Sie sind auf Seite 1von 52

Accepted Manuscript

Direct Numerical Simulation of Mass Transfer in Bubbly Flows

D. Deising, D. Bothe, H. Marschall

PII: S0045-7930(18)30145-2
DOI: 10.1016/j.compfluid.2018.03.041
Reference: CAF 3802

To appear in: Computers and Fluids

Received date: 29 September 2017


Revised date: 19 February 2018
Accepted date: 12 March 2018

Please cite this article as: D. Deising, D. Bothe, H. Marschall, Direct Numerical Simulation of Mass
Transfer in Bubbly Flows, Computers and Fluids (2018), doi: 10.1016/j.compfluid.2018.03.041

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and
all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Highlights

• extensive parameter study of species transfer from single rising bubbles

• deduction of improved Sherwood correlation for interfacial species trans-

T
fer

IP
• adaptive mesh refinement, load balancing and multi-criterion refine-

CR
ment

• improved algebraic Volume-Of-Fluid method using iso-surface recon-


struction US
AN
• flux-based window technique/moving reference frame
M
ED
PT
CE
AC

1
ACCEPTED MANUSCRIPT

Direct Numerical Simulation of


Mass Transfer in Bubbly Flows
D. Deisinga,b , D. Bothea , H. Marschalla,∗

T
a
Mathematical Modeling and Analysis, TU Darmstadt, Darmstadt, Germany

IP
b
Engys Ltd., London, United Kingdom

CR
Abstract

US
Process design, especially in the chemical industry, often requires a detailed
understanding and accurate quantification of interfacial mass transfer. Ex-
perimental investigations, however, are often infeasible as required measure-
AN
ment quantities and locations are inaccessible. Thus, Direct Numerical Simu-
lations (DNS) have become a viable tool for the detailed study of mass trans-
M

fer processes. This paper presents a numerical modelling framework for the
simulation of dilute species transfer based on an algebraic Volume-of-Fluid
ED

method, aiming at the deduction of an improved closure model for interfa-


cial species transfer from single rising bubbles. To this end, the Continuous
PT

Species Transfer (CST) model (Deising et al., 2016) and state-of-the-art high
performance computing techniques are employed.
CE

Keywords: algebraic Volume-of-Fluid, interfacial species transfer, adaptive


mesh refinement, rising single bubbles, Sherwood correlation
AC


Corresponding author
Email address: marschall@mma.tu-darmstadt.de (H. Marschall)

Preprint submitted to Computers & Fluids April 6, 2018


ACCEPTED MANUSCRIPT

1. Introduction

In many industrial processes – especially in the chemical and bio-techno-


logical industry – two-phase gas-liquid systems are employed. Examples are

T
aerated bioreactors, which usually do not have any moving parts and aerated

IP
stirred tank reactors (Deckwer, 1985), which are stirred by one or multiple
impellers and play a major role e.g. in waste water treatment, where often

CR
continuous-flow stirred tank reactors are employed. One other two-phase
system of major industrial interest is the dispersed gas-liquid flow in bubble

US
column reactors. This reactor type is widely employed in the chemical and
biochemical industry, whenever mass transfer is needed and the underlying
AN
chemistry is comparably slow. Bubble column reactors are often utilized in
production processes of base chemicals as oxidation, hydrogenation, phos-
genation, alkylation and hydroformylation with a total annual production
M

volume of more than 108 t (Dudukovic, 2007).


A detailed understanding of the interaction between the hydrodynamics
ED

and interfacial mass transfer processes is of major importance for design and
optimization of such reactors. Experimental studies are expensive and in
PT

many cases not feasible as the measurement locations or desired measure-


ment quantities may not be accessible. Consequently, local insight into such
CE

processes by means of experiments is often infeasible. Therefore, numerical


methods for Direct Numerical Simulations of two-phase flows have become
AC

an important means for studying two-phase flow systems.


For the numerical study of mass transfer, many different methods have
been developed in recent years. These cover the physisorption of a dilute
species (Davidson and Rudman, 2002; Darmana et al., 2006; Francois and

2
ACCEPTED MANUSCRIPT

Carlson, 2010; Figueroa-Espinoza and Legendre, 2010; Haroun et al., 2010;


Aboulhasanzadeh et al., 2012; Marschall et al., 2012; Bothe and Fleckenstein,
2013; Deising et al., 2016) and the chemisorption of a dilute species (Khinast,
2001; Khinast et al., 2003; Koynov et al., 2005; Deshpande and Zimmerman,

T
2006; Darmana et al., 2007; Radl et al., 2008; Alke and Bothe, 2009; Onea

IP
et al., 2009; Bothe et al., 2011). Typically, in both cases mass transfer effects

CR
on the phase continuity and momentum balance equations are neglected.
Thus, it is emphasised here that the term species transfer implies the numer-

US
ical transport of an inert scalar and should be strictly distinguished from the
expression ’mass transfer ’. Volume effects of mass transfer are accounted for,
e.g., by Bothe and Fleckenstein (2013) and Hayashi and Tomiyama (2011).
AN
The underlying two-phase flow simulation methods to simulate species trans-
fer can be categorised into Front Tracking (Khinast, 2001; Khinast et al.,
M

2003; Koynov et al., 2005; Tukovic and Jasak, 2008; Darmana et al., 2007;
Radl et al., 2008; Aboulhasanzadeh et al., 2012) Level-Set (Yang and Mao,
ED

2005; Deshpande and Zimmerman, 2006), Volume-of-Fluid (VOF) methods


(Bothe et al., 2003, 2004; Haroun et al., 2010; Marschall et al., 2012; Bothe
PT

and Fleckenstein, 2013; Deising et al., 2016; Weiner and Bothe, 2017), and
Arbitrary Lagrangian-Eulerian (ALE) interface tracking (Lehrenfeld, 2015;
CE

Bäumler, 2014; Weber, 2016; Weber et al., 2017).


All these methods can be categorized further according to the underly-
ing mathematical model formulation and numerical discretization technique
AC

used. This immediately affects the conceptual implementation of species/-


mass transfer in the respective methods, resulting in distinct method-specific
advantages and disadvantages. The majority of models employ a single-field

3
ACCEPTED MANUSCRIPT

model formulation for an averaged concentration. The other formulation is


the two-field model, where two distinct concentration fields – one for each
phase – are present and the transport equations for both phases are coupled
via interfacial transmission and jump conditions.

T
This paper aims at the deduction of an improved closure model for the

IP
interfacial transfer of a dilute species from single rising bubbles. This re-

CR
quires the conduction of a comprehensive 3D parameter study which comes
at high computational costs but enables a detailed insight into the underly-

US
ing physical processes of species transfer at single rising bubbles. Unlike in
previous numerical and experimental investigations, in this work the effects
of changes in bubble area and in the local concentration gradients onto the
AN
species transfer are studied separately, allowing of further enhancement of
existing mass transfer correlations.
M

A major challenge in the Direct Numerical Simulation of two-phase gas


liquid flows arises from the separation of characteristic hydrodynamic and
ED

mass transfer length scales due to the thin concentration boundary layer in
close vicinity to the gas-liquid interface. In order to alleviate this problem,
PT

local dynamic adaptive mesh refinement and dynamic load balancing tech-
niques for massively parallel computations using a domain decomposition
CE

approach are employed in this work.


The implementation is accomplished in the C++ library OpenFOAM-4.x.
AC

2. Mathematical and Numerical Modelling

We employ an algebraic VOF method in combination with the Continuous


Species Transfer (CST) method, a single-field model for the direct numerical

4
ACCEPTED MANUSCRIPT

simulation of interfacial species transfer. Method details are published in


Marschall et al. (2012); Deising et al. (2016); Hill et al. (2017), where the
first two publications cover method derivation and validation and the third
one describes the utilized discretization approach on unstructured meshes

T
of general topology. For the sake of brevity, we only provide an executive

IP
summary of the underlying mathematical model and numerical method.

CR
2.1. Sharp interface continuum model

Basis of the utilised modelling framework is the so-called sharp inter-

US
face model, in which material properties are assumed to experience a sharp
discontinuous jump at the location of the phase-separating interface. The
AN
interface Σ (t) is herein assumed to be a surface (i.e. with zero thickness),
which separates two immiscible Newtonian and incompressible fluid phases.
M

We further restrict our considerations to isothermal conditions (i.e., disregard


phase change due to evaporation or condensation) and to dilute, chemically
ED

inert multi-component systems (i.e., disregard chemical reactions) of a pure


system (i.e., in particular void of contamination by surface active agents).
Under above assumptions, the balances for mass and linear momentum
PT

read
CE

∇· u = 0 in Ω\Σ(t) , (1)

∂t (ρu) + ∇· (ρu ⊗ u) = −∇p + ∇· τ + ρg in Ω\Σ(t) , (2)


AC

where Ω = Ω+ (t) ∪ Ω− (t) and Ω+/− (t) denote the respective bulk regions.
For the considered Newtonian fluid flow, the viscous stress tensor reads
 
T
τ = η ∇u + (∇u) with the dynamic fluid viscosity η. The corresponding

5
ACCEPTED MANUSCRIPT

interface jump conditions are valid at the fluid interface Σ (t) and read

JuK = 0 at Σ(t) , (3)

JpI − τ K · nΣ = σκnΣ at Σ(t) , (4)

T
where J·K denotes the common jump bracket, nΣ the outer unit vector normal

IP
to the interface and κ = ∇Σ · (−nΣ ) twice the mean interface curvature.

CR
To account for the transfer of a dilute species i, we introduce an additional
set of so-called inert scalar transport quantities for the species concentrations.

US
The corresponding set of transport equations in the bulk phases read as

∂tci + ∇· (ci u) = ∇· (Di ∇ci ) in Ω\Σ(t) (5)


AN
with the interfacial jump conditions

J−Di ∇ci K · nΣ = 0 at Σ(t) , (6)


M

c+ −
i = Hci at Σ(t) , (7)
ED

where c+
i and ci denote the local concentrations adjacent to the interface in

Ω+ (t) and Ω− (t) respectively. The above set of equations for the transport
PT

of species concentration is valid under the assumption that the interface does
not hold mass. The last jump condition represents the so-called Henry law,
CE

describing the concentration jump at the interface, assuming local thermo-


dynamic equilibrium. The Henry coefficient is set to be constant, which is a
valid approximation for most disperse gas-liquid flows.
AC

2.2. Conditional volume averaging


As the conditional volume averaging (CVA) technique builds the foun-
dation for numerical modelling in the scope of one-field formulations within

6
ACCEPTED MANUSCRIPT

the Finite Volume framework, the method is shortly presented here: to ar-
rive at consistent single field models, application of the CVA technique to
the local instantaneous governing equations of the mathematical model is
indispensable. Application of this technique consists of two steps. The equa-

T
tion is first ’conditioned’ for phase discrimination by multiplication with the

IP
phase-indicator function χ1 (respectively χ2 = 1 − χ1 ), which renders the

CR
equation valid throughout the entire computational domain. In a second
step, volume-averaging is applied to this conditioned equation.

US
The phase indicator, introduced to distinguish between the two phases 1
and 2, separated by a fluid interface SΣ (x, t), is defined as

AN

1 if x ∈ Ω1 (t) at time t
χ1 (x, t) ≡ (8)

0 otherwise.
M

The jump set of χ1 (x, t) is assumed to be sufficiently regular, so that in a


mathematically weak sense (Drew, 1983; Evans and Gariepy, 1992)
ED

∇χ1 (x, t)
nΣ = . (9)
|∇χ1 (x, t) |
The spatial derivative of the indicator function is thus well-defined and can
PT

be expressed in terms of the interface normal vector nΣ pointing from phase 1


into phase 2 and the interface delta function δΣ as ∇χ1 = −nΣ δΣ . Introducing
CE

the notation for volume-, phasic- and surface-averaged quantities


Z
Φ :=
1
Φ (x + η, t) dη χ1 ≡ |V1 | =: α1 , (10)
AC

|V | V |V |
Z
1 1
Φ := Φ (x + η, t) dη and (11)
|V1 | V1
Z Z
z}|{ 1 1 1
Φ := Φ(x, t) dS = lim Φ(x, t) dS (12)
|SΣ | SΣ aΣ δV →0 δV SΣ

7
ACCEPTED MANUSCRIPT

with aΣ representing the interfacial area density, and applying Gauss and
Leibnitz rules to derivatives (Drew and Passman, 1999; Jakobsen, 2008), the
conditional volume-averaged equation for species transfer with respect to
phase 1 then reads

T
IP
χ1 ∂t c + χ1 ∇· (cu) = χ1 ∇· (D∇c) + χ1 R (13)
  
1 1 z }| {1

CR
⇔ ∂t(α1 c1 ) + ∇· α1 c1 u1 = ∇· α1 D ∇c1 + α1 R − D∇c · nΣ aΣ .

Analogously, volume-averaging of the general jump condition yields

z }| {1

z}|{
c = H z}|{
1
c .
2
US
z }| {2
D∇c · nΣ aΣ + D∇c · nΣ aΣ = 0 , (14)

(15)
AN

Above equations are the basis for the derivation of a consistent one-field for-
mulation. The volume-averaged jump conditions are herein used for closure
M

modelling of the interfacial terms. More details can be found in Deising et al.
(2016).
ED

Application of the conditional volume averaging technique to the local


balance equations of mass, momentum and the interface transport equation
PT

lead to the well-known set of VOF equations

∇· u = 0 (16)
CE

∂t α1 + ∇· (α1 u) = 0 (17)
z}|{
(18)
AC

∂t ( ρ u) + ∇· ( ρ u u) = −∇hpi + ρg + ∇· hτ i + fΣ aΣ ,

z}|{
where h· · ·i denotes mean (or mixture) values and fΣ aΣ denotes the volu-
metric force density due to surface tension. Here, we employ the continuous
surface force model (CSF) of Brackbill et al. (1992). Note, that while above

8
ACCEPTED MANUSCRIPT

VOF equations resemble the local instantaneous balance equations, they are
indeed mixture equations derived using the CVA technique.

2.3. Single-field formulation for species transport

T
Employing the immersed interface concept which is based on Peskin

IP
(1977), Equation (13) is summed up over both phases 1 and 2 in order to ob-

CR
tain a single field formulation for interfacial species transfer. Exploiting the
surface averaged interfacial jump condition (Eqn. 14), defining the mixture
concentration c ≡ α1 c1 +α2 c2 and ignoring chemical reactions, the following

US
– yet unclosed – single field model is derived
 1 2

AN
∂t c + ∇· ( c u) = ∇· α1 D ∇c1 + α2 D ∇c2 . (19)

The basic idea to arrive at a closed single field formulation for interfacial
M

species transfer, which is readily usable for Finite Volume discretization, is to


relate the phasic averaged concentrations at the interface employing Henry’s
ED

law. The local instantaneous concentration jump at the interface is given


in Equation (7), connecting the one-sided concentrations of both sides of
the interface. To employ this law to phasic averaged concentrations, Haroun
PT

et al. (2010) and Marschall et al. (2012) introduce the follwing approximation
z}|{
CE

1
c+ c c1
H = − = z}|{2 ≈ 2 . (20)
c c c

By inserting (20) into the definition of the mixture concentration c, we can


AC

now reformulate the phasic averaged concentrations in terms of the mixture

9
ACCEPTED MANUSCRIPT

concentration and volumetric phase fraction as


 
1 2 1 1 − α1
c := α1 c + α2 c ≈ c α1 +
H
c 1 c
c1 ≈ 1−α1 , c2 ≈ . (21)

T
α1 + H H α1 + 1−α
H
1

IP
Based on above closure relation, it was found in Deising et al. (2016) that
the r.h.s. of (19) can be rewritten in different ways, depending on the

CR
definition of the mean diffusion coefficient which can be chosen arbitrar-
ily. The mean value calculations typically applied are the arithmetic or

cient hDih ≡
1
D D
2
2

α1 D +α2 D
1,
US
harmonic mean. When choosing a harmonic mean mixture diffusion coeffi-
the following single field formulation for interfacial
AN
species transfer is obtained:
 
1 − H1
∂t c + ∇· ( c u) = ∇· (hDih ∇c) − ∇· hDih c∇α1 . (22)
α1 + 1−α 1
M

This equation was obtained in Haroun et al. (2010) and Deising et al. (2016)
independently, using different mathematical derivations. It is further shown
ED

in Deising et al. (2016), that a different formulation can be obtained when


defining an arithmetic mean mixture diffusion coefficient, which leads to
PT

    
1 2 1
∂t c + ∇· ( c u) = ∇· (hDia ∇c) + ∇· D − D α1 − 1 ∇c
α1 + 1−α
H
1

" !! #
CE

1 2 2
c 1 D −D 1 D
+ ∇· 1−α1 − D − ∇α1 .
α1 + 1−α
H
1
H α1 + H
H
(23)
AC

The model formulation in Equation 22 has proven to be beneficial in terms


of accuracy, being second order accurate for a Henry coefficient H ≡ 1 and
reducing to first order for H → ∞, due to the closure relation (21), while
(23) is generally of first order. For typically high Henry coefficients, both

10
ACCEPTED MANUSCRIPT

methods yield very similar results. In under-resolved cases, the derived sin-
gle field species transfer models, termed Continuous Species Transfer (CST)
model(s), will lead to an overprediction of the species transfer. For details, it
is referred to Marschall et al. (2012); Deising et al. (2016); Hill et al. (2017).

T
IP
3. Numerical Solution Procedure

CR
3.1. Algebraic VOF

The employed two-phase solver is based on the interFoam solver but con-

US
tains a series of enhancements and additions. To briefly introduce the utilized
interFoam solver, some comments are given here on the employed algorithm
AN
and available literature. Main focus in this introduction is the flux-based
formulation and the algebraic advection algorithm.
The interFoam solver is based on the VOF equations (16 – 18) and suit-
M

able for the numerical simulation of immiscible incompressible two-phase


flows of Newtonian fluids. The implementation of the governing equations
ED

in the interFoam solver resembles (16 – 18) – due to the utilized technique
of equation mimicking. However, the solver utilizes a flux-based formulation
PT

for the set of equations, i.e. does not solve for the velocity u but for the face
flux defined as Ff := Sf · u f . The discretization of the surface tension force
CE

(CSF model) is flux-based, leading to a formally balanced-force algorithm.


Descriptions of the algorithm and the solver in general can be found e.g. in
AC

Ubbink (1997); Rusche (2002); Deshpande et al. (2012).


The interFoam solver employs an algebraic VOF method, meaning that
the transport of the volumetric phase fraction field α is handled by solving the
respective transport equation numerically without reconstructing the inter-

11
ACCEPTED MANUSCRIPT

face. Algebraic VOF methods are widely employed also in commercial CFD
codes as Ansys CFX and Fluent but usually exhibit difficulties proprietary
with the transport of sharp fields. The algebraic advection in the solver-
family interFoam utilizes a flux-corrected transport (FCT) method (Zalesak,

T
1979) called MULES (multidimensional universal limiter with explicit solu-

IP
tion) (Weller, 2006). This FCT is unique in the way the limiter is calculated

CR
at each cell face, using an iterative approach. However, unlike the standard
interFoam solver, this work utilizes the Interface Capturing Scheme CICSAM

US
(Ubbink, 1997) in a blended manner (Deising et al., 2016) for advection of
the phase fraction and concentration fields due to its advantage in accuracy
and consistent advection.
AN

3.2. Hydrodynamic solver modification & validation


3.2.1. Moving reference frames
M

To simulate single rising bubbles it is sufficient, also efficient and thus


ED

common practice to employ methods that allow to follow the bubble during
the simulation. To achieve this, one can either move the computational do-
main with the bubble (Moving Window Technique) as e.g. Bothe and Flecken-
PT

stein (2013); Weiner and Bothe (2017), while observing the flow system from
the same inertial frame, or transform the governing equations into a frame
CE

of reference in which the bubble barycentre is stationary (Moving Reference


Frame (MRF) technique). Application of the Moving Window Technique
AC

to unstructured hexahedral meshes is not straightforward. On unstructured


meshes with polyhedra of general shape, this approach is even no longer
viable. A suitable approach on unstructured meshes is a flux-based win-
dow technique introduced in our previous work (Deising et al., 2016), named

12
ACCEPTED MANUSCRIPT

Inertial Reference Frame (IRF) technique, where only the phase fraction ad-
vection equation is transformed into the moving reference frame.
A sound theoretical basis for the transformation of equation systems into
reference frames can be found in Muschik and Restuccia (2001); Essén (2002)

T
and an example for a concrete implementation in terms of a non-constant

IP
non-rotating frame is given e.g. in Rusche (2002); Cariglino (2013)

CR
For validation, the two different reference frame formulations have been
compared against each other for a path-instable rising bubble at Eo = 1.2

US
and Mo = 1.267e-10, resulting in Remax ≈ 580. The simulations are con-
ducted on a relatively coarse mesh with a maximum local resolution by cells
of 1/36 bubble diameter size in the region around the interface. Figure 1
AN
shows the results of rise velocity over time and bubble trajectory. Here, the
bubble trajectory was projected onto cylindrical coordinates for easier com-
M

parison. The resulting path of the bubbles shows a zigzag behaviour for both
approaches. However, the plane in which the bubble rises is different. While
ED

the bubble trajectory is very similar for both reference frames (cf. Figure
1b), the rise velocity shows significant differences regarding its transient be-
PT

haviour (cf Fig. 1a). The change in the bubble velocity over time is about
2.5 times larger for the MRF. The cause for this effect is most probably the
CE

additional acceleration term in the momentum equation for the MRF formu-
lation. The bubble shape and velocity streamlines over time for the MRF
case are shown in Figure 1c.
AC

Comparison with Mougin and Magnaudet (2002), who investigate path


instability under very similar (although not identical) conditions, yields good
agreement for both formulations. This is because their contribution only

13
ACCEPTED MANUSCRIPT

0.35
MRF
0.3 IRF
300

vertical position [mm]


0.25
rise velocity [m/s]

0.2
200

T
0.15

IP
0.1 100

0.05 MRF
IRF

CR
0 0
0 0.2 0.4 0.6 0.8 1 1.2 −2 0 2 4 6 8 10
time [s] radial position [mm]

(a) Magnitude of bubble velocity. (b) Projected bubble trajectory.

US
AN
M

(c) Streamlines at t = 1.2 s, 1.22 s, 1.24 s, 1.26 s using MRF.


ED

Figure 1: Comparison of MRF and IRF for a path-instable single rising bubble

includes discussion of the bubble path and temporal behaviour of the lateral
PT

velocity components, which are both very similar for the MRF and IRF
with differences being less than 10%. The amplitude of the change in rise
CE

velocity can only be estimated from (Mougin and Magnaudet, 2002, Fig.
6), according to which the bubble rise velocity predicted by the IRF is in
AC

better agreement. Thus, the IRF supposedly yields more accurate results.
Also, it is numerically more stable as it avoids having a controller-dependent
acceleration term in the momentum equation.

14
ACCEPTED MANUSCRIPT

3.2.2. Improved curvature calculation


The utilized CSF model (Brackbill et al., 1992) in its flux-based formu-
lation formally allows for a balanced-force implementation which strongly

T
reduces parasitic currents in the velocity field. However, this requires an ac-
curate method for the calculation of interface curvature. The implementation

IP
in interFoam computes the curvature from the divergence of the interface nor-

CR
mal, approximated by a simple gradient computation of the volumetric phase
fraction field which is then interpolated to the face centres. This estimation

US
of the curvature is very inaccurate and deviations of the obtained to the exact
curvature are up to several hundred percent (≈ +/− 300 . . . 500 %). Applica-
tion of smoothing algorithms for the phase fraction field prior to calculating
AN
the interface normal as suggested e.g. by Ubbink (1997); Rusche (2002) are
not able to circumvent this problem and offer only a small enhancement of
M

the calculated interface normals and therefore the curvature.


On structured grids, the curvature can be estimated with sufficient accu-
ED

racy by using height functions (Popinet, 2009), which, however, is not trivial
to adapt to unstructured grid arrangements. Therefore, an iso-surface recon-
PT

struction of the phase fraction field based on the work of Batzdorf (2015)
is employed here, which reconstructs the interface as a volume fraction iso-
CE

contour at α1 = 0.5. The iso-surface is utilized to calculate the interface


normal vectors, which are smoothed and distributed in the interface area by
interpolation. Finally, the curvature is calculated as before from
AC

X
κ=− (nΣ )f,iso · Sf . (24)
f

Actual cell cutting to obtain an interface surface is not performed, the al-

15
ACCEPTED MANUSCRIPT

gorithm is only used for calculating the interface normal vectors and the
interface curvature in cells containing the interface. The reconstruction algo-
rithm of Batzdorf (2015) is incorporated in the solver framework developed
in this work and used for the sole purpose of increasing the accuracy of the

T
interface normal vector and thus the curvature calculation. Further changes

IP
on the surface tension model are not included, as the employed approach

CR
ensures balance of pressure and surface tension forces by using the same
discretization stencils for surface tension and pressure equation.

US
To prove that the underlying algorithm allows for force balancing, sim-
ulations of a 2D liquid disc in zero gravity as presented in Francois et al.
(2006) are performed. The volumetric phase fraction field is initialized nu-
AN
merically exact due to computing the phase fraction values by cutting of a
cylindrical surface with the utilized mesh, employing the library presented in
M

Maric et al. (2013). Tables 1 and 2 show the resulting maximum (spurious)
velocities for a prescribed exact curvature and calculated curvature.
ED

Table 1: Error in maximum velocity |u|max (in m/s) after one time step for the inviscid
static drop in equilibrium when the exact curvature is specified
PT

ρ1 /ρ2 interCSTFoam CSF cell-centred (Francois et al., 2006) SSF face-centred (Francois et al., 2006)
1 9.81e − 19 5.19e − 5 5.43e − 19
103 9.11e − 17 6.15e − 3 4.44e − 18
CE

105 7.17e − 15 6.91e − 3 2.71e − 19

The main question regarding hydrodynamics is whether the algebraic


AC

VOF framework is capable of accurately simulating a single rising bubble


in quiescent liquid at moderate Reynolds numbers. To this end, numerical
simulations for the water-air system are performed and compared to exper-

16
ACCEPTED MANUSCRIPT

Table 2: Effect of the fluid density ratio on the error in maximum velocity |u|max after
one time step for the viscous static drop in equilibrium

ρ1 /ρ2 interFoam interCSTFoam interCSTFoam CSF, height function SSF, convolution


isoSurface exact curvature (Francois et al., 2006) (Francois et al., 2006)

T
10 1.99e − 4 1.93e − 5 1.17e − 18 4.40e − 7 2.30e − 6

IP
103 2.47e − 3 5.03e − 4 7.14e − 17 6.22e − 7 2.59e − 6
105 2.01e − 1 5.30e − 2 5.24e − 13 6.25e − 7 2.59e − 6

CR
imental data of Duineveld (1995), who conducted a series of measurements
in an ultra-pure water system (without influence of surfactant). The investi-

US
gated range of bubble diameters is db ∈ [0.3 mm . . . 1 mm]. For the numerical
simulations, the bubble diameters db = {0.35, 0.5, 0.65, 0.85, 0.95 mm} are
AN
chosen. The comparison in Figures 2a and 2b show the results for bub-
ble rise velocity and bubble aspect ratio, respectively. For the simulations,
the blended CICSAM advection scheme and the iso-surface-based curvature
M

model are used. It can be seen that the numerical results deviate only slightly
ED

40
2 Duineveld 1995
35 Moore 1965
standard interFoam
1.8
rise velocity [cm/s]

interCSTFoam
aspect ratio [-]

30
PT

1.6
25
1.4
20 Duineveld 1995
Moore 1965 1.2
CE

15 standard interFoam
interCSTFoam 1
0.4 0.6 0.8 1 0.4 0.6 0.8 1
radius [mm] radius [mm]
AC

(a) Rise velocity. (b) Bubble shape.

Figure 2: Comparison of numerical results to measurements of Duineveld (1995)

from the experimental results. It should be noted that this relatively good

17
ACCEPTED MANUSCRIPT

agreement is obtained only by employing the improved curvature calculation.


For the original solver framework, the discrepancies in rise velocity and shape
are far more severe (cf. Figure 2 standard interFoam).
Hence, it is shown that the algebraic VOF method implemented in the

T
interFoam solver framework as utilized in this work is capable of simulating

IP
rising bubble hydrodynamics with sufficient accuracy.

CR
3.3. Parallelisation and HPC

Most gas-liquid systems encountered in nature and those utilized in indus-

US
trial processes exhibit high Peclet numbers, which means that the transport
by advection is several orders of magnitude faster than the diffusive transport,
AN
leading to extremely thin concentration boundary layers being typically 2 to
3 orders of magnitude thinner than the hydrodynamic boundary layer. The
M

limiting factor for numerical simulations is thus the Schmidt number, which
directly relates to the decrease in concentration boundary layer thickness rel-
ED

ative to the hydrodynamic boundary layer. To resolve the thin concentration


boundary layer at the fluid interface, High Performance Computing (HPC)
techniques are employed in this work, namely dynamic local adaptive mesh
PT

refinement (AMR) and dynamic load balancing. Application of AMR signifi-


cantly reduces the overall mesh size, whereas dynamic load balancing ensures
CE

efficient usage of cores in parallel computations. The latter is important since


the bubble interface and wake exhibit significant changes during the bubble
AC

rise (cf. Fig. 6a). The simultaneous application of both techniques massively
reduces the overall computation time. To quantify the speed-up (computa-
tion time reduction factor) compared to utilizing a uniform mesh of the same
resolution at the interface, a comparison for the 3D dam break tutorial case

18
ACCEPTED MANUSCRIPT

with obstacle was conducted (cf. Figure 3). The reference solution was

100

T
30
speed-up [-]

IP
10

CR
3 Uniform mesh
Linear speed-up
AMR & LB
1
1 2 4 8

US 16 32 64 128 256 512


num. of cores [-]

Figure 3: Scale-up plot for 3D dam break with obstacle using uniform mesh and AMR
AN
with load balancing

performed on a minimum of 8 cores, thus the speed-up equals 1 for this


M

simulation. To correctly assess the statistical computational cost of AMR


and load balancing, the test case was simulated for 0.5 seconds of physical
ED

time. This is essential in order to obtain a correct speed-up behaviour, as


the additional computational cost strongly depends on the number of refined
PT

cells per iteration, the mapped flux used in the pressure correction equation
and the load balancing interval. These, however, strongly depend on the
CE

local solution and thus can only be reliably assessed by simulating a physical
problem rather than by conventional HPC assessment over a few iterations.
AC

Due to the very high overall simulation time for this study (cf. Table 3),
simulations on less than 8 cores have not been performed.
In the employed OpenFOAM-library both techniques are already present
in principle but required modifications in order to function together. In the

19
ACCEPTED MANUSCRIPT

Table 3: Computational time (in hours) for uniform and dynamic mesh on different num-
bers of cores

mesh type 1 core 8 cores 16 cores 32 cores 64 cores 128 cores 512 cores
uniform mesh − 233 122.4 77.8 52.2 26.5 7.35

T
dynamic mesh 63.4 7.83 5.45 3.15 − − −

IP
recent few years, two different implementations to achieve load balancing

CR
in OpenFOAM have been accomplished, at TTD TU Darmstadt (Batzdorf,
2015), and at Purdue University (T.G. Voskuilen). The former is based on

US
the use of a special decomposition method (clustered decomposition) to main-
tain the refinement history required for AMR, while the latter includes the
load balancing directly into the mesh refinement algorithm. Above proposed
AN
methods have been unified in the scope of this work into one library frame-
work and further improved through major enhancement of the multi-criterion
M

refinement library originally introduced by T.G. Voskuilen and addition of


a surface field mapping algorithm in cooperation with M.Sc. Daniel Ret-
ED

tenmaier (TU Darmstadt). The improvements introduced here increase the


method’s robustness, usability and performance.
PT

3.3.1. Local Dynamic Adaptive Mesh Refinement


The introduced multi-criterion refinement allows the user to specify multi-
CE

ple refinement criteria and respective target refinement levels for this criterion
by means of a dictionary file, the common input format for OpenFOAM. For
AC

each criterion, minimum and maximum values of the respective quantity and
the number of maximum refinement steps can be specified. Further, domain
regions can be refined by defining geometrical shapes as boxes and spheres.
An exemplary dictionary entry, leading to the refined mesh shown in Figure

20
ACCEPTED MANUSCRIPT

dynamicFvMesh
dynamicRefineBalancedFvMesh ;
refinementControls
{
enableRefinementControl true ;
interface
(
alpha1 (2 5)
);
fields
(

T
alpha1 ( 0 . 0 1 1.1 3)
C1 ( 0 . 0 0 1 0 . 0 5 2 )
);

IP
gradients
(
alpha1 ( 0 . 0 1 2 2)
);
curls
(

CR
U ( 1 0 0 1 e +05 3 )
);
regions
(
cylinderToCell
{
p1 (0.015 0.015 0.015);
p2 (0.015 0.033 0.015);

US
radius 0.006;
}
);
}

(a) Multi-criterion refinement. (b) Species concentration


AN
Refine mesh based on vorticity, wake behind a rising air
interface position, concentration bubble simulated with local
field and region with different adaptive grid and dynamic
M

refinement levels. load balancing at t = 0.18s.


ED

Figure 4: Local dynamic adaptive mesh refinement

5b, is given in Figure 5a for the DNS of species transfer from a single bubble.
PT

Local adaptive mesh refinement leads to new cells and faces, at which
values of the respective transport quantities have to be specified. For fields
CE

stored in cell centres, a simple conservative algebraic mapping is employed,


which for refined cells copies the value of the parent cell onto each child cell
AC

and for unrefined cells performs a simple averaging onto the agglomerated
cell. For face-centred fields, however, the employed mapping algorithm in the
OpenFOAM library simply copies the value from local face zero of the parent
cell onto all newly generated faces of the child cells. As the solver relies on

21
ACCEPTED MANUSCRIPT

a divergence free volumetric flux field, the field needs to be corrected after
each AMR step in order to ensure a divergence free flux. For this, a pressure
correction step is introduced after each AMR step, which requires a good
approximation of the flux field on newly generated faces. In this work, the

T
flux field is re-computed from the face-centred velocity field uf , obtained by

IP
projecting the interpolated velocity tangential to the face and using the flux

CR
normal to the face, i.e.

Ff
uf = (I − nf ⊗ nf ) (u)f (CD) + nf , (25)

US |Sf |

which is mapped onto the newly generated faces by averaging of neighbor face
values. This procedure significantly increases the robustness of AMR com-
AN
pared to the standard flux mapping in OpenFOAM, where the face fluxes
are reconstructed from the cell-centred velocity field employing linear inter-
M

polation.
In terms of computational cost per time step, the mesh update takes
ED

about 12% of the time. However, the subsequent flux correction takes about
25−30% of the computation time per time step. The flux mapping algorithm
PT

introduced here significantly reduces that cost due to a better initial solution
and thus faster solver convergence. To maintain a reasonable computational
CE

cost while ensuring that the interface remains well within the refinement
region, AMR is executed every 20 Iteration for all simulations presented here.
Thus, the overall computational cost increase due to dynamic refinement is
AC

less than 5%.

22
ACCEPTED MANUSCRIPT

3.3.2. Dynamic Load Balancing


For efficient parallel computing, the decomposition of the mesh onto the
computation cores is modified at run time, when a user-defined imbalance

T
threshold is reached. This mainly involves the combination of two steps which
are available within the OpenFOAM framework. At first, a decomposition is

IP
created, which is basically a list of integer values that specify for each cell in

CR
the domain the respective core it belongs to. In a second step, the mesh and
all fields are communicated to the cores based on the aforementioned list.

US
However, several problems with the current load balancing method occur.
One main drawback of the mesh balancing approach introduced by (T.G.
Voskuilen) is that all simulation steps, including the pre-processing, need to
AN
be executed in parallel in order to correctly maintain the refinement history
(cell refinement level). Also, a redistribution onto a different number of
M

cores is not possible as in their introduced framework, reconstruction of the


processor data results in an incorrect refinement history. Thus, pre- and
ED

post-processing as well as the simulation have to be performed on the same


number of cores. An additional limitation is that their presented algorithm
PT

only works with the scotch decomposition method.


To alleviate this, work has been done at TTD TU Darmstadt (Batzdorf,
CE

2015) who introduced an additional clusteredDecomposition method. Essen-


tially, this decomposition ensures that all refined cells stemming from the
same base cell remain on one core (thus the term clustered ), which enables
AC

decomposition and reconstruction of the refinement history. Drawback is


that decompositions obtained by different available algorithms have to be
corrected to achieve clustering. This correction is currently only available for

23
ACCEPTED MANUSCRIPT

T
IP
CR
(a) Core distribution at time (b) Core distribution at time
t = 0.02 s.
US t = 0.17 s.

Figure 5: Dynamic load balancing using clustered hierarchical decomposition


AN
hierarchical and simple decompositions. This problem can more generically
be overcome within the current code structure of OpenFOAM by modify-
M

ing the decomposition weights of the different decomposition methods rather


than the methods themselves. Recently, this problem has been remedied
ED

within the OpenFOAM library by introducing a decomposition constraint


for the refinement history. However, the utilized approach in this work has
PT

proven to be faster and significantly more robust. To enable correct mapping


of surface fields, the existing library was modified to flip the sign of surface
CE

fields only when they are flux quantities, thus allowing to recompute the
volumetric flux field from the mapped surface velocity field.
AC

4. Simulation Results

This Section presents the results for species transfer from single rising
bubbles in a quiescent liquid (infinite medium) using the presented VOF/CST

24
ACCEPTED MANUSCRIPT

single-field model. Here, the term infinite medium refers to a sufficiently


large computational domain, such that the bubble rise velocity is not af-
fected significantly by the boundary conditions at domain boundaries. The
considered species properties range from low Schmidt and Peclet numbers

T
to moderate and high numbers (Sc ∈ {0.03, 0.1, 0.3, 1, 3, 10, 30, 100, 300},

IP
Pe ∈ [10, 104 ]), with realistic diffusion coefficient ratios (up to 104 ) and Henry

CR
coefficients (between 5 and 30).
A detailed model verification and validation study has been shown in

US
Deising et al. (2016) and thus omitted in this paper.

4.1. Parameter study


AN
Main aim of the presented work is to deduce an improved closure relation
for species transfer from rising single bubbles. To this end, an extensive
M

parameter study has been performed to obtain insight into the influence of
the Reynolds and Schmidt numbers onto the species transfer process.
ED

Making use of local adaptive mesh refinement and dynamic load balanc-
ing strategies, in this work the concentration boundary layer is sufficiently
resolved for Peclet numbers up to Pe = 104 (cf. Figure 9b), while still main-
PT

taining a somewhat reasonable overall simulation time of about 8000 to 10000


core hours per case (3 to 4 weeks on 16 cores). The maximum resolution
CE

around the interface for all cases in the parameter study is about 224 cells
per bubble diameter. The total mesh cell count is about 4·106 cells, although
AC

it depends for each case – due to the dynamic adaptive mesh refinement –
on the respective bubble shape. Figure 6 shows the concentration profiles
around rising single bubbles for two different material parameter settings.
In addition to the species concentration profiles, concentration iso-surfaces

25
ACCEPTED MANUSCRIPT

T
IP
CR
(a) Eo = 0.3, Mo = 10−11 .
US (b) Eo = 5, Mo = 10−5 .

Figure 6: Species concentration field with iso-contours and streamlines (H = 5, Pe ≈ 8000)


AN
and velocity streamlines are plotted in order to give a better insight into the
flow structure around both bubbles. It can be seen that the concentration
M

profile looks similar, while the flow around both bubbles is entirely different.
The velocity and pressure fields for case Eo = 0.3, Mo = 10−11 are given in
ED

Figure 7. Figure 8 shows the cell values (original data, non-smoothed) of the
species concentration at Pe ≈ 6300. In Figure 8b, additionally, the employed
PT

computational mesh is shown, using three levels of refinement in the vicinity


of the interface and seven refinement levels in total.
CE

4.2. Species transfer correlation

4.2.1. Species transfer modelling


AC

A common engineering approach for the modelling of species transfer in


two-phase flows is to introduce a species transfer coefficient kl :

ṁ = kl AΣ (cl,Σ − cl,∞ ) = kl Sr Asphere (cl,Σ − cl,∞ ) , (26)

26
ACCEPTED MANUSCRIPT

where AΣ denotes the interfacial area and (cl,Σ − cl,∞ ) is the liquid-sided
difference between the interface and the bulk concentrations and represents
the driving force of the interfacial species transfer. The area AΣ can be
expressed as the area of a spherical volume-equivalent bubble Asphere times

T
the surface ratio Sr := AΣ /Asphere . This approach introduces two unknowns

IP
which need modelling, the species transfer coefficient kl and the interfacial

CR
area AΣ or surface ratio Sr .
In order to calculate kl from experimental studies of mass transfer, the

US
interfacial area needs to be obtained. While for spherical bubbles, AΣ can
be obtained by visual observation and measurement of the bubble diame-
ter, the bubble area for non-spherical bubbles or bubbles undergoing shape
AN
changes (oscillation, wobbling) can only be estimated. From the simulation
results however, all relevant information is readily available. Additionally,
M

information about local species transfer at a certain position on the inter-


ED
PT
CE
AC

(a) Velocity field. (b) Pressure field.

Figure 7: Velocity and pressure field for case Eo = 0.3, Mo = 10−11 at t = 0.1379 s

27
ACCEPTED MANUSCRIPT

T
IP
CR
(a) Concentration field.
US (b) Logarithmic concentration
field.
AN
Figure 8: Concentration field for case Eo = 0.3, Mo = 1e-11, Sc = 10 at t = 0.1379 s

face can be obtained, which is impossible in experimental studies. Moreover,


M

commonly only the global mass transfer (kl A) is measured, which lumps the
effects of changes in local concentration gradients and increase in bubble area
ED

and thus contains less information about the actual physics. Due to the lack
of detailed information, experimental results are only partially suitable for
PT

the deduction of mass transfer closure models for detail-reduced simulation


methods (e.g. two-fluid models).
CE

An approach often used in literature for the quantification of mass transfer


is to apply similarity theory which leads to the dimensionless local Sherwood
AC

(∇c·nΣ ) db
number (Shloc ), defined as Shloc := cΣ −c∞
from which the global Sherwood
R
number can be defined as Sh := 1

Shloc dS ≡ kl db
D
. An important differ-

ence between this work and the vast majority of literature is, that we define
the Sherwood number based on the actual bubble area rather than the area

28
ACCEPTED MANUSCRIPT

of a volume-equivalent sphere, thus enabling to study the effects of change


in local concentration gradient and change in interface area separately.
A multitude of different correlations for the mass transfer coefficient or
Sherwood number are available in literature. A comprehensive list of rele-

T
vant correlations is given in (Green and Perry, 2007, pp. 5-71 to 5-73) or in

IP
Colombet et al. (2014). Further relevant correlations can be found in Lochiel

CR
and Calderbank (1964); Oellrich et al. (1973); Clift et al. (1978); Takemura
and Yabe (1998). These correlations are summarized in Table 4.

US
AN
M
ED
PT
CE
AC

29
ACCEPTED MANUSCRIPT

T
IP
CR
Table 4: Sherwood correlations for mass transfer from single rising bubbles

Comments
Correlation

Sh = 1.0(Re Sc)1/3
US
E = Empirical, S = Semiempir-
ical, T = Theoretical

[T] solid sphere,


contaminated spherical
Ref.

McCabe et al. (2005); Sherwood et al.


(1975); Green and Perry (2007)
AN
bubble

Sh = 1.13(Re Sc)1/2 [T] small (spherical) bubbles Sherwood et al. (1975); Green and Perry (2007)
 
db
30

Sh = 1.13(Re Sc)1/2 [S] medium to large bubble, Johnson et al. (1969); Sherwood et al.
0.45 cm + 0.2db
M

carbon dioxide & butene in (1975); Green and Perry (2007)


water, 0.6 cm ≤ db ≤ 4 cm

Sh = 2 + 0.31 Gr1/3 Sc1/3 [S] Calderbank and Moo- Calderbank and Moo-Young (1961);
ED

d3 |ρG − ρL |g Young correlation, carbon Geankoplis (1993) Kirwan (1987); Trey-


with Ra = Gr Sc = b
µL DL
dioxide in water/glycerol, bal (1980); Shah et al. (1982); Green and
db < 2.5 mm Perry (2007)
PT

Sh = 0.42 Gr1/3 Sc1/2 [S] Calderbank and Moo- Calderbank and Moo-Young (1961);
Young correlation, carbon Geankoplis (1993); Green and Perry
dioxide in water/glycerol, (2007); Kirwan (1987); Lee (1992); Shah
CE

db > 2.5 mm et al. (1982)


AC
ACCEPTED MANUSCRIPT

T
IP
CR
!0.116
db g 1/3
Sh = 2 + 0.061 Sc0.546 Re0.779
s 2/3
[E] Hughmark correlation, Hughmark (1967); Treybal (1980);
DL
Res with slip velocity, Green and Perry (2007)
air/water-glycerol system

Sh = 2 + 0.651

Sh = 2 +
(Re Sc)1.72

1 + (Re Sc)1.22

0.232(Re Sc)1.72
US
[T]

[T]
Oellrich correlation, small
bubbles, Re → 0, Sc → ∞

Oellrich correlation, large


Oellrich et al. (1973); Clift et al. (1978);
Fleckenstein and Bothe (2015)

Oellrich et al. (1973); Clift et al. (1978);


AN
1 + 0.205(Re Sc)1.22
bubbles, Re → ∞, Sc → 0 Fleckenstein and Bothe (2015)

Sh = 1.13(Re Sc)1/2 f (χ) [T] aspect ratio χ, Lochiel and Calderbank (1964); Colom-
for oblate spheroids bet et al. (2014)
31

2 1/2 2χ1/3 (χ2 −1)1/2


M

with f (χ) = 3 (1 + k) χ(χ2 −1)1/2 +ln (χ+(χ2 −1)1/2 )


eχ2 − χ sin−1 e
k=− , e = (1 − χ−1 )1/2
e − χ sin−1 e

Sh = 1.13(Re Sc)1/2 f (χ) [S] aspect ratio χ ∈ [1, 3], Figueroa-Espinoza and Legendre (2010);
ED

Colombet et al. (2014)


χ
with f (χ) = 0.524 + 0.88χ − 0.49χ2 + 0.086χ3 500 ≤ ( )1/3 Re ≤ 1000,
8
Sc > 100
" #1/2
PT

2 1
Sh = 1.13 1 − (2.5 + Pe1/2 ) [E] nearly spherical bubbles, Takemura and Yabe
3 (1 + 0.09 Re2/3 )3/4
0 < Re ≤ 100, Sc  1 (1998); Colombet
et al. (2014)
CE
AC
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
32

M
ED
PT
CE
AC
ACCEPTED MANUSCRIPT

It can be seen from Table 4, that many different forms of correlations


exist in literature and that each correlation has its respective narrow range
of validity. It should also be noted, that especially empirical correlations can
produce entirely wrong results when applied outside their range of validity.

T
While the number of independent dimensionless numbers describing mass

IP
transfer from rising bubbles is fixed, almost all reported correlations include

CR
only some dependencies and are thus strongly limited in their applicability.
Only very few correlations account for the actual shape of the rising bubble,

US
while the vast majority of mass transfer correlations only account for the ratio
of hydrodynamic to diffusion boundary layer thickness (Schmidt number) and
ratio of inertia/buoyancy to viscous forces (Reynolds or Grashof number).
AN
The most common form of a Sherwood correlation is

Sh = 2 + a Reb Scc or Sh = a Reb Scc . (27)


M

It models the mass transfer in dependence of the Schmidt and Reynolds


ED

numbers only, disregarding the effects of bubble shape.


To obtain a full description of the species transfer, additional influence
PT

parameters are to be considered. In order to obtain all relevant influence


parameters for a species transfer correlation, a dimensional analysis has been
CE

performed. Considering the simplified species transport Equation (5), re-


called here for the reader’s convenience,
AC

∂t c + ∇· (cu) = ∇· (D∇c) ,

the non-dimensional form then reads

urel trel Drel trel


∂t∗ c∗ + ∇∗ · (c∗ u∗ ) = ∇∗ · (D∗ ∇∗ c∗ ) . (28)
xrel x2rel

33
ACCEPTED MANUSCRIPT

Defining urel trel = xrel then leads to the non-dimensional species transport
equation with one non-dimensional parameter

Drel νrel
∂t∗ c∗ + ∇∗ · (c∗ u∗ ) = ∇∗ · (D∗ ∇∗ c∗ ) (29)

T
νrel xrel urel
1
= ∇∗ · (D∗ ∇∗ c∗ ) .

IP
Re Sc

Thus, the non-dimensional species concentration transport is only dependent

CR
on the Peclet number (Pe = Re Sc). From this result, the Sherwood corre-
lation (which is a non-dimensionalised form of the concentration gradient at

US
the interface) is assumed to be a function of the Peclet number only.
However, the dimensionless velocity profile and the bubble shape is a func-
AN
tion of the Eötvös and Reynolds numbers. Thus, the Sherwood correlation
is assumed to take the following form:
M

Sh = f (Re, Eo) · g(Pe) ,

where f (Re, Eo) denotes a correction factor describing the influence of changes
ED

in the local velocity field due to bubble deformation. This form of correlation
can also be found in literature, e.g. in Lochiel and Calderbank (1964), who
PT

introduce an aspect ratio to include the bubble shape influence onto the ve-
locity field. From the above reasoning, it can be deduced that the influence
CE

of the Schmidt number for a given set of material parameters (and thus fixed
velocity field) is approximately constant. From the solution of a potential
AC

flow field around a sphere, the Sherwood number is expected to be a function


of Sca with a close to 0.5.

34
ACCEPTED MANUSCRIPT

4.2.2. Numerical results


In order to improve existing species transfer correlations, the results
of the conducted parameter study are utilised to obtain insights into the

T
evolution of the dynamically changing bubble surface area and its depen-
dency on the global Sherwood number. The main results are visualised

IP
in Figures 9a and 9b. Figure 9a shows that the interfacial area increases

CR
1.7
Eo = 0.3
102
Sc = 0.3
Eo = 1
1.6 Sc = 1
Eo = 2

US
Sc = 3
Eo = 5
Sc = 10
1.5 Eo = 10
Sc = 30
Eo = 20 101.5 Sc = 100
correlation
1.4 a · sqrt(Pe)
Sr

Sh
AN
1.3
101
1.2

1.1
M

1 0 100.5
10 101 102 101 102 103 104
Re Pe

(a) Bubble area closure. (b) Sherwood number closure.


ED

Figure 9: Surface ratio and Sherwood number as function of Reynolds and Peclet number
PT

significantly with increasing Reynolds and Eötvös numbers. The correla-


tion curves plotted in 9a take the form Sr = 1 + 0.0011 Eo Reg(Eo) , with
CE

g(Eo) ≈ 0.726 + 0.317 1+(1.6081 Eo)1.426 ∈ [0.726, 1]. Here, the influence of the
Eötvös number is dominant, which is not surprising. The well-known Grace
AC

diagram (Clift et al., 1978) for instance visualises – although not quantifies –
this effect. Figures 9b and 11a on the other hand show that the global Sher-
wood number decreases with increasing bubble deformation (i.e. increasing
Eötvös number). However, Figure 9b also shows that the decrease in global

35
ACCEPTED MANUSCRIPT

102
Eo = 0.3 , Mo = 3 · 10−7
Eo = 0.3 , Mo = 10−8
Eo = 0.3 , Mo = 10−10
Eo = 0.3 , Mo = 10−11
Eo = 1 , Mo = 10−5
101.5
Eo = 1 , Mo = 10−7
Eo = 1 , Mo = 3 · 10−9

T
Sh

Eo = 2 , Mo = 10−4
Eo = 2 , Mo = 10−6
Eo = 2 , Mo = 5 · 10−9

IP
1
10
Eo = 5 , Mo = 10−3
Eo = 5 , Mo = 10−5
Eo = 10 , Mo = 10−2
Eo = 10 , Mo = 3 · 10−5

CR
Eo = 20 , Mo = 3 · 10−2
100.5
10−2 10−1 100 101 102
Sc

Figure 10: Dependency of global Sherwood number from Schmidt number

US
Sherwood number due to bubble deformation is strictly limited by a lower
AN
bound, while the upper bound can be approximately by Higbie’s penetration

theory (Higbie, 1935), i.e. by Shmax = √2π Pe. Our simulations show that the
upper limit is only surpassed for (nearly) spherical path-instable bubbles (cf.
M

Figure 11a). We further investigated the dependency of the global Sherwood


ED

102 0.6

0.55
PT

101.5
Exponent
Sh

0.5
Eo = 0.3 , Mo = 10−10
Eo = 0.3 , Mo = 10−11
CE

101 Eo = 1 , Mo = 3 · 10−9
Eo = 0.3
Eo = 2 , Mo = 5 · 10−9 0.45 Eo = 1
−5
Eo = 5 , Mo = 10 Eo = 2
Eo = 10 , Mo = 3 · 10−5 Eo = 5
penetration model (Higbie) Eo = 10

100.5 0.4
AC

0 100 200 300 400 500 600 700


101 10 2
103 104
Pe Re

(a) Bubble shape dependency. (b) Schmidt number exponent.

Figure 11: Influence of bubble shape onto global Sherwood number

36
ACCEPTED MANUSCRIPT

number on the Schmidt number. It was found from our simulation results
that this functional relation follows in good approximation an exponential
law with a constant exponent and a material system-dependent pre-factor.
This means that the global Sherwood number for any given material system

T
is seen to solely depend on the Schmidt number by

IP
Sh = a · Scb , (30)

CR
where a = g (Eo, Re) and b have been found to be almost constant. This then
results with a set of curves, where all our simulation data points lie on straight

US
lines in the double-logarithmic plot as shown in Figure 10. Fitting b for
different cases yields values of approximately 0.5 (cf. Figure 11b) and thus, as
AN
expected, varies only slightly from the theoretical value of 0.5, obtained, e.g.,
in the analytical solution of potential flow around a sphere. However, Figure
M

11b shows that the exponent does indeed slightly increase with increasing
Reynolds and Eötvös number (i.e. increasing bubble deformation). The
ED

functional relation of the pre-factor a can be described in good agreement by

a = γ(Sr ) Re0.5 , (31)


PT

hence depending on the surface ratio and Reynolds number. This form is
also suggested by Lochiel and Calderbank (1964) based on analytical in-
CE

vestigations of species transfer from ellipsoidal bubbles. Unlike in Lochiel


and Calderbank (1964), the reduction in Sherwood number due to bubble
AC

deformation is found in our simulation results to have strict a lower limit


and reduction factors obtained by the proposed correlation of Lochiel and
Calderbank (1964) are far too high, especially for Eo > 5 and moderate to
high Reynolds numbers. A further finding is that in addition to the Peclet

37
ACCEPTED MANUSCRIPT

number, the Schmidt number is found to also affect overall mass transfer,
although this effect is small (cf. Figure 11b).
Eo = 0.3, Sc = 1
102
Eo = 1, Sc = 1

T
Eo = 2, Sc = 1
Eo = 5, Sc = 1
Eo = 10, Sc = 1

IP
Eo = 0.3, Sc = 10
101.5 Eo = 1, Sc = 10
Eo = 2, Sc = 10
Sh ·Sr

Eo = 5, Sc = 10

CR
Eo = 10, Sc = 10
Eo = 0.3, Sc = 100
Eo = 1, Sc = 100
101
Eo = 2, Sc = 100
Eo = 5, Sc = 100
Eo = 10, Sc = 100

100.5
101 102
Pe US
103 104

Figure 12: Non-dimensional kl A-coefficient as a function of the Peclet number


AN

Since nearly all Sherwood correlations reported in literature are based on


M

the area of volume-equivalent spheres, they are a non-dimensional form of


kl A and thus comparison with numerical results has to account for this fact.
ED

Figure 12 shows our numerical results and a comparison to the correlation of


Oellrich et al. (1973) is given in Figure 13. It can be seen, that our numerical
results agree well with the correlations proposed in Oellrich et al. (1973), in
PT

both upper and lower bound, as well as the dependency of the Sherwood
number on the Schmidt number.
CE

Lastly, our numerical results are compared to the correlation proposed in


Takemura and Yabe (1998) in Figure 14. Here, to better distinguish beteen
AC

the different curves, the numerical results are plotted separately for each
Eötvös number against the correlation of Takemura and Yabe (1998). As
the correlation is only valid for large Schmidt numbers, deviation from the
numerical results at small Schmidt numbers can be ignored. However, espe-

38
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Figure 13: Comparison of simulation results with correlation of Oellrich et al. (1973)
PT

cially for very small and very large Reynolds numbers, a deviation from the
proposed correlation can be seen. The deviation at large Reynolds numbers
CE

(cf. Figure 14a, Mo = 10−11 ) is to be expected as the correlation of Take-


mura and Yabe (1998) is derived from experimental data for Re < 100. The
source of deviation for small Reynolds numbers (cf. Figure 14c, Mo = 10−4 )
AC

remains unclear. Overall, it can be seen that the numerical results agree well
with the correlation.

39
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Figure 14: Comparison of simulation results with correlation of Takemura and Yabe (1998)
PT

5. Summary and Conclusion

A modelling framework for the direct numerical simulation of dilute species


CE

transfer from single bubbles rising in quiescent liquids is presented, employing


the algebraic Volume-of-Fluid and the Continous Species Transfer methods.
AC

The utilized flow solver is based on OpenFOAM interFoam which however


has been significantly enhanced to meet the specific requirements. The sim-
ulations are conducted in a moving reference frame, where only the phase
fraction advection is solved in this frame and the linear momentum is solved

40
ACCEPTED MANUSCRIPT

in the inertial frame of a fixed observer. This setup is similar to the moving
window technique employed, e.g., in Bothe and Fleckenstein (2013).
In order to resolve the thin concentration boundary layer, state-of-the-art
high performance computing techniques are employed. The computational

T
mesh is dynamically refined at the interface and in the bubble wake, based

IP
on a multi-criterion refinement strategy. To ensure computational efficiency,

CR
dynamic load balancing is performed based on a clustered hierarchical de-
composition.

US
A parameter study at single rising bubbles has been conducted in order
to investigate the influence of the local flow field and the Schmidt number
onto the species transfer process. It is found that the bubble shape strongly
AN
influences the transfer. Stronger deformation leads to a significant decrease
in the global Sherwood number, while at the same time leading to an increase
M

of interfacial area. Further, it is deduced that the influence of the Schmidt


number Sc onto the species transfer is well described by a power law with
ED

constant exponent. Thus, the global Sherwood number is always a function


of Sca , with a being a function of Eötvös Eo and Morton Mo numbers, respec-
PT

tively. Further it is shown that a is always close to 0.5 but slightly increasing
with bubble deformation and rise velocity.
CE

One major aspect of this contribution is the identification of the neces-


sity to separately investigate the two different causes for changes in global
mass/species transfer: changes in local concentration gradients and changes
AC

in interfacial area. The numerical results presented in this contribution high-


light the importance of accounting for bubble shape in mass transfer corre-
lations. As reported in Lochiel and Calderbank (1964), bubble deformation

41
ACCEPTED MANUSCRIPT

is shown to lead to a substantial reduction in the Sherwood number but at


the same time lead to an increase of interfacial area. Unlike reported in their
theoretical study, the reduction factor of Sherwood number based on bubble
deformation is found to be strictly limited and much smaller, especially for

T
higher Eötvös numbers (Eo > 5) and moderate to high Reynolds numbers.

IP
Further, the presented numerical results indicate that the correlations pro-

CR
posed by Oellrich et al. (1973) and Takemura and Yabe (1998) are a good
estimate in a large parameter range. The numerical findings of this work

US
can be utilized to improve the accuracy of detail-reduced simulation meth-
ods. An example is the two-fluid model, which requires an accurate species
transfer correlation over a wide range of bubble shapes and local Reynolds
AN
numbers and also an accurate description of the local interfacial area, which
should also account for changes due to interface deformation.
M

References
ED

B. Aboulhasanzadeh, S. Thomas, M. Taeibi-Rahni, and G. Tryggvason. Mul-


tiscale computations of mass transfer from buoyant bubbles. Chem. Eng.
PT

Sci., 75:456–467, 2012.

A. Alke and D. Bothe. 3D Numerical Modeling of Soluble Surfactant at


CE

Fluidic Interfaces Based on the Volume-of-Fluid Method. Fluid Dynamics


and Materials Processing, 5(4):345–372, 2009.
AC

S. Batzdorf. Heat transfer and evaporation during single drop impingement


onto a superheated wall . PhD thesis, TTD, TU Darmstadt, 2015.

K. Bäumler. Simulation of single drops with variable interfacial tension.

42
ACCEPTED MANUSCRIPT

PhD thesis, Applied Mathematics III, Friedrich-Alexander Universität


Erlangen-Nürnberg, 2014.

D. Bothe and S. Fleckenstein. A Volume-of-Fluid-based method for mass

T
transfer processes at fluid particles. Chem. Eng. Sci., 101:283–302, 2013.

IP
D. Bothe, M. Koebe, K. Wielage, and H.J. Warnecke. VOF-simulations of

CR
mass transfer from single bubbles and bubble chains rising in aqueous so-
lutions. In 4th ASME-JSME Joint Fluids Engineering Conference, number
FEDSM2003-45155, July 6-11 2003.
US
D. Bothe, M. Koebe, K. Wielage, J. Prüss, and H.-J. Warnecke. Direct
AN
numerical simulation of mass transfer between rising gas bubbles and wa-
ter. In Martin Sommerfeld, editor, Bubbly Flows: Analysis, Modelling
and Calculation, pages 159–174, Berlin, Heidelberg, 2004. Springer Berlin
M

Heidelberg.
ED

D. Bothe, M. Kröger, and H.-J. Warnecke. A VOF-based conservative


method ffor the simulation of reactive mass transfer from rising bubbles.
Fluid Dynamics and Materials Processing, 7(3):303–316, 2011.
PT

J.U. Brackbill, D.B. Kothe, and C. Zemach. A continuum method for mod-
CE

eling surface tension. J. Comput. Phys., 100:335–354, 1992.

P.H. Calderbank and M.B. Moo-Young. The continuous phase heat and mass
AC

transfer properties of dispersions. Chem. Eng. Sci., 16:39–54, 1961.

F. Cariglino. External aerodynamic simulations in a rotating frame of refer-


ence. PhD thesis, Politechnico di Torino, 2013.

43
ACCEPTED MANUSCRIPT

R. Clift, J.R. Grace, and M.E. Weber. Bubbles, Drops, and Particles. New
York ; London : Academic Press, 1978.

D. Colombet, D. Legendre, F. Risso, A. Cockx, and P. Guiraud. Dynamics

T
and mass transfer of rising bubbles in a homogenous swarm at large gas

IP
volume fraction. J. Fluid Mech., 763:254–285, 2014.

CR
D. Darmana, N.G. Deen, and J.A.M. Kuipers. Detailed 3D modeling of mass
transfer processes in two-phase flows with dynamic interfaces. Chem. Eng.
Technol., 29(9):1027–1033, 2006.

US
D. Darmana, W. Dijlhuizen, N.G. Deen, M. van Sint Annaland, and J.A.M.
Kuipers. Detailed 3D Modelling of Mass Transfer Processes in Two-Phase
AN
Flows with Dynamic Interfaces. In 6th International Conference on Mul-
tiphase Flow, ICMF 2007, 2007.
M

M.R. Davidson and M. Rudman. Volume-of-Fluid calculation of heat or


mass transfer across deforming interfaces in two-fluid flow. Numer. Heat.
ED

Tr. B-Fund., 41(3-4):291–308, 2002.

W.D. Deckwer. Reaktionstechnik in Blasensäulen. Otto Salle Verlag GmbH


PT

& Co, 1985.


CE

D. Deising, H. Marschall, and D. Bothe. A unified single-field model frame-


work for Volume-Of-Fluid simulations of interfacial species transfer applied
AC

to bubbly flows. Chem. Eng. Sci., 139:173 – 195, 2016.

K.B. Deshpande and W.B. Zimmerman. Simulations of mass transfer limited


reaction in a moving droplet to study transport limited characteristics.
Chem. Eng. Sci., 61:6424–6441, 2006.

44
ACCEPTED MANUSCRIPT

S.S. Deshpande, L. Anumolu, and M.F. Trujillo. Evaluating the performance


of the two-phase flow solver interFoam. Comput. Sci. Discov., 5(1):014016,
2012.

T
D.A. Drew. Mathematical modeling of two-phase flow. Annu. Rev. Fluid

IP
Mech., 15:261–291, 1983.

CR
D.A. Drew and S.L. Passman. Theory of Multicomponent Fluids. Springer-
Verlag New York, 1999.

US
M. Dudukovic. Relevance of multiphase reaction engineering to modern tech-
nological challenges. Ind. and Eng. Chem. Res., 46:8574–8686, 2007.
AN
P.C. Duineveld. The rise velocity and shape of bubbles in pure water at high
Reynolds number. J. Fluid Mech., 292:325–332, 6 1995. ISSN 1469-7645.
M

Hanno Essén. 21st Century Mechanics – Motion in accelerated reference


frames. Technical report, Royal Institute of Technology, Department of
ED

Mechanics, Stockholm, Sweden, 2002.

L.C. Evans and R.F. Gariepy. Measure Theory and Fine Properties of Func-
PT

tions. In Steven G. Krantz, editor, Studies in Advanced Mathematics. CRC


Press, 1992.
CE

B. Figueroa-Espinoza and D. Legendre. Mass or heat transfer from spheriodal


AC

gas bubbles rising through a stationary liquid. Chem. Eng. Sci., 65:6296–
6309, 2010.

S. Fleckenstein and D. Bothe. A Volume-of-Fluid-based numerical method

45
ACCEPTED MANUSCRIPT

for multi-component mass transfer with local volume changes. J. Comput.


Phys., 301:35–58, 2015.

M.M. Francois and N.N. Carlson. The Balanced-Force Volume Tracking Algo-

T
rithm and Global Embedded Interface Formulation for Droplet Dynamics

IP
with Mass Transfer. In ASME 2010 3rd Joint US-European Fluids Engi-
neering Summer Meeting collocated with 8th International Conference on

CR
Nanochannels, Microchannels, and Minichannels, pages 81–88. American
Society of Mechanical Engineers, 2010.

US
M.M. Francois, S.J. Cummins, E.D. Dendy, D.B. Kothe, J.M. Sicilian, and
M.W. Williams. A balanced-force algorithm for continuous and sharp in-
AN
terfacial surface tension models within a volume tracking framework. J.
Comput. Phys., 213:141–173, 2006.
M

C.J. Geankoplis. Transport processes and unit operations. Prentice Hall, 3rd
edition, 1993.
ED

D.W. Green and R.H. Perry. Chemical Engineers’ Handbook. McGraw-Hill,


8th edition, 2007.
PT

Y. Haroun, D. Legendre, and L. Raynal. Volume of Fluid method for interfa-


cial reactive mass transfer: Application to stable liquid film. Chem. Eng.
CE

Sci., 65(10):2896 – 2909, 2010.


AC

K. Hayashi and A. Tomiyama. Interface tracking simulation of mass transfer


from a dissolving bubble. JCMF, 3(4):247–262, 2011.

R. Higbie. The rate of absorption of a pure gas into a still liquid during short
periods of exposure. Trans. Am. Inst. Chem. Engrs., 35:365–389, 1935.

46
ACCEPTED MANUSCRIPT

S. Hill, D. Deising, T. Acher, H. Klein, D. Bothe, and H. Marschall.


Boundedness-preserving implicit correction of mesh-induced errors for VoF
based heat and mass transfer. ArXiv e-prints, 2017.

T
G.A. Hughmark. Holdup and mass transfer in bubble columns. Ind. Eng.

IP
Chem. Process Des. Dev., 6:218–220, 1967.

CR
H.A. Jakobsen. Chemical Reactor Modeling: Multiphase Reactive Flows.
Springer-Verlag Berlin Heidelberg, 2008.

US
A.I. Johnson, F. Besik, and A.E. Hamielec. Mass transfer from a single rising
bubble. Can. J. Chem. Eng., 47:559–564, 1969.
AN
J.G. Khinast. Impact of 2-D bubble dynamics on the selectivity of fast gas-
liquid reactions. AIChE J., 47:2304–2319, 2001.
M

J.G. Khinast, A. Koynov, and T.M. Leib. Reactive mass transfer at gas-liquid
interfaces: impact of micro-scale fluid dynamics on yield and selectivity of
ED

liquid-phase cyclohexane oxidation. Chem. Eng. Sci., 58:3961–3971, 2003.

D.J. Kirwan. Handbook of separation process technology, chapter Mass trans-


PT

fer principles, pages 60–128. John Wiley & Sons, 1987.


CE

A. Koynov, J.G. Khinast, and G. Tryggvason. Mass transfer and chemical


reactions in bubble swarms with dynamic interfaces. AIChE J., 51(10):
AC

2786–2800, 2005.

J.M. Lee. Biochemical engineering. Prentice Hall, 1992.

47
ACCEPTED MANUSCRIPT

C. Lehrenfeld. On a space-time extended Finite Element Method for the


solution of a class of two-phase mass transport problems. PhD thesis,
IGPM, RWTH Aachen, 2015.

T
A.C. Lochiel and P.H. Calderbank. Mass transfer in the continuous phase

IP
around axisymmetric bodies of revolution. Chem. Eng. Sci., 19(7):471 –
484, 1964. ISSN 0009-2509.

CR
T. Maric, H. Marschall, and D. Bothe. voFoam - A geometrical Volume
of Fluid algorithm on arbitrary unstructured meshes with local dynamic
US
adaptive mesh refinement using OpenFOAM. ArXiv e-prints, May 2013.
AN
H. Marschall, K. Hinterberger, C. Schüler, F. Habla, and O. Hinrichsen. Nu-
merical simulation of species transfer across fluid interfaces in free-surface
flows using OpenFOAM. Chem. Eng. Sci., 78(0):111 – 127, 2012.
M

W.L. McCabe, J.C. Smith, and P. Harriott. Unit operations of chemical


ED

engineering. McGraw-Hill, 7th edition, 2005.

G. Mougin and J. Magnaudet. Path Instability of a Rising Bubble. Phys.


PT

Rev. Lett., 2002.

W. Muschik and L. Restuccia. Changing the observer and moving materi-


CE

als in continuum physics: objectivity and frame-indifference. Technische


Mechanik, 22(2):152–160, 2001.
AC

H. Oellrich, H. Schmidt-Traub, and H. Brauer. Theoretische Berechnung des


Stofftransports in der Umgebung einer Einzelblase. Chem. Eng. Sci., 28:
711–721, 1973.

48
ACCEPTED MANUSCRIPT

A. Onea, M. Wörner, and D.G. Cacuci. A qualitative computational study of


mass transfer in upward bubble train flow through square and rectangular
mini-channels. Chem. Eng. Sci., 64:1416–1435, 2009.

T
C.S. Peskin. Numerical analysis of blood flow in the heart. J. Comput. Phys.,

IP
25:220–252, 1977.

CR
S. Popinet. An accurate adaptive solver for surface-tension-driven interfacial
flows. J. Comput. Phys., 228:5838–5866, 2009.

US
S. Radl, A. Koynov, G. Tryggvason., and J.G. Khinast. DNS-based pre-
diction of the selectivity of fast multiphase reactions: Hydrogenation of
AN
nitroarenes. Chem. Eng. Sci., 63:3279–3291, 2008.

H. Rusche. Computational Fluid Dynamics of Dispersed Two-Phase Flows at


M

High Phase Fractions. PhD thesis, Department of Mechanical Engineering,


Imperial College of Science, Technology & Medicine, Exhibition Road,
ED

London SW7 2BX, December 2002.

Y.T. Shah, B.G. Kelkar, S.P. Goobole, and W.-D. Deckwer. Design parame-
PT

ters estimations for bubble column reactors. AIChE J., 28:353–379, 1982.

T.K. Sherwood, R.L. Pigford, and C.R. Wilke. Mass transfer. McGraw-Hill,
CE

1975.
AC

F. Takemura and A. Yabe. Gas dissolution process of spherical rising gas


bubbles. Chem. Eng. Sci., 53(15):2691 – 2699, 1998.

R.E. Treybal. Mass-transfer operations. McGraw-Hill, 3rd edition, 1980.

49
ACCEPTED MANUSCRIPT

Z. Tukovic and H. Jasak. Simulation of Free-Rising Bubble with Soluble Sur-


factant Using Moving Mesh Finite Volume/Area Method. In Proceedings
of 6th International Conference on CFD in Oil & Gas, Metallurgical and
Process Industries, no. CFD08-072, 2008.

T
IP
O. Ubbink. Numerical Prediction of Two Fluid Systems with Sharp Inter-
faces. PhD thesis, Department of Mechanical Engineering, Imperial College

CR
of Science, Technology & Medicine, January 1997.

P. Weber, H. Marschall, and D. Bothe. Highly accurate two-phase species

US
transfer based on ALE Interface Tracking. Int. J. Heat Mass Transfer,
104:759–773, 2017.
AN
P.S. Weber. Modeling and Numerical Simulation of Multi-Component Two-
Phase Fluid Systems with Ionic Species. PhD thesis, TU Darmstadt, 2016.
M

PhD thesis.

A. Weiner and D. Bothe. Advanced subgrid-scale modeling for convection-


ED

dominated species transport at fluid interfaces with application to mass


transfer from rising bubbles. J. Comput. Phys., 347:261–289, 2017.
PT

H.G. Weller. A new approach to VOF-based interface capturing methods


for incompressible and compressible flow. Technical report, OpenCFD
CE

Limited, 2006.
AC

C. Yang and Z.-S. Mao. Numerical simulation of interphase mass transfer


with the level set approach. Chem. Eng. Sci., 60(10):2643–2660, 2005.

S.T. Zalesak. Fully multidimensional flux-corrected transport algorithms for


fluids. J. Comput. Phys., 31(3):335 – 362, 1979.

50

Das könnte Ihnen auch gefallen