J. Fluid Mech., 16(02), 161176. doi: 10.1017/S0022112063000665
Moore, D. W., 1965. The velocity of rise of distorted gas bubbles in a liquid of small viscosity. J.
Fluid Mech., 23(04), 749-766. doi:10.1017/S0022112065001660
Mukundakrishnan, K., Eckmann, D. M., Ayyaswamy, P. S., 2009. Bubble motion through a
generalized power-law fluid flowing in a vertical tube. Ann. N. Y. Acad. Sci., 1161(1), 256267. doi: 10.1111/j.1749-6632.2009.04089.x
O'Brien, M. P., Gosline, J. E., 1935. Velocity of large bubbles in vertical tubes. Ind. Eng. Chem.,
27(12), 1436-1440. doi: 10.1021/ie50312a013
Peebles, F. N., Garber, H. J., 1953. Studies on the motion of gas bubbles in liquids. Chem. Eng.
Progr., 49(2), 88-97.
Reddy, C.R., Kishore, N., 2012. Wall retardation effects on flow and drag phenomena of
confined spherical particles in shear-thickening fluids. Ind. Eng. Chem. Res., 51(51), 1675516762. doi: 10.1021/ie302707s
Rosenberg, B., 1950. The drag and shape of air bubbles moving in liquids: David W. Taylor
Model basin, USA. doi: 10.5962/bhl.title.47807
Rybczynski, W., 1911. Uber die fortschreitende Bewegung einer flussigen Kugel in einem zahen
Medium. Bull. Int. Acad. Sci. Cracovie, 1, 40-46.
Ryskin, G., Leal, L. G., 1984. Numerical solution of free-boundary problems in fluid mechanics.
Part 2. Buoyancy-driven motion of a gas bubble through a quiescent liquid. J. Fluid Mech.,
148, 19-35. doi: 10.1017/S0022112084002226
Saboni, A., Alexandrova, S., Mory, M., 2010. Flow around a contaminated fluid sphere. Int. J.
Multiphase Flow, 36(6), 503-512. doi: 10.1016/j.ijmultiphaseflow.2010.01.009
Soo S.L., 1999. Two Phase Flow in Complex Systems. 1st edition, John Wiley and Sons, USA.

You might also like