Numerical Simulation of 3D Bubbles Rising in Viscous Liquids
using a Front Tracking Method
Jinsong Hua
a
, Jan F. Stene
b
and Ping Lin
b
a
Institute of High Performance Computing, 1 Science Park Road, #01-01 The Capricorn,
Singapore 117528
b
Department of Mathematics, National University of Singapore, 2 Science Drive 2,
Singapore 117543
Submitted for Publication to
Journal of Computational Physics
May 2007
Manuscript correspondence should be directed to Dr. Jinsong Hua.
Postal address: Institute of High Performance Computing
1 Science Park Road, #01-01 The Capricorn
Singapore 117528
Email address: [email protected].edu.sg
Telephone number: +65-6419 1537
Fax number: +65-6778 0522
1
Numerical Simulation of 3D Bubbles Rising in Viscous Liquids
using a Front Tracking Method
Jinsong Hua
a
, Jan F. Stene
b
and Ping Lin
b
a
Institute of High Performance Computing, 1 Science Park Road, #01-01 The Capricorn,
Singapore 117528
b
Department of Mathematics, National University of Singapore, 2 Science Drive 2,
Singapore 117543
Abstract
The rise of bubbles in viscous liquids is not only a very common process in many
industrial applications, but also an important fundamental problem in fluid physics. An
improved numerical algorithm based on the front tracking method, originally proposed by
Tryggvason and his coworkers, has been validated against experiments over a wide range
of intermediate Reynolds and Bond numbers using an axisymmetric model (Hua and
Lou, J. Comput. Phys. 222:769-795, 2007). In the current paper, this numerical algorithm
is further extended to simulate 3D bubbles rising in viscous liquids with high Reynolds
and Bond numbers and large density and viscosity ratios at the physical order of the
typical multi-fluid system of air bubbles in water. To facilitate the simulation, mesh
adaptation is implemented for both the front mesh and the background mesh, and the
governing Navier-Stokes equations for incompressible, Newtonian flow are solved in a
moving reference frame attached to the rising bubble. Specifically, the flow equations are
solved using a finite volume scheme based on the Semi-Implicit Method for Pressure-
Linked Equations (SIMPLE) algorithm, and it appears to be robust even in the range of
high Reynolds numbers and high density/viscosity ratios. The 3D bubble surface is
tracked explicitly using an adaptive, unstructured triangular mesh. The model is
integrated with the software package PARAMESH, a block-based adaptive mesh
refinement (AMR) tool developed for parallel computing. PARAMESH allows
background mesh adaptation as well as the solution of the governing equations in parallel
on a supercomputer. The interpolations between the front mesh and the background mesh
are done with Peskin’s distribution function. The current model has also been applied to
Corresponding author, E-mail: [email protected]
2
simulate a number of cases of 3D gas bubbles rising in viscous liquids, e.g. air bubbles
rising in water. The simulation results are compared with experimental observations both
in aspect of terminal bubble shapes and terminal bubble velocities. In addition, we
applied this model to simulate the interaction between two bubbles rising in a liquid. The
simulation results provide us with more physical insights into the complex bubble rising
behavior in viscous liquids.
Keywords: Computational fluid dynamics; Incompressible flow; Multiphase flow; Bubble
rising; SIMPLE algorithm; Front tracking method; Adaptive mesh refinement; Moving
reference frame
3
1. Introduction
Multiphase flows are numerous in both everyday life and engineering practice [34].
Typical examples in nature include raindrops in air and gas bubbles in water, whereas
chemical reactions, combustion and petroleum refining are examples of multiphase flows
in industry. One very basic example of such flow is the rise of a single gas bubble in an
otherwise quiescent viscous liquid. The understanding of the flow dynamics of this
system is of great importance in engineering applications and to the fundamental
understanding of multiphase flow physics. Rising bubbles have long been studied
theoretically [8, 25], experimentally [1] as well as computationally through numerical
modeling [36]. While all these efforts have provided us with valuable insights into the
dynamics of bubbles rising in viscous liquids, there are still many questions that remain
unanswered due to the involvement of complex physics. The behavior of a bubble rising
in a viscous liquid is not only affected by the physical properties such as density and
viscosity of both phases [6], but also by the surface tension on the interface between the
two phases and by the bubble shape evolution [27, 2]. The difficulties in describing and
modeling the complex behavior of a rising bubble are to a large extent due to the strong
nonlinear coupling of factors such as buoyancy, surface tension, bubble/liquid
momentum inertia, viscosity, bubble shape evolution and rise history of the bubble. In
addition, the physics of the behavior of bubbles is of a three-dimensional nature. Hence,
most of the past theoretical works were done with a lot of assumptions, and the results are
only valid for certain flow regimes [25, 43]. The experimental works were limited by the
available technologies to monitor, probe and sense the moving bubbles without
interfering with their physics [1, 38, 44].
With the rapid advance of computing power and the development of robust numerical
methods, first principle based numerical simulations promise great potential in extending
our knowledge of the fundamental system of a single bubble rising in a viscous liquid.
However, there are still great challenges and difficulties in simulating such a system
accurately. This may be attributed to the following facts: (i) the sharp interface between
the gas bubble and the surrounding liquid should be tracked accurately without
introducing excessive numerical smearing; (ii) the surface tension gives rise to a singular
source term in the governing equations, leading to a sharp pressure jump across the
4
interface; (iii) the discontinuity of the density and viscosity across the fluid interface may
lead to numerical instability, especially when the jumps in these properties are high. For
example, the density ratio of liquid to gas could be as high as 1000; (iv) the geometric
complexity caused by bubble deformation and possible topological change is the main
difficulty in handling the geometry of interface; a large bubble may break up into several
small ones, and a bubble may also merge with other bubbles; (v) the complex physics on
the interface, e.g. the effects of surfactants, film boiling and phase change (heat and mass
transfer) and chemical reactions. Fortunately, various methods for multiphase flow have
been developed to address these difficulties, and each method typically has its own
characteristic strengths and weaknesses. Comprehensive reviews of numerical methods
for multiphase/interfacial flow simulation have been given by Scardovelli and Zaleski
[37] and Annaland et al. [41]. Most of the current numerical techniques applied in the
simulation of multiphase/interfacial flows have been developed with focus on the
following two aspects: (i) capturing/tracking the sharp interface, e.g. interface capturing,
grid fitting, front tracking or hybrid methods; and (ii) stabilizing the flow solver to handle
discontinuous fluid properties and highly singular interfacial source terms, e.g. the
projection-correction method [41] and the SIMPLE algorithm [6, 17].
The volume of fluid [14, 4], level-set [28, 42, 29] and phase-field [19] approaches fall
into the first category of front capturing methods. In these methods the interface is
captured using various volume functions defined on the grid used to solve the “one-fluid”
formulation of the governing equations for multiphase flow. Since interface capturing
uses the same grid as the flow solver, it is relatively easy to implement. However, the
accuracy of this approach is limited by the numerical diffusion from the solution of the
convection equation of the volume function. Various schemes have been developed to
advect, reconstruct / reinitialize the volume function to improve the accuracy in
calculating the interface position. One example is the high-order shock-capturing scheme
used to treat the convective terms in the governing equations [18]. Although the explicit
reconstruction of the interface is circumvented, the implementation of such high-order
schemes is quite sophisticated, and they do not work well for the sharp discontinuities
encountered in multiphase/interfacial flows. In addition, a relatively fine grid is needed in
the vicinity of the interface to obtain good resolution.
5
The second category of approaches tries to track the moving interface by fitting the
background grid points to the interface. The fitting is achieved through re-meshing
techniques such as deforming, moving, and adapting the background grid points. This
method is also well-known as “boundary-fitting approach”, and the “boundary” here
refers to the interface between the fluids. The grid-fitting approach is capable of
capturing the interface position accurately. Early development on this approach was done
by Ryskin and Leal [36]. Curvilinear grids were used to follow the motion of a rising
bubble in liquid. This method is suitable for relatively simple geometries undergoing
small deformations, and applications to complex, fully three-dimensional problems with
unsteady deforming phase boundaries are very rare. This is mainly due to difficulties in
maintaining the proper volume mesh quality and in handling complex interface geometry
such as topological change. In spite of these difficulties, recent work by Hu et al. [16]
showed some very impressive results on 3D simulations of moving spherical particles in
liquid.
The third category is the front tracking method. This approach solves the flow field
on a fixed grid and tracks the interface position in a Lagrangian manner by a set of
interface markers. These interface markers can be free particles without connection, or
they can be logically connected elements, possibly containing accurate geometric
information about the interface such as area, volume, curvature, deformation, etc. A front
tracking technique was proposed by pioneer researchers Glimm and his coworkers [11,
12, 13]. They represent the front interface using a set of moving markers and solve the
flow field on a separate background grid. The background grid is modified only near the
front to make background grid points coincide with the front markers of the interface. In
this case, some irregular grids are reconstructed and special finite difference stencils are
created for the flow solver, increasing the complexity of the method and making it more
difficult to implement. Independently, another front tracking technique was developed by
Peskin and collaborators [31, 10]. In their method, the interface is represented by a
connected set of particles which carry forces, either imposed externally or adjusted to
achieve a specific velocity at the interface. A fixed background grid is kept unchanged
even near the front interface, and the interface forces are distributed onto the background
to solve the “one-fluid” formulation of the fluid flow.
6
A number of combinations and improvements of these basic approaches have been
proposed to enhance the capabilities in dealing with the sharp, moving interface, where
complex physical phenomena and processes could occur. One of the most promising
approaches is arguably the front tracking method proposed by Tryggvason and his
collaborators [46, 45]. Actually, this method may be viewed as a hybrid of the front
capturing and the front-tracking techniques: a fixed background grid is used to solve the
fluid flow, while a separate interface mesh is used to track the interface position
explicitly. The tracked interface carries any jumps in the fluid properties, such as density
and viscosity, and any interfacial forces, such as surface tension. Fluid properties are then
distributed onto the fixed background grid according to the position of the interface. The
surface tension can be calculated according to the geometry of the interface and is also
distributed onto the background grid in the vicinity of the interface.
Besides the numerical techniques employed to capture/track the moving interface, it
is also very important to develop a stable numerical method to solve the governing
equations of the flow field. Some investigators have considered simplified models such
as Stokes flow [33], where inertia is completely ignored, and inviscid potential flow [15],
where viscous effects are ignored in. In both cases, the motion of deformable boundaries
can be simulated with boundary integral techniques. However, when considering the
transient Navier-Stokes equations for incompressible, Newtonian fluid flow, the so-called
“one-fluid” formulation for multiphase flow has proved most successful [4, 42, 46].
Popular modern methods that use the “one-fluid” formulation include the projection-
correction method [45, 41] and the SIMPLE algorithm [6, 17]. Various
multiphase/interfacial flow problems have been successfully simulated by the front
tracking method [45] with a projection-correction flow solver. It appears that previously
reported results have been limited to flows with low to intermediate Reynolds numbers
(<100) and small density ratios (<100) [5]. It is thus natural to re-examine the approach
and to make it more robust and applicable to wider flow regimes. Some revised versions
of the project-correction method have been proposed to improve its capability in handling
situations with large density and viscosity ratios [41]. Recently, Hua and Lou [17] tested
a SIMPLE-based algorithm to solve the incompressible Navier-Stokes equations. The
simulation results indicated that the newly proposed method could robustly solve the
7
Navier-Stokes equations with large density ratios up to 1000 and large viscosity ratios up
to 500.
Hua and Lou [17] presented extensive simulations and model validation on a single
bubble rising in a quiescent liquid. The comprehensive simulations show good results in
wide flow regimes with high density and viscosity ratios, and the algorithm is as such
promising in the direct numerical simulation of multiphase flow. Unfortunately, the
previous validation studies were limited to the 2D axisymmetric model where fluid flows
and bubble shapes are axisymmetric. Hence, it would be interesting to investigate the
robustness of the proposed numerical approach for multiphase flow in flow regimes of
higher Reynolds and Bond numbers where the bubble may not be axisymmetric anymore.
Therefore, a fully three-dimensional modeling approach is proposed in this paper. In
addition, other features such as mesh adaptation, moving reference frame and parallel
programming are introduced to enhance the model capability in simulating the rise of a
3D bubble in a viscous liquid.
The numerical algorithm proposed by Tryggvason and co-workers [46, 45], and
extended further by Hua and Lou [17], is adopted in this paper. The handling of the
moving interface in this method may be characterized as a hybrid of interface tracking
and capturing. The governing Navier-Stokes equations are solved on a fixed Cartesian
grid with an adaptive block structure, while the interface is represented by a set of
explicitly tracked front markers. These markers form an adaptive triangular surface mesh
that is advected with a velocity interpolated from the surrounding fluid. An illustration of
such a mesh system is shown in Figure 1. A single set of the governing equations are
solved in the entire computational domain by treating the two fluids as one single fluid
with variable fluid properties across the interface – often referred to as the “one-field” or
“one-fluid” approach. The interface is assumed to have a given finite thickness (normally
about two to four times the background grid size) so that jumps in the fluid properties
across the surface can be reconstructed smoothly by solving a Poisson equation.
A parallel adaptive mesh refinement (AMR) tool, PARAMESH [24], is integrated
with the modified SIMPLE flow solver, and the governing equations are solved in a non-
inertial moving reference frame attached to the rising bubble. The AMR feature allows a
relatively high-resolution mesh in the vicinity of the bubble surface. The non-inertial
8
moving reference frame technique translates the computational domain with the rising
bubble, allowing the computational domain to be relatively small and always centered
around the bubble. The latter feature is particularly useful for studying the path instability
of a rising bubble or the interaction of multiple bubbles, which may need very long
simulation periods. For example, it is observed in experiments that the paths of
millimeter-sized, rising air bubbles in water normally stabilize after a rise distance of 50-
100 times the initial bubble diameter. If a stationary frame was to be applied to simulate
this situation, the computational domain would be huge compared to the domain of
interest. Even though an AMR feature is adopted, the total number of grid points can still
be a big burden, slowing down the simulation.
The problem of a single bubble rising in a viscous liquid has been widely used as a
typical validation case for the development of new numerical methods for multiphase
flow [6, 40, 41]. Due to numerous experimental and numerical studies in the past, the
physical understanding of the bubble rise behavior in liquid has been well-established in
some flow regimes, e.g. regimes with lower Reynolds and Bond numbers [7, 1].
However, due to the complexity in multiphase flow physics and the difficulties in both
experiments and simulations, the behavior of a rising bubble with high Reynolds number
is not understood well [26, 21]. In this paper, we first validate our model through
comparing computational results of a single bubble rising with experimental results [1] in
aspects of both bubble shapes and terminal velocities. In addition, we apply the 3D model
to simulate air bubbles rising in water, and we compare the terminal velocities of the
bubbles predicted in our simulations with experimental results within a large range of
diameters: from 0.5 mm to 30 mm. There have been few numerical studies on this except
some recent ones [9, 21]. However, since air bubbles rising in water is such a common
process both in our daily life and in industrial applications, a better understanding of this
process is of great importance.
The rest of this paper is organized as follows. In Section 2 we present the governing
equations as well as the numerical method we apply to solve these equations. Numerical
results presented in Section 3 include a detailed sensitivity analysis of the computational
set-up as well as validation through a comparison of our numerical predictions with
available experimental data. Finally, we recapitulate our main findings in Section 4.
9
2. MATHEMATICAL MODEL AND NUMERICAL METHOD
2.1. Governing Equations
The problem of gas bubbles rising in liquids studied in this paper can be described as
an isothermal, multi-fluid system with two incompressible and immiscible Newtonian
fluids. We will use one single set of governing equations for the entire flow domain
where we treat the different fluids as one single fluid with material properties varying
across the interface. With this “one-fluid” approach, there is no need to deal with the
jump conditions across the interface when the governing equations are solved. However,
we will have to calculate the fluid property distributions and include the surface tension
as a singular source term in the solution domain through the use of a delta function before
the equations can be solved.
The mass conservation for the whole domain under the incompressibility condition
may be expressed in form of volume flux conservation,
0
=
u . (1)
The momentum conservation (Navier-Stokes equations) takes the form,
gxxnuuuu
u
)()()]([
)(
T
lf
dsp
t
ρρδσκμρ
ρ
++++−∇=+
Γ
, (2)
where u is the fluid velocity,
ρ
is the fluid density,
l
ρ
is the density of the liquid phase,
p
is the pressure,
μ
is the fluid viscosity,
is the surface tension coefficient,
κ
is the
interface curvature,
n
is the unit normal vector to the interface,
g
is the gravitational
acceleration, and )
f
x(x
δ
is a delta function that is defined as the product of three one-
dimensional delta functions:
)(z)()( yx)(
δ
δ
δ
δ
=
x , ),zy,(x
=
x . The subscript refers to a
point on the interface
Γ . It is worth pointing out that the material properties density
f
ρ
and viscosity
μ
will be discontinuous across the interface, and there will generally be a
jump in the pressure
p
across the interface as well. Note that the surface tension term is
a singular term that only comes into effect on the interface between the two fluids.
We non-dimensionalize the equations by introducing dimensionless characteristic
variables as follows,
10
D
x
x =
*
,
gD
u
u =
*
, t
D
g
=
*
τ
,
l
ρ
ρ
ρ
=
*
,
gD
p
p
l
ρ
=
*
,
l
μ
μ
μ
=
*
,
κ
κ
D=
*
,
g
g
g =
*
,
where is the diameter of a sphere with the same volume as the bubble and
D g=g
.
Thus we may re-express the Navier-Stokes equations as
gxxnuuuu
u
)1()(
1
)]([
1)(
*
T
*
++++−∇=+
Γ
ρδκμρ
ρ
ds
BoRe
p
t
f
, (3)
in which the superscript * has been omitted for convenience. Note that the non-
dimensional Reynolds and Bond numbers used here are thus defined as
l
l
Dg
Re
μ
ρ
2/32/1
*
= and
σ
ρ
2
*
gD
Bo
l
= .
By studying the non-dimensional formulation, it can be noticed that the flow is entirely
characterized by the following four dimensionless parameters: The density and viscosity
ratios of the fluids, the Reynolds number and the Bond number. It is also noted that the
definition of the non-dimensional Reynolds number in the experimental works is
different from the one defined here. Normally, experimental works prefer the following
non-dimensional numbers: Eotvos number (
E
, also known as Bond number); Morton
number (
M
) and Reynolds number ( ), defined as Re
σ
ρ
2
gD
E
l
= ,
3
4
M
σρ
μ
l
l
g
= , and
l
l
DU
μ
ρ
=Re ,
where is the terminal rise velocity of the bubble measured in the experiments.
U
2.2. Treatment of the Discontinuities across the Interface
When the governing Navier-Stokes equations are solved numerically on a fixed grid,
the values of the density and viscosity on these grid points are required. It is a reasonable
assumption that each fluid is incompressible, and fluid properties such as density and
viscosity are constant in each fluid phase. Hence the density and viscosity are physically
discontinuous across the interface between the two immiscible fluids, and this abrupt
jump at grid points adjacent to the interface has traditionally caused great problems in
11
many numerical methods. In Tryggavson et al. [46, 45] a fixed background mesh is
adopted to solve the governing flow equations, and a separated mesh is applied to track
the position of the interface as well as the discontinuities across the front. An illustration
of such a mesh system is shown in Figure 1. The discontinuities across the front are
distributed from the front mesh to the background mesh, and continuous distributions of
the fluid properties on the fixed background mesh can be reconstructed. The singular
source term on the front is distributed to the background grid similarly, and the governing
equations can thus be solved on the fixed background grid using any preferred numerical
approach.
Let us first assume zero interface thickness. Consider the associated reconstruction of
the field distribution of material properties at time
t in the whole domain through
a certain indicator function . Let the indicator function be zero in the liquid phase
and one in the gas phase. We may then write,
),( tb x
I
),(
tx
),()(),(
tIbbbtb
lbl
xx
+
=
, (4)
where is either fluid density or viscosity, and the subscripts
l
and
b
refer to liquid
and gas phase, respectively. Further let
),( tb x
Ω
be the domain of the gas phase and let
Γ
be
the interface between the two phases. The indicator function may then be expressed as
Ω
=
)(
)(),(
t
dtI xxxx
δ
. (5)
Taking the gradient of the indicator function and applying Stokes’ theorem, we get
ΓΩ
=
=
)()(
)()(),(
tt
ddtI sxxnxxxx
δδ
, (6)
where
n is the outer unit normal vector of the interface. Taking the divergence yields a
Poisson equation for the indicator function:
Γ
=
)(
2
)(),(
t
dtI sxxnx
δ
. (7)
We solve this equation and then calculate the distribution of material properties from
equation (4).
Unverdi and Tryggvason [46] addressed the sharp jump in fluid properties across the
interface in their front tracking algorithm. They introduced an artificial thickness of the
interface inside which the material properties vary continuously from one fluid to the
12
other. According to this idea, a distribution function is introduced to approximate
the delta function
)(xD
)(x
δ
with the assumption of an artificial thickness of the interface. We
here adopt the traditional Peskin distribution function [31],
<+
=
=
otherwise. 0
2 if )
2
cos(1)4(
)(
)(
3
1
3
h
h
h
D
ff
i
f
xxxxΠ
xx
π
. (8)
Substituting the Peskin distribution function into equation (7), the indicator function can
be reconstructed by solving the Poisson equation. The resulting indicator function will
then be zero in the pure liquid phase, vary continuously from zero to one in the artificial
thickness region, and one in the gas phase.
Besides the discontinuity in fluid properties across the bubble interface, the surface
tension, , a singular source term on the bubble interface, brings
another great challenge for numerical methods in multiphase flow. In the current study
the net force caused by surface tension on the surface elements is calculated, thus
circumventing the high-order derivatives involved in curvature calculations. Figure 2
shows the surface tension force exerted on a central surface element (E0) by its
neighboring elements (E1, E2 and E3). The surface tension force acting on an edge
shared between the central element and a neighboring element can be calculated by
Γ
= dsF
f
)( xxn
δσκ
σ
3 and 2 1,i )(
0,
=
×
=
iii
ntF
σ
, (9)
where is the vector of edge
i and its unit outer normal. Hence, for a central
element E0, the net force caused by surface tension can be expressed as,
i
t
0,i
n
==
×==
3
1
0,
3
1
,
)(
i
ii
i
iE
ntFF
σ
σ
. (10)
According to equation 10, the net surface tension force on all surface elements can be
calculated and then distributed to the background mesh for solving the momentum
equations:
)()(
,
=
k
kE
XD xFxF
σσ
(11)
where represents the mass centre of the k-th element used to triangulate the bubble
surface.
k
X
13
2.3. Tracking the Moving Interface
With the techniques introduced in Sections 2.1 and 2.2, the governing equations can
be solved on a fixed background grid to obtain the flow field. An adaptive, unstructured
triangular mesh (front markers) is used to represent the interface between the two fluid
phases. Hence, the velocity of the moving front markers can be obtained by interpolation
from the flow field on the background mesh, and then the front mesh points can be
advected in a Lagrangian manner. Thus the front moves with the same velocity as the
surrounding fluid, and the so-called no-slip condition of the interface is satisfied. In this
paper, the interpolation is carried out using the same distribution function as the one used
for the transfer of fluid properties to the background grid:
=
x
xxxuxu )(),(),(
fff
Dtt , (12)
t
n
ff
n
f
Δ+=
+
n1
uxx . (13)
As the front marker points are advected, the mesh size and quality may consequently
change. The resolution of the front mesh has a strong effect on the information exchange
with the fixed background grid, which may eventually affect the accuracy of the
simulation results. Therefore, it is of key importance that the front mesh has a more or
less constant quality and uniform size throughout the duration of the simulation. To
ensure this, the front quality is examined at each time step and adapted when necessary.
In this paper, the resolution of the triangular mesh for the 3D surface of the bubble is
maintained more or less uniform through adaptation as the interface evolves.
2.4. Mesh Adaptation
As two sets of mesh are applied in the current front tracking method, the resolution of
the front mesh and the background mesh near the front plays an important role in
resolving the interfacial physics of the multiphase flow. From physical principles it is
known that fluid particles on the bubble interface will move downwards towards the
bottom of the bubble as the bubble rises. Similarly, the mesh points also move
downwards on the bubble surface in the front tracking method as the bubble rises. As a
result, the mesh on the upper part of the bubble becomes coarser. On the other hand, the
14
mesh at the lower part of the bubble becomes increasingly dense. Figure 3(a) shows the
variation of front mesh quality as the bubble rises in liquid without front mesh adaptation.
It is obvious that the accuracy will be affected when the mesh on the top is too coarse,
and that the dense fine mesh at the bottom of the bubble will consume excessive
computing power without much benefit in accuracy. Thus, the front mesh adaptation as
shown in Figure 3(b) is essential to ensure the accuracy and efficiency of the simulation.
In this aspect, three basic operations are adopted to adapt the front mesh, namely edge
swap, edge split and edge deletion. For long edges, the edge swap operation as shown in
Figure 4(a) is a simple and easy operation to improve the mesh quality. In the edge split
operation as shown in Figure 4(b), a new point is generated by surface fitting of the
existing neighboring mesh points. This new point is then inserted into the two associated
meshes, and new and finer triangles are generated to replace the old ones. Thus edge split
is important to refine the front mesh. For the deletion of short edges as shown in Figure
4(c), the triangles associated with the short edge will be deleted, and the resulted gap will
be sealed through merging the old nodes of the short edge. Hence, edge deletion is
important to coarsen the front mesh. In addition, consistent checking of the mesh
connectivity is also important to ensure the accuracy in calculating the surface tension.
Accordingly, the resolution of the background mesh also plays an important role in
capturing the flow behavior – particularly so in the vicinity of the interface. If the
background mesh resolution is too low, then the detailed flow dynamics will not be
captured reasonably well, resulting in unreliable and inaccurate simulations. Therefore, it
is desirable to have relatively high-resolution grids, particularly near the interface, while
coarse grids may be used away from the interface. This is achieved in our model by the
use of the block-based adaptive mesh refinement (AMR) tool PARAMESH [24]. In this
study, the refining and coarsening of the grid blocks is based on whether there exists a
bubble front within the blocks. An example of the block-wise Cartesian mesh refinement
generated by PARAMESH can be seen in Figure 5. It can be seen that fine background
grids are located in the vicinity of the bubble, while coarser background grids are applied
in regions further away from the bubble front. This feature makes it more efficient to
solve the governing equations and to capture the flow physics near the interface
accurately. An excellent feature of PARAMESH is that the Cartesian grid at all levels of
15
blocks has the same structure. Hence, once the flow solver is developed for one grid
block, it can be easily applied to all levels of blocks. In addition, the different blocks can
be distributed to different CPUs in an MPI parallel environment, which speed up the
problem solving cycle.
2.5. Flow Solver
The numerical method for the governing flow equations is one of the key components
in the simulation. Traditionally, an explicit projection-correction method based on
second-order central differences on a regular, staggered Cartesian grid has been used
along with the front tracking approach [46]. However, it seems that this approach is
unable to handle the large density ratios typical of systems in industrial applications as
well as in nature. In search of a more robust solver, Hua and Lou [17] implemented a
modified version of the classical SIMPLE method [30] for axisymmetric multiphase
flow. Further details about the modified SIMPLE algorithm can be found in the past
works [17, 6]. Computational results indicate that this approach is more robust than the
projection correction method – especially for multiphase flows with large density ratios
such as the air-water system. This improvement is most probably due to the fact that the
SIMPLE algorithm avoids solving the problematic pressure equation directly. Instead, the
pressure and the velocity are corrected iteratively based on the governing equations.
2.6. Moving Reference Frame
In many applications of multiphase flow it is often desirable to study the long-term
behavior and evolution of the moving interface between the fluids. In such applications
the front may move a considerable distance, and the study of the rise path of an air bubble
in water is a typical example. The computational domain must then be correspondingly
large to accommodate such extensive movement. However, in a three-dimensional model
with a high-resolution grid, a large computational domain is computationally very
expensive. Computational cost may therefore limit the domain size and thus also long-
time simulations.
To remedy this problem, we have therefore incorporated a moving reference frame
into our numerical algorithm. The idea is to move the reference frame together with the
16
front such that the front (e.g. a bubble) remains more or less fixed in the computational
domain. The size of the computational domain may then be chosen independently of the
duration of the simulation, and this will in turn reduce the computational cost
significantly. As a result, we may carry out long-time simulations of moving interface
problems which could not be done in a stationary reference frame.
Figure 6 illustrates a moving reference frame. There, the frame
XY
stands for a
stationary reference frame and the frame
Y
X
for a moving reference frame. The
positions of a monitoring point in the frames
XY
and
Y
X
are represented as and
, respectively, which are correlated with the position of the moving reference frame
( ) according to equation (14). The velocity of the monitoring point
p
x
p
x
m
x
P
is in the
frame
)t,(xu
XY
and in the moving frame
),'( txu
Y
X
, and the velocity of the moving frame is
. The following correlation can be obtained: )(
m
u t
pmp
xxx
+
=
(14)
),()(),(
ttt
m
xuuxu
+
=
. (15)
Allowing translational, but not rotational, movement of the frame in the present
study, the following is the updated governing flow equations in the moving reference
frame:
gxxn
uu
u
uu
u
)1()(
*
1
)]([
*
1)(
T
+
+
+
+
=+
+
Γ
ρδκ
μρρ
ρ
ds
Bo
Re
p
dt
d
t
f
m
. (16)
0
=
u (17)
The moving front will generally be accelerating and so will the moving reference
frame. Thus the frame of reference in which we solve the governing equations is no
longer an inertial frame, and we must therefore modify the momentum equations to take
into account the acceleration of the frame. In addition, according to equation 15, when
the governing equations are solved in a moving reference frame, the velocity condition on
the boundary ( ) should be modified as
'
B
x )(),(),'(' ttt
mBB
uxuxu
=
.
17
Notice that the additional term on the left-hand side
dtd
m
u , which denotes the
acceleration of the moving reference frame, is added in Equation (16). We aim to
choose so that the rising bubble remains as fixed as possible in the moving frame, i.e.
ideally the acceleration of the frame is equal to the acceleration of the bubble. The bubble
acceleration is of course unknown, so we need to approximate this acceleration at each
time step. We shall adopt the prediction as presented by Rusche [35], namely
m
a
m
a
t
n
m
n
m
Δ
Δ
=
+
+
1
1
u
a
, (18)
where
tt
n
d
n
d
n
dd
n
m
Δ
Δ
=Δ
+
1
2
0
1
1
xxxx
u
λλ
. (19)
Here is the position of the centre of mass of the bubble relative to the moving frame
at time step
j
d
x
j
, , and
1
)(
=Δ
nnn
ttt
1
λ
and
2
λ
are appropriate under-relaxation factors.
It was found that
1
λ
= 1.0
2
=
λ
gave good results.
2.7. Solution Procedure
We may now summarize the main steps in advancing the solution from one time step
to the next as follows:
(1)
The velocity of the front marker points,
n
f
u , is calculated through interpolation of
the fluid velocity field
n
u
according to Equation (12).
(2)
The front is advected to its new position
1+n
f
x by using the normal interface
velocity
n
f
u found in step (1) – see Equation (13). The front elements are then
subject to examination for adaptation and topological change. Meanwhile, volume
conservation is enforced.
(3)
The indicator function
)(
11 ++ n
f
n
is computed based on the interface position
1+n
f
x . This is done by solving the Poisson problem in Equation (7) with the
discrete delta distribution from Equation (8). Subsequently, the distribution of the
I x
18
density
1+n
ρ
, the viscosity
1+n
μ
and the surface tension
1+n
σ
F is updated on the
flow solver grid points.
(4)
We find the velocity field
1+n
u and the pressure
1+n
p by solving the mass
continuity and momentum equations using a modified version of the SIMPLE
algorithm. Appropriate boundary conditions are applied.
(5)
Repeat steps (1) to (4) to advance the solution to time
2+n
t .
3. RESULTS AND DISCUSSION
In this Section, we report various numerical results for different purpose, e.g. model
sensitivity analysis, validation case, and model capabilities exploration. A summary of all
the simulations cases can be found in Table I, in which all simulation parameters are
listed.
3.1. Sensitivity Analysis - Size of the Computational Domain
Extensive experiments have been performed in the past to study the rise and
deformation of single bubbles in quiescent liquids. Often these experiments have been
done in large containers with a size of at least 20 bubble diameters in each spatial
direction to avoid wall containment effects [1]. In this paper, we intend to validate
simulation results against such experiments. To achieve this, the computational domain
should also be rather large to avoid any significant effects caused by the wall
confinement. On the other hand, if the domain is chosen too large, excessive computing
time is needed to complete the simulation. To analyze the influence of the domain size, a
number of numerical tests were run with domain sizes ranging from two to twelve bubble
diameters in each of the spatial dimensions. The grid resolution was kept constant and
accommodated approximately twenty cells inside the bubble in each direction. All
computations were carried out in a moving reference frame.
The terminal rise velocity and the terminal bubble shape were used to assess the wall
confinement effects for various domain sizes. The aim of this sensitivity analysis is to
find the smallest possible computational domain in which wall containment effects have
negligible impact on the bubble terminal velocity and shape. Figure 7 presents the
19
simulation results of the terminal bubble shape and the rise velocity using different
domain sizes under the conditions of
0.243Bo*
=
, 24.15*R
=
e , 1000/ =
bl
ρ
ρ
and
100/ =
bl
μ
μ
(Case A3). There is a notable change in the terminal bubble shape and rise
velocity as the domain size is increased from two to six bubble diameters. When the
domain size is increased beyond six bubble diameters, no significant change in the
simulation results is observed. Moreover, it is noted that the wall confinement has a
strong effect on the terminal velocity for small domain sizes from two to six bubble
diameters. However, the change in terminal velocity is only around 1% when increasing
the computational domain size from eight to ten bubble diameters in each spatial
dimension. Based on these observations we conclude that a domain size with side length
of eight bubble diameters should be sufficient in our simulations. Actually, experimental
results by Krishna et al. [23] also indicate that wall effects on a rising bubble is negligible
when the diameter of the liquid container is larger than eight bubble diameters.
3.2. Sensitivity Analysis - Grid Resolution
PARAMESH divides the computational domain into a number of blocks in each
spatial direction. Each block consists of a certain number of grid cells in each direction,
and the governing equations are discretized and solved numerically on these grid cells.
For a computational domain of fixed size, there are therefore two ways to change the grid
resolution using PARAMESH: either by changing the number of blocks, which is
determined by the maximum number of refinement levels, or by changing the number of
cells in each block. In our grid sensitivity analysis we kept the maximum refinement level
fixed and changed the number of cells in each block. All simulations were done in a
moving, cubic computational domain with side length equal to eight bubble diameters
using equal grid spacing in each of the spatial directions.
The results from the grid sensitivity analysis can be found in Figure 8 under the
conditions of ,
0.243Bo* = 24.15*R
=
e , 1000/
=
bl
ρ
ρ
and 100/
=
bl
μ
μ
(Case A3). It is
noted that the terminal bubble velocity is highly sensitive to the mesh resolution up until
16 cells per bubble in each space direction. However, increasing the number of cells from
16 to 20 yields less than 1% change in terminal velocity. The terminal bubble shape also
has a very strong dependence on the grid resolution - especially when the grid resolution
20
is relatively low. On the other hand, we can even see a slight change in shape when we
increase the number of cells per bubble diameter from 16 to 20. However, increasing the
number of cells per bubble diameter further to 32 does not lead to any visually detectable
change in bubble shape. Based on the above we find it sufficient to use a resolution of 20
cells per bubble diameter in our simulations.
3.3. Comparison of Results Obtained in a Stationary and a Moving Reference Frame
Since a moving reference is introduced in this paper to perform long-time simulations
of single bubbles rising as well as two-bubble interactions, it is important to evaluate the
accuracy and impact of using a moving reference as opposed to a stationary frame. For
this purpose, simulation results for Case A4 obtained in a stationary reference frame are
compared with those obtained in a moving reference frame.
The bubble shapes predicted in both stationary and moving frames at different time
steps are shown in Figure 9. There are no visually observable differences between the
results from the two frames. A comparison of the velocity profiles of bubbles rising in a
stationary and a moving reference frame is shown in Figure 10, and they are in
reasonable agreement. Note that the velocity of the moving frame has been added to the
rise velocity obtained in the moving reference frame to enable a direct comparison with
the rise velocity in the stationary frame. In addition to comparing bulk behavior of rising
bubbles, we would also like to compare the detailed flow patterns around the bubbles.
A comparison of the streamlines predicted in a stationary and a moving frame can be
found in Figure 11. Since one frame is at rest and one is moving, comparing streamlines
would only be sensible if we modify the velocity in one of the frames to account for the
different velocities of the frames. Here we have chosen to subtract the velocity of the
moving reference frame from the velocity field computed in the stationary frame before
comparison. It can be seen from Figure 11 that an excellent agreement of the streamline
patterns in the stationary and the moving frame is obtained. Furthermore, pressure
distributions predicted in a stationary and a moving frame are compared in Figure 12, and
again the results agree well.
21
Based on the various tests and comparisons of different flow characteristics carried
out and described above, we can conclude that the use of a moving reference frame yields
numerical results equivalent to those obtained in a stationary reference frame.
3.4. Model Validation with Experiments
In this section we will compare experimental results available in the literature [1]
with predictions obtained by our numerical method. In Figure 13 we compare observed
and predicted terminal bubble shapes for a range of Reynolds and Bond numbers, and the
results agree very well. In Table II we compare the associated terminal rise velocities,
and again there is reasonable agreement between experiments and our numerical
predictions. However, note that the relative deviation in Case A1 is a little bit high. This
may be due to the low rise velocity where the relative error will be high even though
there is no change in the simulation accuracy.
Air bubbles rising in water are common in many industrial processes. Examples in
chemical engineering include bubble columns, loop reactors, agitated stirred reactors,
flotation, or fermentation reactors. For the design of efficient two-phase reactors, detailed
knowledge of bubble sizes and shapes, slip velocities, internal circulations, swarm
behaviors, bubble induced turbulence and mixing, and bubble size distributions
(including coalescence and breakup) is of fundamental importance. In such industrial
applications, bubbles often have non-spherical and even dynamic shapes as well as
asymmetric wake structures. Extensive experimental studies have been performed to
study air bubbles rising in water [7, 44]. Their measurements of the terminal rise velocity
of air bubbles in water are presented in Figure 14 as a function of the bubble size. It is
found that the measurements of the terminal velocity vary significantly (or bifurcation)
when the bubble size is greater than 0.5 mm and smaller than 10 mm. Traditionally this
variation has been explained by the presence of surfactants [7], but more recently both
Wu and Gharib [47] and Tomiyama et al. [44] attributed this variation to the manner in
which the initial bubbles were generated. The issue continues to be a matter of discussion
- refer to Yang and Prosperetti [48].
Due to difficulties in measuring the physical properties on the bubble, a fundamental
understanding of the system of a single bubble rising in high Reynolds number regimes is
22
not well-established. With the recent rapid increase in computing power, numerical
simulations of two-phase flows based on continuum mechanics models with moving free
interfaces have become feasible and proved extremely useful for a better understanding
of fundamental processes and phenomena. However, numerical modeling of the multi-
fluid system of air bubbles rising in water is still quite challenging due to the large
density ratio of water to air, the low liquid viscosity of water, high Reynolds numbers,
and large bubble deformations. Koebe et al. [21] started early trials of 3D direct
numerical simulation of air bubbles rising in water at high Reynolds number using the
volume of fluid (VOF) method. They studied bubbles with diameters from 0.5 mm to 15
mm, and their numerical predictions on the terminal rise velocity of the bubbles agree
reasonably with experimental data. However, they introduced some initial white noise in
the simulations, which may introduce non-physical perturbations to the simulation
system. The recent work by Dijkhuizen et al. [9] reported their trial on simulation of
single air bubbles rising in initially quiescent pure water using both a 3D front tracking
method and a 2D VOF method for bubble diameters ranging from 1 mm to 8 mm. The
calculated terminal rise velocites by the 3D front tracking method are quite close to the
experimental observations by Tomiyama et al. [44], but they over-predicted the velocity
for bubble diameters larger than 3 mm.
In this paper, we use the front tracking method with features of mesh adaptation and
moving reference frame, allowing a finer mesh in the region of the bubble surface.
Consequently, better accuracy is obtained in the current simulations. We simulate a single
air bubble rising in initially quiescent pure water with the bubble diameters ranging from
0.5 mm to 30 mm. The numerically predicted rise velocities of the bubbles agree well
with the upper bound of the experimental measurements by Tomiyama et al. [44] within
the whole range of different bubble sizes. When the bubble diameter is in the range from
2.0 mm to 10 mm, oscillation of the bubble rise velocity and the bubble shape is also
predicted in the simulations. The terminal bubble rise velocity is calculated through
averaging the instantaneous rise velocity over a period of time. Since we assume the
initial bubble shape to be spherical and the surface tension coefficient to be constant, the
bifurcation of the bubble rise velocity is not revealed in the current simulation. However,
this is an interesting topic to be explored in the future.
23
3.5 Numerical Studies on the Interaction between Two Rising Bubbles in Viscous Liquid
The problem of a single bubble rising in a viscous liquid is an ideal case for numerical
model validation. However, the final goal when developing a numerical model for
multiphase flow is not just investigating the flow behavior of single bubbles rising in
viscous liquids, but also investigating multi-fluid systems with multiple bubbles. With
the confidence from validating the current model for a single bubble rising in a viscous
liquid, we would like to extend this model to explore the complex interaction between
two bubbles rising in a liquid. Figures 15 and 17 illustrate the simulation of the
interaction of two initially spherical bubbles rising in a quiescent liquid due to buoyancy.
In the first simulation, one smaller bubble is initially located 2.5D above a bigger
bubble in vertical direction, and 0.5D axis-off from the bigger bubble in the horizontal
direction of Y. Here, D represents the effective diameter of the bigger bubble. The
diameter of the smaller bubble is half that of the bigger bubble. The flow conditions for
the bigger bubble are as follows:
6.134*R
=
e
,
0.115Bo*
=
1181/
=
bl
ρ
ρ
and (Case
B1). Figure 15 shows the temporal bubble shape evolution of two rising bubbles. As the
bigger bubble has a higher rise velocity, it will catch up with the smaller bubble (Figure
15(
5000/ =
bl
μμ
0.4=
τ
)). When they are close enough, the trailing bigger bubble is significantly
affected by the low-pressure zone in the wake of the leading smaller bubble. The trailing
bubble undergoes large deformations and moves towards the bottom wake zone of the
leading bubble (Figure 15(
0.6=
τ
)). Finally, the trailing big bubble merges with the
leading smaller bubble, and a toroidal bubble ring is formed (Figure 15(
0.10=
τ
)).
Similar bubble shape evolution patterns have also been predicted by other numerical
predictions [40]. In addition, Figure 16 shows the temporal variation of the position of the
bubbles in both vertical and horizontal directions. It can be seen from Figure 16(a) that
the trailing bigger bubble has a higher rise speed than the smaller leading bubble. The
interesting finding is that when the two bubbles are close enough, then the rise speed of
both bubbles increases significantly. After the coalescence of the two bubbles, the
resulting merged bubble returns to a normal situation of a single bubble. The lateral
movement of the trailing bubble caused by the leading bubble can be seen in Figure
16(b). Even though initially the leading bubble moves slightly away from the trailing
24
bubble laterally, this distance is quite small. However, the trailing bubble, despite its big
size, is significantly affected by the leading bubble and moves towards it.
In the second simulation, the smaller bubble is initially located 2.5D above the bigger
bubble in vertical direction, and 1.0D center-off from the big bubble in the horizontal
direction of Y. Here, D represents the effective diameter of the bigger bubble. The
diameter of the smaller bubble is half that of the bigger bubble. The flow conditions for
the bigger bubble are as follows:
24.15*R
=
e
,
0.243Bo*
=
, and
1181/ =
bl
ρρ
5000/ =
bl
μ
μ
(Case B2). Figure 17 shows the temporal evolution of the two rising
bubbles. As a result, in this case the bigger bubble finally overtakes the smaller one, since
the big bubble is further horizontally off-set the small one. It can be seen from Figures 17
and 18(b) that the leading bubble first starts moving laterally away from the trailing
bubble, whereas the trailing bubble then starts moving towards the leading bubble before
the overtaking occurs (Figure 17(
0.4
=
τ
, 0.6
=
τ
, 0.8
=
τ
)) . After the bigger bubble has
overtaken the smaller one, the smaller bubble is significantly affected by the wake of the
bigger bubble. In fact, the smaller bubble is attracted to the wake of the bigger bubble,
resulting in a highly deformed and elongated bubble shape as shown in Figure 17
(
0.12=
τ
,
0.14=
τ
,
0.16=
τ
). It is also noticed from Figure 18(a) that the smaller bubble
is accelerated and rises fast in the wake of the bigger bubble, and finally it catches up and
merges with the bigger bubble (Figure 17(
0.18
=
τ
, 0.20
=
τ
, 0.22
=
τ
)). In this case, the
smaller trailing bubble has little effect on the rising speed of the bigger leading bubble.
4. CONCLUSION
The numerical algorithm used in this paper is a further extension of the algorithm
given in [17] for simulating 3D gas bubbles rising in viscous liquids at high Reynolds and
Bond numbers for systems with large density and viscosity ratios such as air/water. To
achieve this, mesh adaptation is implemented for both the front mesh and the background
mesh, and a moving reference frame attached to the rising bubble is used to solve the
governing incompressible Navier-Stokes equations. The solution method is a finite
volume scheme based on the Semi-Implicit Method for Pressure-Linked Equations
25
(SIMPLE), and the solver appears to be robust even in the range of high Reynolds
numbers and high density/viscosity ratios. The bubble surface is tracked explicitly using
an adaptive, unstructured triangular mesh. The model is integrated with the software
package PARAMESH, a block-based adaptive mesh refinement (AMR) tool developed
for parallel computing. It includes features such as background mesh adaptation and
parallel implementation of solvers for the governing equations on supercomputers. The
interpolations between the front mesh and the background mesh are done with Peskin’s
traditional approximation of the delta function. The current model has been applied to
simulate a number of examples of 3D gas bubbles rising in viscous liquids, e.g. air
bubbles rising in water. The simulation results are compared with experimental
observations in aspect of both the terminal bubble shape and the terminal bubble velocity.
In addition, we use this model to simulate the interaction between two bubbles rising in
liquid. The simulation results provide us with some physical insights into the complex
behavior of bubbles rising in viscous liquids.
REFERENCES
1.
D. Bhaga and M.E. Weber, Bubbles in viscous liquids: shapes, wakes and
velocities,
J. Fluid Mech. 105, 61 (1981).
2.
T. Bonometti and J. Magnaudet, Transition from spherical cap to toroidal bubbles,
Phys. Fluids. 18, 052102 (2006).
3.
L.A. Bozzi, J.Q. Feng, T.C. Scott, and A.J. Pearlstein, Steady axisymmetric
motion of deformable drops falling or rising through a homoviscous fluid in a
tube at intermediate Reynolds number,
J. Fluid Mech. 336, 1 (1997).
4.
J.U. Brackbill, D.B. Kothe, and C. Zenmach, A continuum method for modeling
surface tension,
J. Comput. Phys. 100, 335 (1992).
5.
B. Bunner, and G. Tryggvason, Dynamics of homogenous bubbly flows. Part I:
Rise velocity and microstructure of the bubbles,
J. Fluid Mech. 466, 17 (2002).
6.
L. Chen, S.V. Garimella, J.A. Reizes, and E. Leonardi, The development of a
bubble rising in a viscous liquid,
J. Fluid Mech. 387, 61 (1999).
26
7.
R. Clift, J.R. Grace, and M.E. Weber, Bubbles, Drops, and Particles, (Academic
Press, San Diego, 1978).
8.
R.M. Davies, and F.I. Taylor, The mechanism of large bubbles rising through
extended liquids and through liquids in tubes,
Proc. R. Soc. Lond. A 200, 375
(1950).
9.
W. Dijkhuizen, E.I.V. van den Hengel, N.G. Deen, M. van Sint Annaland, and
J.A.M. Kuipers, Numerical investigation of closures for interface forces acting on
single air-bubbles in water using volume of fluid and front tracking method
models,
Chem. Eng. Sci. 60, 6169 (2005).
10.
A.L. Fogelson, and C.S. Peskin, A fast numerical method for solving the three-
dimensional Stokes equations in the presence of suspended particles,
J. Comput.
Phys.
79, 50 (1988).
11.
J. Glimm, O. McBryan, R. Menikoff, and D.H. Sharp, Front tracking applied to
Rayleigh–Taylor instability,
SIAM J. Sci. Statist. Comput. 7, 230 (1986).
12.
J. Glimm, J.W. Grove, B. Lindquist, O. McBryan, and G. Tryggvason, The
bifurcation of tracked scalar waves,
SIAM J. Sci. Stat. Comput. 9, 61 (1988).
13.
J. Glimm, J.W. Grove, X.L. Li, and D.C. Tan, Robust computational algorithms
for dynamic interface tracking in three dimensions,
SIAM J. Sci. Comput. 21,
2240 (2000).
14.
C.W. Hirt and B.D. Nichols, The Volume of Fluid (VOF) method for the
Dynamics of Free Boundaries,
J. Comput. Phys. 39, 201 (1981).
15.
T.Y. Hou, J.S. Lowengrub, and M.J. Shelley, Boundary integral methods for
multicomponent fluids and multiphase materials,
J. Comput. Phys. 169, 302
(2001).
16.
H.H. Hu, N.A. Patankar, and M.Y. Zhu, Direct numerical simulation of fluid-solid
system using the arbitrary Lagrangian-Eularian technique,
J. Comput. Phys. 169,
427 (2001).
17.
J. Hua and J. Lou, Numerical simulation of bubble rising in viscous liquid. J.
Comput. Phys.
222, 759 (2007).
27
18.
M. Ida, An improved unified solver for compressible and incompressible fluids
involving free surfaces. Part I. Convection.
Computer Physics Communications
132, 44 (2000).
19.
D. Jacqmin, Calculation of two-phase Navier-Stokes flows using phase-field
modeling,
J. Comput. Phys. 155, 96 (1999).
20.
S.I. Kang, and G.L. Leal, Numerical solution of axisymmetric, unsteady free-
boundary problem at finite Reynolds number. I. Finite difference scheme and its
application to the deformation of a bubble in a uniaxial straining flow,
Phys.
Fluids
30, 1929 (1987).
21.
M. Koebe, D. Bothe, J. Pruess, and H.-J. Warnecke, 3D direct numerical
simulation of air bubbles in water at high Reynolds number, In Proceedings of the
2002 ASME Fluids Engineering Division Summer Meeting (FEDSM2002-
31143), Montreal, Quebec, Canada, July 14-18, 2002.
22.
A. Koynov, J.G. Khinast, and G. Tryggvason, Mass transfer and chemical
reactions in bubble swarms with dynamics interface,
AIChE J. 51, 2786 (2005).
23.
R. Krishna, M.I. Urseanu, J.M. van Baten, and J. Ellenberger, Wall effects on the
rise of single gas bubbles in liquids,
Int. Comm. Heat Mass Transfer 26, 781
(1999).
24.
P. MacNeice, K.M. Olson, C. Mobarry, R. deFainchtein, and C. Packer,
PARAMESH: A parallel adaptive mesh refinement community toolkit.
Comput.
Phys. Commun.
126, 330 (2000).
25.
D.W. Moore, The rise of a gas bubble in viscous liquid, J. Fluid Mech. 6, 113
(1959).
26.
G. Mougin and J. Magnaudet, Wake-induced forcesand torqueson a zigzagging /
spiraling bubble,
J. Fluid Mech. 567,185 (2006).
27.
M. Ohta, T. Imura, Y. Yoshida, and M. Sussman, A computational study of the
effect of initial bubble conditions on the motion of a gas bubble rising in viscous
liquids,
Int. J. Multiphase Flow 31, 223 (2005).
28.
S.J. Osher and J.A. Sethian, Fronts propagating with curvature dependent speed.
Algorithms based on Hamilton-Jacobi formulations,
J. Comput. Phys. 79, 12
(1988).
28
29.
S. Osher and R.P. Fedkiw, Level set methods: An overview and some recent
results,
J. Comput. Phys. 169, 463 (2001).
30.
S.V. Patankar, Numerical Heat Transfer and Fluid Flow, (Hemisphere, New
York, 1980).
31.
C.S. Peskin, Numerical analysis of blood flow in the heart, J. Comput. Phys. 25,
220 (1977).
32.
C.S. Peskin and B.F. Printz, Improved volume conservation in the computation of
flows with immersed boundaries,
J. Comput. Phys. 105, 33 (1993).
33.
C. Pozrikidis, Interfacial dynamics for stokes flow, J. Comput. Phys. 169, 250
(2001).
34.
A. Prosperetti, Bubbles, Phys. Fluids 16, 1852 (2004).
35.
H. Rusche, Computational Fluid Dynamics of Dispersed Two-Phase Flows at
High Phase Fractions, Ph.D. Thesis, Imperial College of Science, Technology &
Medicine, 2002.
36.
G. Ryskin and L.G. Leal, Numerical simulation 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 (1984).
37.
R. Scardovelli and S. Zaleski, Direct numerical simulation of free-surface and
interfacial flow.
Annu.Rev. Fluid. Mech. 31, 567 (1999).
38.
W.L. Shew and J.-F. Pinton, Viscoelastic effects on the dynamics of a rising
bubble,
J. Stat. Mech. P01009, (2006).
39.
S. Shin, and D. Juric, Modelling three-dimensional multiphase flow using a level
contour reconstruction method for front track with connectivity.
J. Comput. Phys.
180, 427 (2002).
40.
M. van Sint Annaland, N.G. Deen, and J.A.M. Kuipers, Numerical simulation of
gas bubbles behaviour using a three-dimensional volume of fluid method.
Chem.
Eng. Sci.
60, 2999 (2005).
41.
M. van Sint Annaland, W. Dijkhuizen, N.G. Deen, and J.A.M. Kuipers,
Numerical simulation of behaviour of gas bubbles using a 3-D front-tracking
method,
AIChE J. 52, 99 (2006).
29
30
42.
M. Sussman, P. Smereka, and S. Osher, A level set approach for computing
solutions to incompressible two-phase flows,
J. Comput. Phys. 114, 146 (1994).
43.
T.D. Taylor, and A. Acrivos, On the deformation and drag of a falling viscous
drop at low Reynolds number,
J. Fluid Mech. 18, 466(1964).
44.
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).
45.
G. Tryggvason, B.B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J.
Han, S. Nas, and Y.J. Jan, A front-tracking method for the computations of
multiphase flow.
J. Comput. Phys. 169, 708 (2001).
46.
S.O. Unverdi and G. Tryggvason, A front-tracking method for viscous,
incompressible, multi-fluid flows.
J. Comput. Phys. 100, 25 (1992).
47.
M.M. Wu and M. Gharib, Experimental Studies on the Shape and Path of Small
Air Bubbles Rising in Clean Water,
Phys. Fluids 14, 49 (2002).
48.
B. Yang, A. Prosperetti, and S. Takagi, The Transient Rise of a Bubble Subject to
Shape or Volume Changes,
Phys. Fluids 15, 2640 (2003).