Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN FLUIDS

Int. J. Numer. Meth. Fluids 2014; 75:184–204


Published online 17 February 2014 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/fld.3890

Efficient computation of the flow around single fluid particles


using an artificial boundary condition

D. Weirich, M. Köhne and D. Bothe*,†


Center of Smart Interfaces, Mathematical Modeling and Analysis, Technische Universität Darmstadt, Petersenstr. 17,
64287 Darmstadt, Germany

SUMMARY
Two-phase flows around fluid particles are often considered to be in infinite domains, to avoid influence
of the domain walls. Numerical simulations, however, must be modeled with a bounded domain, thus
introducing artificial boundaries. Modeling of fluid flow in a domain with such artificial boundaries requires
a careful choice of suitable boundary conditions. Slip boundary conditions for example can have a large
impact on the computational results if the domain is chosen to be too small, because they model imperme-
able walls. This paper introduces an artificial boundary condition for simulations of the flow around single
rising or settling fluid particles based on the approximated decay behavior of the velocity and the pressure
field in the surrounding liquid. This is applied to the simulation of rising gas bubbles in systems with a
Reynolds number of up to 50, and the outcome is compared with experimental results and simulations with
slip boundary condition. It is found that domain size can be reduced by a factor of about two compared with
slip boundary conditions without loss of accuracy. Copyright © 2014 John Wiley & Sons, Ltd.

Received 19 December 2012; Revised 15 October 2013; Accepted 25 January 2014

KEY WORDS: two-phase flows; bubbles; Navier–Stokes equations; incompressible flow; finite volume;
hydrodynamics

1. INTRODUCTION

Processes in multiphase flows often span a wide range of spatial and temporal scales. Numerical
computations, however, cannot cover all scales in a single simulation. In particular, DNS cannot
usually be used for the simulation of an entire apparatus. In such cases, scale-reduced numerical
descriptions employing, for example, the Euler–Lagrange or the Euler–Euler method are necessary.
In these approaches, the finest scales are not resolved, but additional models are needed to describe
especially the interphase momentum exchange on these scales. With DNS, the elementary processes
on the finest scales can be investigated if highly resolved computations are feasible.
A prototype example for such a process is the rising or settling of single fluid particles due
to buoyancy. To avoid influences of the surrounding region to the observed process, the flow
around the fluid particle is assumed to be in an unbounded domain. Numerical calculations, on the
other side, need to truncate the problem to a bounded computational domain. Hence, the domain
boundary is placed inside the surrounding liquid instead at a physical boundary. Consequently, the
boundary conditions at infinity are to be replaced with conditions on the arbitrarily chosen domain
boundary . In most cases, the domain boundary is not chosen in accordance with the physical fluid
flow, that is, the streamlines are in most cases not tangential to the bounding surface of the computa-
tional domain. Nevertheless, no-slip or perfect slip boundary conditions are often chosen, which rule
out flows that cross the domain boundary. This is a distortion of the real fluid flow at the artificial

*Correspondence to: D. Bothe, Center of Smart Interfaces, Mathematical Modeling and Analysis, Technische Universität
Darmstadt, Petersenstr. 17, 64287 Darmstadt, Germany
† E-mail: bothe@csi.tu-darmstadt.de

Copyright © 2014 John Wiley & Sons, Ltd.


ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 185

boundary and can impact the flow field not only in the region near the boundary but also near the
fluid particle and thus lead to corrupted results. The influence is strong for small domain sizes and
decreases with increasing domain size. Increasing the domain to reduce effects from the boundary
conditions means increasing also the computational effort and simulations with a fine resolution can
become difficult and need a huge amount of computational time. With adaptive mesh refinement on
unstructured meshes, the problem of increased computational effort can be reduced, but adaptive
mesh refinement in a parallel code requires redistributing of the work load of the processes, which
introduces new problems and requires large implementation effort.
Another possibility is to use an artificial boundary condition (ABC), taking the fluid flow over
the boundary into consideration. The disturbance of the flow at the domain boundary is reduced,
and better results are achieved also with a small domain. Ideally, we would prescribe the fluid flow
on the artificial boundary in a way that exactly reproduces the exterior solution. Then, the artificial
boundary does not produce any disturbance to the flow field at all. As such, boundary conditions
lead to the same results as the unbounded problem independent of the size of the truncated domain
are called exact ABCs. For some flow problems, exact ABCs can be constructed by considering
fluid flow as a propagation of waves. On physical boundaries, these waves are reflected; however,
on artificial boundaries, no reflection should occur. So, exact ABCs must allow outgoing waves to
leave the domain but not reenter it. Because of their property of absorbing waves, such ABCs are
often called absorbing or nonreflecting boundary conditions. Engquist and Majda [1] present a way
to derive such ABCs for general classes of wave equations: They construct the solution of fluid flow
in the Fourier space as a superposition of waves and select on the boundary only the waves that leave
the domain and drop those waves that would enter the domain. Application of the inverse Fourier
transformation leads to a nonlocal operator, and thus, the boundary conditions constructed in such a
way are nonlocal in space and, for instationary problems, in time.
Another method for deriving exact ABCs is presented in [2] by Keller and Givoli. They use
the so-called Dirichlet-to-Neumann map, a nonlocal operator that expresses the normal derivative
of the solution of the exterior problem on the artificial boundary by the boundary value of the
solution itself. They use this approach to derive ABCs for the reduced wave equation in two and
three dimensions for circular and spherical boundaries.
The principal advantage of exact ABCs is their accuracy. But they are difficult to derive for most
problems, and in many cases, only certain geometries are feasible. The main problem however is
the nonlocality. This makes numerical implementations difficult and can cause high computational
costs. Even more severe is the fact that almost nothing is known about the well-posedness of the
resulting initial boundary value problems. For these reasons, exact ABCs are in most cases not used.
Instead, local ABCs are employed because these are in general easier to implement and produce
lower computational costs.
One way to obtain appropriate local ABCs is to approximate nonlocal conditions. The equations
in the transformed space must be modified in such a way that the boundary condition in the physical
space can be written as a local expression. However, these local conditions share the geometric
restrictions with their corresponding exact ABCs. Engquist and Majda [1] use a Taylor expansion in
the transformed space to approximate their exact boundary condition under the assumption of planar
waves hitting the domain boundary nearly from normal direction. In first order approximation, no
transversal wave modes are considered, but with higher orders, more transversal terms occur in the
boundary condition.
The approach of Engquist and Majda [1] is used by many different authors and adopted to a broad
range of problems. For instance, Halpern [3] starts with nonlocal approximate ABCs under the
assumption of small viscosity and derives local ABCs in a similar way to Engquist and Majda [1].
Tourrette [4] applies this method to the linearized two-dimensional compressible Navier–Stokes
system in a half-space. The full incompressible Navier–Stokes equations are treated by Jin and
Braza [5], and the resulting ABCs are applied to an unsteady free shear flow in the transition towards
turbulence. The velocity vector is considered as a transported wave quantity on the boundary, and
incoming waves are dropped. A Padé approximation by the characteristic velocity of the waves is
used to derive a local operator. The boundary condition derived in this way maintains the nonlin-
ear character of the Navier–Stokes equations. Dorodnicyn [6] uses the selection of wave modes

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
186 D. WEIRICH, M. KÖHNE AND D. BOTHE

to construct exact ABCs for the one-dimensional linearized Euler equation and extends it to an
approximate boundary condition for the two-dimensional linearized Euler equation. This nonreflect-
ing boundary condition is extended to nonlinear systems and adjusted to the nonlinear problem of
viscous gas flow over solid bodies.
Local ABCs can also be derived from the asymptotic behavior of the far-field solution in the phys-
ical space. Hagstrom [7] presents such an asymptotic boundary condition for wave propagation in
dissipative systems. A small set of dominant wave groups with minimum decay rates are identified,
and their asymptotic behavior is used to derive a local ABC. Sa and Chang [8] derived asymptotic
boundary conditions for two-dimensional incompressible flows around solid bodies. The stream
function at the artificial boundary is calculated as an integral series expansion around the free stream
velocity. Note that in higher dimensions a stream function formulation is no longer available.
Another approach is taken by Toulopoulos and Ekaterinaris [9] for simulations of the two-
dimensional, compressible Euler equation. With upwind discretization, an approximation of the time
variation of the outgoing waves at the artificial boundary is estimated. Then, the time variations of
the state variables are related to the time variation of the waves at the boundary.
Papanastasiou, Malamataris, and Ellwood [10] describe an ABC-type approach for the FEM. In
FEM, surface integrals occur on the boundaries that result from the reformulation as a variational
problem. In general, these boundary integrals suggest natural boundary conditions. Papanastasiou
et al. [10] do not explicitly impose a boundary condition at all and avoid the integration by parts
for equations associated with the artificial boundary by extending the validity of the weak formu-
lation of the momentum and energy equations to the boundary. Although, at the level of the partial
differential equations, the omission of boundary conditions is not possible and leads to an underde-
termined problem, Renardy [11] has shown that this ‘free’ boundary condition on the discrete level
is equivalent to a boundary condition for the PDE, which extrapolates the solution from the bulk to
the artificial boundary. Even for high Reynolds numbers, it can be shown that boundary layers are
avoided with this approach.
A similar free boundary condition, called the ‘do-nothing condition’, is proposed by Heywood,
Rannacher, and Turek [12]. They use a variational formulation in which a particular choice of a
function space corresponds to a choice of a specific boundary condition. The boundary condition at
an outflow boundary is determined by a space of solenoidal functions that are chosen to carry the flux
from the inflow boundary through the domain to the outflow boundary. An alternative formulation
of this boundary condition allows to prescribe a pressure drop on the domain instead of the flux at
the inflow boundary.
Bao [13] approximates the two-dimensional incompressible flow around an obstacle by the flow in
an infinite channel. On the lateral boundaries, a slip condition is used. Inflow and outflow conditions
are designed using the continuity equation and the normal stress.
Besides the use of ABCs, in some numerical simulations, an absorbing boundary region is used
within which waves and turbulent disturbances are gradually reduced. Hu, Li, and Lin [14] use the
perfectly matched layer technique for nonlinear Euler and Navier–Stokes systems. In the bound-
ary region, the difference between the nonlinear fluctuation and a prescribed pseudo mean flow is
damped exponentially.
With such absorbing boundary regions, the solution can be modified in a way that allows the use of
periodic boundary conditions even if the underlying problem is nonperiodic. In the absorbing region,
the solution is modified until it has a periodic course with minimal back effects to the meaningful
solution in the inner domain. Schlatter, Adams, and Kleiser [15] use a so-called windowing method
for this purpose to simulate nonperiodic problems for the two-dimensional incompressible Navier–
Stokes equations via periodic boundary conditions. A windowing function that has a value of one
in the inner domain and that exponentially decays to zero in the boundary region is multiplied with
the solution, damping it to zero near the boundary. Then, periodic boundary conditions can be used
without impacting the nonperiodicity of the flow in the computational domain itself.
Coclici, Sofronov, and Wendland [16] use an absorbing boundary region in simulations of two-
dimensional and three-dimensional exterior compressible flows. At large distances to the surrounded
body, convective forces dominate over viscous effects. Therefore, in the boundary region, inviscid

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 187

flow is modeled with the linearized Euler equations while in the near-field region the compressible
Navier–Stokes equations are considered.
Concerning adaptive boundary conditions for stationary exterior flow problems, we refer to [17],
which also contains a brief overview of the literature on ABCs for flow problems.
Exact ABCs cannot be found for every problem, and in most cases, the nonlocality prevents the
practical use of this type of boundary condition. For the problem of rising or settling fluid particles,
exact ABCs are not available. The local boundary conditions based on the asymptotic behavior
presented previously can also not be used for the case presented in this paper. To use the local ABC
from Hagstrom [7], a dispersion relation would be needed for the outer problem and must then be
approximated by dominant wave groups. In the boundary condition by Sa and Chang [8], the free-
stream velocity must be given. The boundary conditions by Toulopoulos and Ekaterinaris [9] cannot
be used here as they are adapted to a finite difference discretization. The approach by Papanastasiou
et al. [10] is based on the FEM. The works of Heywood et al [12]. and Bao [13] present only inflow
and outflow conditions; ABCs suited for lateral walls are not presented in these works.
The present paper introduces a novel ABC for the simulation of exterior flow around fluid parti-
cles. This ABC is based on the far-field decay behavior of the fluid flow around single fluid particles
rising or settling under gravity conditions. The decay behavior can be approximated by a Robin-type
boundary condition, where the geometry of the domain is encoded into the coefficients appearing in
the ABC. This makes the boundary condition useable for rather general convex domains, allowing
a reduction of the domain size by a factor of about two (in the directions normal to the walls with
ABCs) without loss of accuracy. In combination with the ABC, a refinement of the resolution near
the fluid particle can save additional computational resources.

2. MATHEMATICAL AND NUMERICAL MODEL OF TWO-PHASE FLOWS

2.1. Continuum mechanical model


The motion of fluid particles is modeled as a two-phase flow with sharp, deformable interface.
We assume the fluids to be Newtonian and incompressible with constant density at isothermal
conditions. For small Mach numbers, the incompressibility assumption is also reasonable inside a
gas phase.
We denote by † D †.t / the sharp interface, which is assumed to be bounded. The continuous
phase is located in the unbounded connected component of R3 n †.t /, which is denoted by C .t /,
while the dispersed phase occupies the not necessarily connected region  .t / WD R3 n N C .t /.
Balance of mass and momentum inside the phases and away from the interface † lead to the
Navier–Stokes equations in  .t / [ C .t /, that is,

@t .v/ C r  .v ˝ v/ D rp C r  S C  g;


(1a)
r  v D 0:

Here,  denotes the densities of the two fluids, which are assumed to be constant in the respective
phase. Moreover, v and p denote the velocity field and the pressure of the flow, respectively, and g
represents the mass specific density of exterior forces, for example, due to gravity. The stress tensor
is given for Newtonian fluids as

S D  .rv C .rv/T /;

where the viscosity  is assumed to be constant for each of the phases. At the interface, balance of
mass and momentum are modeled via the jump conditions

.v  v† /  n† D 0;
(1b)
 v ˝ .v  v† / C p I  S  n† D   n† C r† ;

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
188 D. WEIRICH, M. KÖHNE AND D. BOTHE

with the interface velocity v† and the unit normal n† at the interface. Whenever the orientation of
the interface is relevant, we assume n† to point from  into C . Furthermore,  D r  n† is
the sum of the local principal curvatures, and r†  is the surface gradient of the surface tension.
The notation

 .x/ D lim . .x C h n† /  .x  h n† //
h!0C

represents for the jump of a physical quantity across the interface †.


In all cases considered in this paper, no phase change occurs, and the surface tension is assumed
to be constant. As both phases have nonzero viscosity, we assume that there is no slip between the
phases at the interface. In this case, the interfacial jump conditions can be simplified to read

v D 0;
(1c)
p I  S  n† D   n† :

Finally, an initial velocity v0 is prescribed, and the fluid is assumed to be at rest far away from the
interface, that is,

v.0/ D v0 ; lim v.x/ D 0: (1d)


jxj!1

2.2. Numerical method


The numerical method employed here is based on the volume-of-fluid (VOF) method by Hirt and
Nichols [18]. The VOF method approximates the solutions to the Navier–Stokes equations for an
incompressible transient two-phase flow. The fundamental idea of this method is to capture the
interface position implicitly by means of a phase indicator function, that is, a scalar function f D
f .t; x/ with, say, f D 1 in the dispersed phase and f D 0 in the continuous phase. Because of the
absence of phase change, the transport of f is governed by the advection equation

@t f C v  rf D 0: (2a)

Therefore, the VOF method inherently conserves phase volume. This is an important issue espe-
cially if the long-term behavior of solutions is to be studied, where other methods like level set or
front tracking could lose too much of the droplet or bubble volume. In the finite volume (FV) dis-
cretization scheme employed here, the cell centered value of f corresponds to the phase fraction
inside a computational cell. On the basis of these values, an approximation of the interface normal
can be computed as

rf
n† D :
krf k

A combination of the fractional volume of the dispersed phase with the interface normal then allows
a phase-volume consistent interface reconstruction with the so-called piecewise-linear interface cal-
culation (PLIC) method [19]. This reconstruction of the phase geometry by local planes inside the
computational cells generates important subgrid-scale information which can be employed for an
accurate geometrical advection of the phase indicator without smearing. Once the phase distribution
is known, a one-field formulation of the Navier–Stokes system is possible in which the interfacial
momentum jump condition acts as a source term in the momentum equations. It reads as

 @t v C .v  r/v D rp C r  S C  g C   n† ı† ;


(2b)
r  v D 0;

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 189

where ı† denotes the interfacial Dirac distribution, discretized as jjrf jj. The material properties
 and  we refer to the phase dependent values, and local values of the density and viscosity are
calculated as

 D d f C c .1  f /;  D d f C c .1  f /

(cf. [20, 21] for the more sophisticated choice of the viscosity in interfacial cells).
In contrast to the model problem (1), which is posed in R3 , the one-field formulation (2) is to
be solved in a bounded domain   R3 , because the FV method is not suitable to be applied to
unbounded domains. Because we may assume g D ge3 for some scalar constant g and because
we are interested in a numerical simulation of single fluid particles that rise or fall due to gravity,
we choose  D Œd; d 2  Œh; h with constants d; h > 0. Thus, we choose the domain of the
flow to be a rectangular tube, which introduces the six artificial boundaries

k˙ D ¹ x 2 @ W xk D ˙d º ; k D 1; 2; 3˙ D ¹ x 2 @ W x3 D ˙h º :

We now need to complement the numerical model (2) by appropriate boundary conditions at the
aforementioned artificial boundaries, which substitute the boundary condition at infinity in (1d).
The two artificial boundaries 3˙ , which are approximately perpendicular to the streamlines of the
expected flow, are treated as inlet respectively outlet, that is, we impose the boundary conditions

@3 vk D 0 on 3C ; k D 1; 2;
p D0 on 3C ;
(2c)
@3 vk D 0 on 3 ; k D 1; 2;
v3 D 0 on 3 ;

if we expect the fluid particle to rise in the surrounding liquid. The remaining lateral artificial bound-
aries k˙ , k D 1; 2 demand a more sophisticated modeling, because there the flow is expected to
be neither perpendicular nor parallel to the boundary, and streamlines are expected to leave and
reenter the domain through these boundaries. A derivation of a suitable set of ABCs is presented in
Section 3.
In this work, the inhouse VOF-code free surface 3D (FS3D), originally developed by Rieber
and Frohn, is used (we refer to [21] for details). The spatial discretization is based on a Cartesian
staggered grid according to the well-known Marcer-and-Cell (MAC) scheme. To keep the interface
sharp, the PLIC reconstruction is used for the advection of the interface. Inside the bulk phases, a
simple first-order upwind scheme can be used for the convective transport of the volume fraction f .
In FS3D, a Godunov splitting is employed for the calculation of the velocity field. The calculation
is performed in three steps. First, all convective acceleration terms are considered by computing the
intermediate velocity field

1  nC1 n 
vK D  v  ıt r  .vn ˝ vn / ; (3a)
nC1

with ıt denoting the size of the timestep. Then, all nonconvective terms are considered, except the
pressure acceleration. This amounts to compute the field

ıt
vO D vK C Œr  SK C   n† ı† : (3b)
nC1

Finally, the pressure acceleration is calculated according to

ıt
vnC1 D vO  rp nC1 ; (3c)
nC1

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
190 D. WEIRICH, M. KÖHNE AND D. BOTHE

where the pressure is calculated from the pressure Poisson equation. For this purpose, the divergence
condition is combined with (3c) to obtain
 
1 r  vO
r  nC1 rp nC1 D : (3d)
 ıt

This pressure Poisson equation has to be complemented by a suitable boundary condition. How-
ever, because of the relation (3c), a boundary condition for the pressure implicitly imposes a
boundary condition for the velocity vnC1 and vice versa. For the two boundaries 3˙ , this has the
following consequences. We complement the set of boundary conditions (2c) by

@3 v3 D 0 on 3C ; @3 p D 0 on 3 (4)

and employ the conditions (2c, 4) for the approximate velocity fields vK and vO as well as for vnC1 .
Indeed, the boundary conditions (2c), together with (3c), imply

nC1  nC1 
@3 p D v3  vO 3 D 0 on 3 :
ıt
The boundary condition for v3 in (4) follows analogously.
Boundary conditions are realized in FS3D with dummy cells, that is, two additional layers of
cells outside the physical domain. Adding these cells, the same spatial discretization scheme can be
employed at the boundary as inside the domain. The boundary conditions are instantiated by setting
appropriate dummy values in those cells. Setting these dummy values from derivatives in normal or
tangential direction to the domain boundary can be performed without interpolation. However, for
the calculation of dummy values from derivatives in oblique directions, values have to be mapped
via interpolation to the dummy cells. Interpolation errors in the pressure boundary condition lead to
a pressure acceleration in (3c) that does not correctly reproduce the velocity boundary conditions.
The velocity boundary condition at the new timestep can then only be fulfilled by a discrete velocity
field that is not divergence free in the cells near the domain boundary. As this divergence is fed back
as a source term into the pressure Poisson equation in the next timestep, small errors result, which
will accumulate over time and finally lead to large errors in the continuity equation. Hence, the
numerical stability is improved, if the pressure and velocity boundary conditions are implemented
using only derivatives in normal direction to the domain boundary.
In FS3D, the computation of the surface tension forces can be performed with the continuum
surface force model of Brackbill et al. [22] or with the continuous surface stress method of Lafaurie
et al. [23]. In both methods, the interface unit normal n† D rf =krf k and the curvature  D
r  n† of the interface are calculated from the derivatives of a smoothed (mollified) approximation
of the volume fraction field f . These methods are prone to inaccuracies and yield results that do
not converge with grid refinement. This causes spurious currents locally at the interface and distorts
the flow field there. To avoid this problem, the surface tension is discretized with a balanced-force
implementation of the continuum surface force method by Francois et al. [24]. In this method,
normals and curvature are calculated from height functions. A fixed 7 x 3 x 3 stencil is used in
[24]. This stencil can be problematic in regions with high curvature in the order of the grid size.
Popinet [25] addresses this issue by calculating local asymmetric stencils for the height functions,
an approach that is also employed in FS3D.

3. ASYMPTOTIC BOUNDARY CONDITIONS

The boundary condition presented in this work is modeled on the basis of the far-field decay behavior
of the velocity field and the pressure for the model problem in the whole space R3 . It is convenient
to consider the governing equations in a frame attached to the (moving) center of mass xp .t / of the
dispersed phase. Hence, we set

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 191

u.t; y/ WD v.t; xp .t / C y/  vp .t /;
q.t; y/ WD p.t; xp .t / C y/   g  .xp .t / C y/ C  ap .t /  .xp .t / C y/;

where vp D xP p denotes the velocity of the center of mass of the particle and ap D vP p denotes its
acceleration. The equations (1) for the transformed quantities u and q then read

 @t u C  u  r u D  u  rq;
r  u D 0;
u D 0;
(5)
q I   .ru C .ru/T /  n D   n C .n ˝ y/.g  ap /;
lim u D vp ;
jyj!1

u.0/ D u0 ;

where the transformed interface  is given as


® ¯
.t / D y 2 R3 W xp .t / C y 2 †.t /

and the jump bracket ŒŒ   now denotes the jump of a quantity across .
To motivate the ABC that we are going to use, and for a simple and instructive derivation, we first
assume that the flow in and around the rising fluid particle has reached a steady state with respect
to the attached frame and that the flow is laminar, that is, the Reynolds number is sufficiently small.
Moreover, we assume the fluid particle to have a spherical shape that implies  D .t / D @BR .0/
for some given radius R > 0. Finally, we may without loss of generality assume g D ge3 . Then,
we have vp D ve3 and ap D 0, and system (5) simplifies to

 u  rq D 0;
r  u D 0;
u D 0; (6)
q I   .ru C .ru/ /  y D 2  y=R  g .y ˝ y/e3 ;
T

lim u D ve3 :
jyj!1

Note that n D y=R and  D 2=R. Now, employing the spherical coordinates
0 1
r sin.
/ cos. /
B C
y D @ r sin.
/ sin. / A ; u D ur er C u e C u e
r cos.
/

with the standard choice for er , e , and e , the solution to (6) is given as

u0r D .a1 r 2 C a2 / cos


; ur D .b1 =r 3 C b2 =r C b3 / cos
;
u0 D .2a1 r 2 C a2 / sin
; u D .b1 =2r 3  b2 =2r  b3 / sin
; (7a)
0 2
q D 10  a1 r cos
C a0 ; q D  b1 =r cos
;

where .u0 ; q 0 /, respectively .u; q/ describe the flow inside, respectively outside the fluid particle.
Note that u0 D u D 0 due to the assumed symmetries.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
192 D. WEIRICH, M. KÖHNE AND D. BOTHE

The constants can be determined from the boundary conditions and the jump conditions at the
interface, resulting in the Hadamard-Rybczynski solution. The constants are

2 vp 1
a0 D ; a1 D ;
R 2 1 C 0 =
vp 1 1 vp 3 0 =
a2 D  ; b1 D  R ; (7b)
2 R 1 C 0 =
2 2 1 C 0 =
vp 2 C 30 =
b2 D ; b3 D vp :
2 1 C 0 =

Note that the rise velocity vp is uniquely determined in terms of the densities  of the fluids, the
radius R of the fluid particle, and the magnitude g of the gravitational forces.
Thus, in this case, we obtain

C C0
u.y/ C vp D C o.jyj1 /; q.y/ D C o.jyj2 / (8)
jyj jyj2

and this asymptotic structure will be used below to determine suitable ABCs. In particular, our
ABC based on the spatial asymptotics of u and q does not depend on the constants (7b). At this
point, in order to judge the range of applicability of the resulting ABC, it is crucial to note that
the asymptotic structure (8) is valid for a large class of two-phase flows and not restricted to the
case where the flow has reached a quasistationary state, is laminar or exhibits a small Reynolds
number. Indeed, the asymptotic structure (8) can be obtained for all quasistationary flows around a
possibly translating and rotating rigid body [26]. Moreover, in this case, the asymptotics does not
depend on a possibly inhomogeneous boundary condition on the fluid–rigid body interface. This
leads to the observation that the aforementioned asymptotics remains valid under the following
relaxed conditions: Whenever there exists a constant velocity vector vp and a radius l > 0, such that

N  .t /  Bl .xp .t //;
 xp .t / D xp .0/ C t vp ; t > 0;

that is, the disperse phase moves within a sufficiently large ball that translates at a constant velocity,
then the aforementioned transformations lead to the equations

 @t u C  u  r u D  u  rq;
r  u D 0;
uj@B .xp .0// D w;
(9)
lim u D vp ;
jyj!1

u.0/ D u0 ;

in R3 n Bl .xp .0// with some possibly inhomogeneous (and time dependent) boundary velocity w.
Indeed, as has been shown in the recent work [27], the solutions to (9) always exhibit the asymptotic
structure (8), despite the fact that we have to deal with a dynamic (nonstationary) flow at an arbitrary
large Reynolds number. Consequently, our ABC is applicable to a large class of two-phase flows.
For example, a rise on a helical path is admissible, as well as a fully instationary rise with bounded
variation against a mean rise velocity. Even a group of bubbles would be admissible if the bubbles
stay in a common, translating ball.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 193

To derive the ABC, we transform back to the reference frame. Then, the asymptotics (8) imply
1 2
@r v C v D o.r1 /; @r .p C gx3 / C .p C gx3 / D o.r1 / as r ! 1; (10)
r r
because

v.t; x/ D u.t; x  xp .t // C vp .t /; p.t; x/ C gx3 D q.t; x  xp .t //:

Here, r D r.t; x/ D jx  xp .t /j denotes the distance to the center of mass of the fluid particle or to
the center of a ball Bl .xp .t // that contains the fluid particle as described previously. Note that this
decay behavior is independent of the choice of the computational domain. However, a formulation
of the boundary condition in normal direction to the domain boundary @ is preferable for the
numerical implementation for reasons explained previously. This will be given next, where we are
interested in the special case in which the center of mass of the fluid particle is located at the origin of
the reference frame; for a rising or falling particle, this can be achieved by a continuous translation
of the reference frame, as it is employed in the numerical simulations. In this case, xp .t /  0
and r D r D jxj, and the aforementioned boundary conditions may be rewritten on the artificial
boundaries k˙ (k D 1; 2) as

˙ @k vl C ˛kl .x/vl D 0 on k˙ ; k D 1; 2; l D 1; 2; 3; (11a)

while the pressure boundary condition becomes

˙ @k p C ˇk .x/p D h.x/ on k˙ ; k D 1; 2: (11b)

Here, the coefficients, which encode the geometry of the domain boundary, are given as
3x1 1 3x2
˛11 D  ; ˛21 D ;
jxj2 x1 jxj2
3x1 3x2 1
˛12 D ; ˛22 D  ;
jxj2 jxj2 x2
   2  (11c)
3 x1 jxj2 C 3x23 3 x2 jxj C 3x23
˛1 D 2  2 ; ˛2 D 2  2 ;
jxj jxj C x23 jxj jxj C x23
x1 x2
ˇ1 D 2 ; ˇ2 D 2 ;
r r
and the inhomogeneous right-hand side h of the pressure boundary condition stems from the term
gx3 . The aforementioned set of conditions is compatible with the splitting scheme discussed in
Subsection 2.2.
The ABC presented here can be combined with other techniques that help enhancing the resolu-
tion near the bubble. One possibility is to reduce the resolution in the outer liquid at some distance
to the rising bubble. In FS3D, the cells near the domain boundary are stretched in normal direc-
tion to the boundary by a factor of up to 2. In dependence of the domain size, the resolution in the
inner region can be increased by up to a factor of 1.5 this way, without increasing the number of
computational cells.
An example of a computational grid with such a varying grid size is shown in Figure 1. The
length of the domain is 7.68 bubble diameter and resolved with 256 cells. The total domain width is
6 bubble diameters with a total of 128 cells. The inner domain has a width of 1.68 bubble diameter
and consists of 56 cells in this direction. The outer region has on each side a width of 2.16 bubble
diameter and 36 cells. Thus, the resolution in the inner domain is 0:3 mm=cell and in the outer domain
is 0:6 mm=cell. A uniform grid with domain size of 6 bubble "diameter" and 128 grid cells would have
a resolution of 0:47 mm=cell. So, using this nonuniform grid, the resolution in the inner region can be
increased by a factor of 1.5.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
194 D. WEIRICH, M. KÖHNE AND D. BOTHE

Figure 1. Example.

In simulations over a significant amount of physical time, the bubble can leave the computational
domain. In this work, we use the so-called window technique to shift the computational domain
along with the rising bubble. When the bubble’s center of mass has moved by more than the width
of a computational cell in rise direction, all field values are copied in such a way that the bubble
remains at the original position. At the upper domain boundary, the velocities in the first layer of
grid cells are set to zero. This way, in principle, simulations over arbitrary physical durations are
possible without the bubble leaving the computational domain.

4. RESULTS AND DISCUSSION

The shape and terminal velocity of a rising bubble can be characterized by the dimensionless Morton
number and Eötvös number. The Morton number is defined as


Mo D g 4c (12)
c2  3

and depends only on the material properties of the fluid and the gas. The Eötvös number is
defined as

 de2
Eo D g (13)

p
with the equivalent bubble diameter de D 3 6V = . Besides the fluid properties, the Eötvös num-
ber only depends on the equivalent bubble diameter. Therefore, instead of the Eötvös number, the
equivalent bubble diameter can also be used to characterize rising bubbles. This allows for a more
direct comparison to experimental values.
To test and study the boundary conditions described in Section 3, we simulate the rise of air bub-
bles with diameters ranging from 4 mm to 10 mm in liquids of different viscosities. We investigate
the dependence of the simulation results on the size of the computational domain and compare it
with simulations performed with slip boundary conditions. Furthermore, the results are compared
with experiments and analytical solutions.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 195

For the calculation of terminal rise velocities and bubble shapes, we use a resolution of 16 cells
per bubble diameter. Test simulations with higher resolutions yield results with the same terminal
velocities, and also, the velocity fields outside the gas bubble show good agreement. Only for the
velocity fields inside the gas bubble a further improvement of the resolution would be necessary.
Experimental investigations of rise velocities and bubble shapes in water-glycerol mixtures with
different viscosities have been carried out by Raymond and Rossant [28]. Simulation results are
compared with these experimental data. The relevant material parameters for the liquid phase
in these systems are shown in Table I. The density of air is taken as d D 1:122 kg=m3 and
the viscosity as d D 1:824  105 Pa  s. We consider several of the fluid systems S1 to S8
investigated in [28].

4.1. Terminal rise velocities


First, we consider the overall center of mass motion of the fluid particle. In numerical simulations,
the terminal velocity of the bubble not only depends on the physical properties of the surrounding
liquid but also depends critically on numerical parameters like the domain size or the resolution of
the underlying grid.
In Figure 2, the dependence of the terminal velocity on the domain size is plotted for slip bound-
ary conditions at the lateral domain boundaries and for the novel ABC. The system used for this
calculation is S1; the bubble diameter is 4 mm. With a Reynolds number of 0.17, we can compare
the results to the analytical solution of Hadamard and Rybczynski, for which
2.c  /gR2 c C
U1 D : (14)
3 c 2 c C 3
With the data for S1, we expect a rising velocity of 2.34 cm/s for the 4 mm bubble.
Both boundary conditions predict nearly the same terminal rise velocity at a domain size of 32
bubble diameters. The terminal velocity at this domain size is about 1% under the value predicted
from Hadamard and Rybczynski. However, for a domain size of 8 bubble diameter, slip boundary
conditions underpredict the terminal velocity by about 10% whereas the ABC produces an error of

Table I. Physical properties of water-glycerol mixture, systems taken from [28].

Viscosity Density Surface tension


System Water .Pa  s/  .kg=m3 /  .N=m/ Mo
S1 traces 0.7 1250 0.063 7
S3 7% 0.24 1245 0.063 1.4
S5 18 % 0.075 1205 0.064 9  104
S6 24 % 0.042 1190 0.064 1  104

Figure 2. Terminal velocity of a 4 mm bubble in system S1 for different domain sizes.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
196 D. WEIRICH, M. KÖHNE AND D. BOTHE

about 4% and thus has still a good accuracy, slightly better than calculations with slip boundary
condition and 16 bubble diameter domain size. Even with a domain size of 4 bubble diameters, the
ABCs give better results than slip boundary conditions with a domain size of 8 bubble diameters.
Thus, in the following computations, to keep the computational effort affordable, a domain size of
8 bubble diameters is used for calculations with the ABCs.
Studying the dependence of the terminal rise velocity on the grid resolution for the system S1
and 4 mm bubble diameter shows no resolution dependency at all for the considered meshes. Even
with a rough resolution of 4 cells per bubble diameter (0:1 cm=cell), we obtain the expected terminal
velocity of 2:34 cm=cell with an accuracy better than 1%. However, with smaller Morton numbers or
larger bubbles, the resolution must be increased to obtain grid independent results. Figure 3 shows
a resolution study performed for the system S5 and a bubble with 10 cm equivalent diameter using
ABCs and slip boundary conditions with varying domain sizes.
On coarse grids too, the terminal rise velocity is always underestimated. But simulations with
ABC in general show higher velocities than simulations with slip boundary conditions. On the finest
grid with a resolution of 32 cells per bubble diameter, the terminal velocity is underpredicted by
less than 2% with ABCs compared with experimental values, almost independent of the domain
size. Slip boundary conditions on the other side underpredict the terminal velocity by more than 4%
using a domain size of 8 bubble diameters and by 10% using a domain size of 4 bubble diameters.
A comparison of numerical and experimental results is shown in Figure 4 for bubble diameters
ranging from 4 mm to 10 mm and the systems S1– S6 described in Table I. Raymond and Rosant
[28] give a relative error for their velocity measurement of ˙2%. As numerical simulations for the
bubbles with 6, 8, and 10 bubble diameter in system S5 and 6 and 8 bubble diameter in System S6

Figure 3. Terminal velocity of a 10 mm bubble in system S5 for different resolutions and boundary
conditions.

Figure 4. Comparison of terminal velocities in experiment and numerical simulation with artificial boundary
condition.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 197

Figure 5. Comparison of bubble aspect ratios from experiments with numerical simulations using the
artificial boundary condition.

have the highest demand concerning grid resolution, in these simulations, a deformed grid as shown
in Figure 1 is used to improve the resolution near the bubble.
Numerical simulations for systems with small terminal velocities, that is, higher Morton num-
ber or smaller bubbles, show good agreement with experimental results. However, in systems with
higher velocities (small Morton numbers or large bubbles), the terminal velocity is underestimated
by up to 2%. Specific calculations with slip boundary conditions but larger domain size show the
same results. Therefore, it is reasonable to assume that the ABC keeps the physics of the bubble
rise intact.

4.2. Bubble shapes


To investigate the bubble shapes in the numerical simulations, we consider the bubble aspect ratio,
defined as the ratio of bubble height to bubble width. In all systems, the bubble has the shape of an
oblate sphere. In Figure 5, we compare the numerical results to the experimental data from Raymond
and Rosant [28]. Like in the results for the terminal velocity, the agreement is quite good for higher
Morton numbers and smaller bubbles. With smaller Morton numbers and larger bubbles, the aspect
ratio in the numerical simulations is higher than the experimentally measured, that is, bubbles are
more deformed in the experiments.
As the increased deformation comes from increased inertial forces related to the higher terminal
velocities, the underestimation of the deformation in numerical simulations is not surprising and can
be explained by the underestimation of the terminal velocity because of insufficient grid resolution.

4.3. Velocity fields


The analytical solution of the two-phase Stokes equations for a rising gas bubble by Hadamard [29]
and Rybczynski [30] is valid in the limit of Re ! 0 (spherical bubble, no inertial effects). This
solution was extended for slightly deformed bubbles by Taylor and Acrivos [31]. Employing sin-
gular perturbation theory, they obtained first-order corrections with respect to the Reynolds number
thus taking small inertial effects into account.
For the system S1 and the 4 mm bubble, the Reynolds number is about 0.17 and thus can be
compared with the analytical solutions. With an aspect ratio of 0.99, the computed bubble shape is
also nearly spherical.
A comparison of streamlines in the surrounding liquid for calculations with a domain size of 4
bubble diameters to the Hadamard–Rybczynski solution is shown in Figure 6, both for the ABC
and the slip boundary condition. The black lines on both sides of the picture correspond to the
Hadamard–Rybczynski reference solution. On the left side of the picture a simulation with slip
boundary condition is shown in red; on the right side of the picture the solution with ABC is shown
in red. A very good overall agreement is achieved with the ABC. Even near the lateral boundaries,
nearly no deviation can be observed. Minor disturbances occur only near the inflow boundary.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
198 D. WEIRICH, M. KÖHNE AND D. BOTHE

Figure 6. Comparison of streamlines: (black) reference solution, (left) slip boundary condition, and (right)
artificial boundary condition.

Slip boundary conditions on the other hand result in a flow that is almost parallel to the domain
boundary near the boundary, thus showing a strong distortion there. Even near the bubble, the
streamlines deviate noticeably from the reference solution. Boundary conditions hence have a sig-
nificant impact on the flow behavior in the near field around the bubble and can severely influence
simulation results.
A more detailed comparison can be made concerning the velocity profiles. In Figures 7 and 8,
the velocity profile is plotted on a cut in direction normal to the boundary 1C taken at the rear end
of the bubble. The velocity component u3 in rise direction is plotted in Figure 7, and the velocity
component u1 in direction normal to the boundary 1C in Figure 8.
The agreement of the simulation results with the analytical Hadamard–Rybczynski solution is
quite good for the velocity component u3 in the rising direction of the bubble for both slip boundary
condition and ABC. Only near the domain boundary a small deviation can be observed for the
simulation with slip boundary condition, while the agreement of the ABC stays good even near the
domain boundary.
In contrast to the velocity component u3 , a larger deviation to the analytical results can be
observed for the normal velocity component u1 at the boundary 1C for the simulation with slip
boundary condition. Near the boundary, the normal components of the velocity field are damped to
zero by the slip boundary condition. In the simulation with the ABC, on the other hand, no damping
of the velocity field from boundary conditions can be observed. Even near the domain boundary, the
agreement with the analytical solution is quite good.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 199

Figure 7. Comparison of velocity component u3 in a plane at the rear end of bubble.

Figure 8. Comparison of velocity component u1 in a plane at the rear end of bubble.

Figure 9. Comparison of velocity component u3 in a plane one bubble diameter behind the bubble.

In Figure 9, the velocity profile for the velocity component u3 is plotted at a distance of one
bubble diameter behind the bubble. The corresponding plot of the velocity component u1 can be
found in Figure 10. Here, in the simulation with slip boundary condition the velocity profile is
underpredicted for both velocity components u3 and u1 compared with the analytical solution. The
velocity component u3 in rising direction of the bubble is underpredicted over the complete range
by more than 5%. The velocity component u1 in normal direction of the artificial boundary 1C is,
as expected, completely damped to zero at the artificial boundary. But also in the center of the com-
putational domain a noticeable deviation of the velocity profile is seen. However, for both velocity

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
200 D. WEIRICH, M. KÖHNE AND D. BOTHE

Figure 10. Comparison of velocity component u1 in a plane one bubble diameter behind the bubble.

Figure 11. Comparison of velocity component u3 in a plane two bubble diameters behind the bubble.

Figure 12. Comparison of velocity component u1 in a plane two bubble diameters behind the bubble.

components the simulation with the ABC shows only minor deviations near the domain boundary
compared to the analytical solutions.
The velocity profile in a distance of two bubble diameters behind the bubble is shown in
Figures 11 and 12. Again, in the simulation with slip boundary condition, a lower velocity profile
can be observed compared with the analytical solutions. Compared with the velocity profiles at a
distance of one bubble diameter in Figures 9 and 10, the deviation of the simulation results to the
analytical solution is even larger. Especially for the normal velocity component u1 , the deviation of
the results are big. With the ABC, again, only minor deviations in the near of the domain boundary
can be observed.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 201

The far-field behavior of the velocity profile can be compared in Figures 13 and 14. At a distance
of 4 bubble diameters behind the bubble, the velocity component u3 is plotted in Figure 13, and u1
is plotted in Figure 14. With increasing distance, simulations with slip boundary condition show a
growing deviation of the velocity component u3 in the complete domain. At a distance of 4 bubble
diameters, the error in the velocity profile is larger than 10%. Simulations with ABC do not show
similar deviations. In the center of the computational domain, almost no deviation of the velocity
profile can be observed even at a distance of 4 bubble diameters. Only at the domain boundary, a
small deviation of simulation results from the analytical results occurs. But even at a distance of
4 bubble diameters, the differences between simulations with ABC to analytical solutions is less
than 2%.
Also for the velocity component u1 , simulations with slip boundary condition show a strongly
growing deviation of the results with increasing distance to the bubble. At the distance of 4 bubble
diameters, the velocity component u1 is almost completely damped to zero over the full compu-
tational domain. Compared with slip boundary condition, the ABC shows a significant reduced
damping of the velocity profile. At a distance of 4 bubble diameters behind the bubble, for the first
time, a significant deviation of the velocity profile is seen in the velocity component u1 . However,
in contrast to results from simulations with slip boundary conditions, with the ABC, the velocity
profile is still similar and close to the velocity profile of the analytical solution.
To compare the deviation of the complete velocity field from both simulations to the Hadamard–
Rybczynski solution, we define the relative field error in the computational domain  according to

v D kvsim  vHR k=kvHR k (15)

Figure 13. Comparison of velocity component u3 in a plane 4 bubble diameters behind the bubble.

Figure 14. Comparison of velocity component u1 in a plane 4 bubble diameters behind the bubble.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
202 D. WEIRICH, M. KÖHNE AND D. BOTHE

Figure 15. Velocity field error in dependence on the domain size.

with vHR denoting the velocity field of the Hadamard–Rybczynski solution and vsim the numerically
computed velocity field. The norm of the velocity field is defined as
Xp
kvk D v.xi /2 :
xi 2

The resulting velocity error for slip boundary condition and ABC for domain size ranging from
4 to 32 bubble diameter is shown in Figure 15. With increasing domain size the field error for
both boundary conditions is shrinking and reaches nearly the same value. For smaller domain sizes,
however, the error increases stronger for the slip boundary condition. For a domain size of 4 bubble
diameters, the field error is about 8% for the slip boundary condition but less than 4% for the ABC.
With a domain size of 8 bubble diameters, the slip boundary conditions result in a field error of
about 3%, and the ABCs in a field error of about 1%.

5. CONCLUSION

Comparison of simulations with the novel ABC with simulations performed with slip boundary con-
dition show good agreement concerning the terminal rise velocity of rising bubbles for domain sizes
larger than 16 bubble diameter. For smaller computational domains, the results differ significantly,
and the ABC leads to better results comparable with those performed with slip boundary conditions
on a domain size increased by a factor of two.
For systems with a Reynolds number from 0.17 up to about 50 and a Morton number ranging from
7 to 1  104 , we have validated the simulation results with experimental values. Good agreement
is achieved for the terminal velocity and the aspect ratios of the bubble shapes. For higher viscous
liquids the agreement is better.
In the simulation for the smallest Reynolds number, the velocity field is compared with the
analytical solutions done by Hadamard–Rybczynski and Taylor–Acrivos. The velocity field in the
surrounding liquid fits well to the numerical results, the inner gaseous phase of the bubble however
compares only qualitatively with the analytical solutions because of the low viscosity there. The
velocity field from simulations with the ABC matches the analytical solutions very accurately, while
the results with slip condition show significant deviations.
With the boundary condition presented here, the impact of domain boundaries on simulation
results with small domain size is reduced. This helps to reduce computational costs of numerical
simulations and improves the gained accuracy. Furthermore, the ABC can be combined with other
computational techniques such as the moving window technique or the refinement of the grid
resolution near the rising bubble. Both methods have been combined with the ABC to improve

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
ARTIFICIAL BOUNDARY CONDITION FOR FLOW AROUND SINGLE FLUID PARTICLES 203

simulation results. Beside the use of a VOF approach in this paper, the ABC presented here can also
be used with other approaches such as level-set or front-tracking methods. In general, the boundary
condition is not restricted to the FV method but can be used also for numerical simulations with
finite difference or finite element methods.
In future work, the ABC presented here will be applied to the simulation of mass transfer from
single rising gas bubbles. Mass transfer depends strongly on the exact form of the bubble and the
hydrodynamics in the surrounding liquid, especially on the rise velocity. Moreover, the thin con-
centration boundary layer at the interface requires a fine resolution. Thus, it is important to have
a high resolution close to the fluid particle to obtain physically sound results. The presented ABC
can reduce the impact from domain boundaries and allows the usage of smaller domain sizes, thus
improving the possible grid resolution.

ACKNOWLEDGEMENT

The authors gratefully acknowledge financial support by the Deutsche Forschungsgemeinschaft


(DFG) within the cluster of excellence EXC 259 ‘Center of Smart Interfaces’.

REFERENCES
1. Engquist B, Majda A. Absorbing boundary conditions for the numerical simulation of waves. Mathematics of
Computation 1977; 31:629–651.
2. Keller JB, Givoli D. Exact non-reflecting boundary conditions. Journal of Computational Physics 1989; 82:172–192.
3. Halpern L. Artificial boundary condition for incompletely parabolic perturbations of hyperbolic systems. SIAM
Journal on Mathematical Analysis 1991; 22:1256–1283.
4. Tourrette L. Artificial boundary conditions for the linearized compressible Navier–Stokes equations. Journal of
Computational Physics 1997; 137:1–37.
5. Jin G, Braza M. A nonreflecting outlet boundary condition for incompressible unsteady Navier–Stokes calculations.
Journal of Computational Physics 1993; 107:239–253.
6. Dorodnicyn LV. Nonreflecting Boundary conditions and numerical simulation of external flows. Computational
Mathematics and Mathematical Physics 2011; 51:143–159.
7. Hagstrom T. Asymptotic boundary conditions for dissipative waves: general theory. Mathematics of Computation
1991; 56:589–606.
8. Sa J-Y, Chang KS. Far-field stream function condition for two-dimensional incompressible flows. Journal of
Computational Physics 1990; 91:398–412.
9. Toulopoulos I, Ekaterinaris JA. Artificial boundary conditions for the numerical solution of the Euler equations by
the discontinuous Galerkin method. Journal of Computational Physics 2011; 230:5974–5995.
10. Papanastasiou TC, Malamataris N, Ellwood K. A new outflow boundary condition. International Journal for
Numerical Methods in Fluids 1992; 14:587–608.
11. Renardy M. Imposing ‘no’ boundary condition at outflow: why does it work? International Journal for Numerical
Methods in Fluids 1997; 24:413–417.
12. Heywood JG, Rannacher R, Turek S. Artificial boundaries and flux and pressure conditions for the incompressible
Navier–Stokes equations. International Journal for Numerical Methods in Fluids 1996; 22:325–352.
13. Bao W. Artificial boundary conditions for incompressible Navier–Stokes equations: a well-posed result. Computer
Methods in Applied Mechanics and Engineering 2000; 188:595–611.
14. Hu FQ, Li XD, Lin DK. Absorbing boundary conditions for nonlinear Euler and Navier–Stokes equations based on
the perfectly matched layer technique. Journal of Computational Physics 2008; 227:4398–4424.
15. Schlatter P, Adams NA, Kleiser L. A windowing method for periodic inflow/outflow boundary treatment of non-
periodic flows. Journal of Computational Physics 2005; 206:505–535.
16. Coclici CA, Sofronov IL, Wendland WL. The far-field modelling of transonic compressible flows. Mathematica
Bohemica 2001; 2:293–305.
17. Heuveline V, Wittwer P. Adaptive boundary conditions for exterior stationary flows in three dimensions. Journal of
Mathematical Fluid Mechanics 2010; 12:554–575.
18. Hirt CW, Nichols BD. Volume of fluid (VOF) method for dynamic boundaries. Journal of Computational Physics
1981; 39:201–225.
19. Rider WJ, Kothe DB. Reconstructing volume tracking. Journal of Computational Physics 1998; 141(2):112–152.
20. Focke C, Bothe D. Direct numerical simulation of binary off-center collisions of shear thinning droplets at high
Weber numbers. Physics of Fluids 2012; 24:073105.
21. Rieber M. Numerische Modellierung der Dynamik freier Grenzflächen in Zweiphasenströmungen. PhD thesis,
Institut für Thermodynamik der Luft- und Raumfahrt, Universität Stuttgart, Germany, 2004.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld
204 D. WEIRICH, M. KÖHNE AND D. BOTHE

22. Brackbill JU, Kothe DB, Zemach C. A continuum method for modeling surface tension. Journal of Computational
Physics 1992; 100:335–354.
23. Lafaurie B, Nardone CN, Scardovelli R, Zaleski S. Modeling merging and fragmentation in multiphase flows with
surfer. Journal of Computational Physics 1994; 113:134–147.
24. Francois MM, Cummins SJ, Dendy ED, Kothe DB, Sicilian JM, Williams MW. A balanced-force algorithm for con-
tinuous and sharp interfacial surface tension models within a volume tracking framework. Journal of Computational
Physics 2006; 213:141–173.
25. Popinet S. An accurate adaptive solver for surface-tension-driven interfacial flows. Journal of Computational Physics
2009; 228:5838–5866.
26. Kyed M. On the asymptotic structure of a Navier–Stokes flow past a rotating body. Journal of the Mathematical
Society of Japan 2014; 66(1):1–16.
27. Deuring P. Spatial decay of time-dependent incompressible Navier–Stokes flows with nonzero velocity at infinity.
SIAM Journal on Mathematical Analysis 2013; 45(3):1388–1421.
28. Raymond F, Rosant J. A numerical and experimental study of the terminal velocity and shape of bubbles in viscous
liquids. Chemical Engineering Science 2000; 55:943–955.
29. Hadamard JS. Mouvement permanent lent d’une sphère liquide et visqueuse dans un liquide visqueux. Comptes
Rendus des Séances de l’ Académie des Sciences 1911; 152:1735–1738.
30. Rybczynski W. Über die fortschreitende Bewegung einer flüssigen Kugel in einem zähen Medium. Bulletin
International de l’Académie des Sciences de Cracovie, Ser. A 1911:40–46.
31. Taylor TD, Acrivos A. On the deformation and drag of a falling viscous drop at low Reynolds number. Journal of
Fluid Mechanics 1964; 18:466–476.

Copyright © 2014 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Fluids 2014; 75:184–204
DOI: 10.1002/fld

You might also like