Вы находитесь на странице: 1из 217

Experimental and numerical investigation of bubble

column reactors
Bai, W.

DOI:
10.6100/IR693280

Published: 01/01/2010

Document Version
Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the author's version of the article upon submission and before peer-review. There can be important differences
between the submitted version and the official published version of record. People interested in the research are advised to contact the
author for the final version of the publication, or visit the DOI to the publisher's website.
• The final author version and the galley proof are versions of the publication after peer review.
• The final published version features the final layout of the paper including the volume, issue and page numbers.
Link to publication

Citation for published version (APA):


Bai, W. (2010). Experimental and numerical investigation of bubble column reactors Eindhoven: Technische
Universiteit Eindhoven DOI: 10.6100/IR693280

General rights
Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners
and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
• You may not further distribute the material or use it for any profit-making activity or commercial gain
• You may freely distribute the URL identifying the publication in the public portal ?
Take down policy
If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately
and investigate your claim.

Download date: 31. Oct. 2017


Experimental and Numerical Investigation of
Bubble Column Reactors
Samenstelling promotiecommissie:

prof.dr. P.J. Lemstra, chairman, Eindhoven University of Technology


prof.dr.ir. J.A.M. Kuipers, promotor Eindhoven University of Technology
dr.ir. N.G. Deen, copromotor Eindhoven University of Technology
prof.dr. R.O. Fox Iowa State University, USA
prof.dr.ir. B.J. Geurts Eindhoven University of Technology
prof.dr. D. Lohse University of Twente
prof.dr. R.F. Mudde Delft University of Technology
prof.dr. A.E.P. Veldman University of Groningen
Experimental and Numerical Investigation of
Bubble Column Reactors

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de


Technische Universiteit Eindhoven, op gezag van de
rector magnificus, prof.dr.ir. C.J. van Duijn, voor een
commissie aangewezen door het College voor
Promoties in het openbaar te verdedigen
op donderdag 23 december 2010 om 16.00 uur

door

Wei Bai

geboren te Shaanxi, China


Dit proefschrift is goedgekeurd door de promotor:

prof.dr.ir. J.A.M. Kuipers

Copromotor:
dr.ir. N.G. Deen

Copyright © 2010 by Wei Bai


All rights reserved. This book, or parts thereof, may not be reproduced in any form
or by any means, electronic or mechanical, including photocopying, recording or any
information storage and retrieval system now known or to be invented, without
written permission from the author.

A catalogue record is available from the Eindhoven University of


Technology Library.

ISBN: 978-90-386-2405-1
Cover design by Jie Fan.
Printed by Ipskamp Drukkers B.V., Enschede, the Netherlands.
For my parents
Contents

Contents vii

1 Introduction 1
1.1 Bubble column reactors . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Experimental techniques . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Computational Fluid Dynamics . . . . . . . . . . . . . . . . . . . 8
1.4 Objectives & outline . . . . . . . . . . . . . . . . . . . . . . . . . 10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Single bubble experiment 19


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Data processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 28
2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3 Bubble properties of heterogeneous bubbly flows in bubble column 41


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Data processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 50
3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4 Numerical analysis of the effect of gas sparging on bubble column


hydrodynamics 63
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

vii
Contents

4.2 Discrete bubble model . . . . . . . . . . . . . . . . . . . . . . . . 65


4.3 Simulation details . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 73
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

5 Discrete bubble modeling of bubbly flows: Swarm effects 89


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.2 Discrete bubble model . . . . . . . . . . . . . . . . . . . . . . . . 92
5.3 Drag coefficient correlations . . . . . . . . . . . . . . . . . . . . . 96
5.4 Simulation details . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.5 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 100
5.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

6 Discrete bubble modeling of bubbly flows: Implementation of


breakup models 117
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
6.2 Discrete bubble model . . . . . . . . . . . . . . . . . . . . . . . . 119
6.3 Coalescence model . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.4 Breakup models & implementation . . . . . . . . . . . . . . . . . 124
6.5 Simulation details . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
6.6 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 131
6.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

7 Numerical investigation of gas holdup and phase mixing in bubble


column reactors 149
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
7.2 Discrete bubble model . . . . . . . . . . . . . . . . . . . . . . . . 151
7.3 Correlations of the gas holdup . . . . . . . . . . . . . . . . . . . . 155
7.4 Phase mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
7.5 Numerical aspects of phase mixing study . . . . . . . . . . . . . 163
7.6 Simulation details . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

viii
Contents

7.7 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 166


7.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
7.A Runge-Kutta-Fehlberg method . . . . . . . . . . . . . . . . . . . 178
7.B Interpolation methods . . . . . . . . . . . . . . . . . . . . . . . . 179
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

Summary 187

Samenvatting 191

总结 195

List of publications 197

Acknowledgement 199

About the author 203

ix
Chapter
1
Introduction
Bubble column reactors are commonly used in chemical, petrochemical, biochemical,
pharmaceutical, metallurgical industries and so on for a variety of processes, i.e.
hydrogenation, oxidation, chlorination, alkylation, chemical gas cleaning, various
bio-technological applications, etc.
Bubble column reactors have advantages of ease of operation, low operating and
maintenance costs as it requires no moving parts, and compactness. Also, they have
the characteristics of high catalyst durability and excellent heat and mass transfer
characteristics. Furthermore, bubble column reactors can be adapted to specific con-
figurations according to practical requirements.
In spite of the simple geometry of bubble column reactors, complex hydrodynamics
and its influence on transport characteristics make it difficult to achieve reliable design
and scale-up of bubble column reactors. Research on bubble column reactors covers
a wide range of activities, i.e. gas holdup, bubble properties, interfacial area, flow
regime, heat and mass transfer, back mixing, pressure drop, etc. Investigations on the
characteristics of bubble column reactors include both experimental and numerical
work. Some experimental techniques and numerical methods used in multiphase
flows are briefly introduced in this chapter. Finally, the objectives and outline of this
thesis will be described.

1
1. Introduction

1.1 Bubble column reactors

In general, bubble columns are used to achieve intimate contact between a


continuous liquid and dispersed gas phase. In its most simple form, a bubble
column reactor consists of a vertical cylindrical vessel filled with a liquid and
a gas distributor at the bottom, as sketched in Figure 1.1. Gas is sparged
through the distributor into the column in the form of bubbles and comes in
contact with the liquid.

Figure 1.1: Sketch of a simple bubble column reactor.

Bubble columns are employed in many industries, such as chemical,


petrochemical, biochemical, pharmaceutical, metallurgical industries and so
on. Besides a number of conventional processes, such as hydrogenation,
oxidation, chlorination, alkylation, chemical gas cleaning and various bio-
technological applications, bubble column reactors are also employed in the
processes of partial oxidation of ethylene to acetaldehyde, oxidation of ac-
etaldehyde to acetic acid, oxidation of p-xylene to dimethylterephthalate,

2
1.1. Bubble column reactors

Fischer-Tropsch synthesis, synthesis of methanol, synthesis of hydrocarbons,


hydrolysis of phosgene, oxychlorination of ethylene to 1,2-dichlooroethane,
ozonization of waste water and biological waste water treatment. In industrial
applications, bubble column reactors usually have very large dimensions. For
instance, bubble column reactors for high-tonnage products have capacities
of 100 − 300 m3 . There are even larger bubble column reactors with capacities
up to 3000 m3 which are employed as fermenters for protein production from
methanol. For waste water treatment, the units can be as large as 20 000 m3 .
Bubble columns offer certain distinct advantages such as ease of operation,
low operating and maintenance costs (as they don’t contain moving parts) and
compactness. Also, high catalyst durability can be achieved and excellent
heat and mass transfer characteristics prevail. Moreover, bubble columns
can easily be adapted. For instance, the liquid phase can be operated in
batch mode, or in co-current or counter-current flow to the gas phase. When
solid particles are suspended in the liquid, a slurry phase is formed and
the bubble column is then referred to as a slurry bubble column reactor. In
addition, according to particular practical needs, bubble column reactors can
be modified to different forms, i.e. a cascade bubble column, a packed bubble
column and a multishaft bubble column, as shown in Figures 1.2(a)–1.2(c).
A simple bubble column reactor incorporating additional perforated plates
forms a cascade bubble column and gas is redistributed over the perforated
plates. Consequently, the redistribution intensifies mass transfer and reduces
the fraction of large bubbles and prevents back mixing in both phases. By
using a packing or static mixers (Figure 1.2(d)), the back mixing can be further
reduced. Multishaft bubble column reactors prevent bulk circulation and
uniform distribution of gas bubbles over the cross section can be achieved.
Furthermore, the circulation can be enhanced via either an internal or external
loop, as shown in Figures 1.2(e)–1.2(f).
In spite of the simple geometry of bubble column reactors, complex hy-
drodynamics and its influence on transport characteristics make it difficult
to achieve reliable design and scale-up of bubble column reactors. There
are many factors, such as column dimensions, column internals design, gas
distributor design, operating conditions, i.e. pressure and temperature, su-

3
1. Introduction

(a) Cascade bubble column (b) Packed bubble column (c) Multishaft bubble col-
umn

(d) Bubble column with (e) Bubble column with in- (f) Bubble column with ex-
static mixers ternal loop ternal loop

Figure 1.2: Types of bubble column reactors.

perficial gas velocity, physical properties, solid particle properties and concen-
tration and so on, which influence the performance of this type of gas-liquid
contactors significantly. During the past decades, scientific interest in bubble
column reactors has increased considerably (Deckwer, 1992; Kantarci et al.,
2005). The research on the bubble columns covers a wide range including
gas holdup (i.e. Fair et al., 1962; Shah et al., 1982; Heijnen and Van’t Riet,
1984; Kawase and Moo-Young, 1987; Krishna et al., 1991; Saxena and Rao,
1991; Ruzicka et al., 2001), bubble characteristics (i.e. Abuaf et al., 1978; Lin
and Fan, 1999; Manera et al., 2009; Guet et al., 2003; Luther et al., 2004), in-
terfacial area (i.e. Kataoka et al., 1986; Tan and Ishii, 1990; Revankar and
Ishii, 1993; Delhaye and Bricard, 1994; Kiambi et al., 2001; Manera et al., 2009),

4
1.2. Experimental techniques

flow regime (i.e. Shah et al., 1982; Shaikh and Al-Dahhan, 2007), heat and
mass transfer coefficients (i.e. Deckwer et al., 1974; Wang and Fan, 1978; Shah
et al., 1982; Heijnen and Van’t Riet, 1984; Zehner, 1986; Verma, 1989; Saxena
and Rao, 1991; Avdeev et al., 1992; Lin and Fan, 1999; Merchuk et al., 1994;
Schlüter et al., 1995; Dudley, 1995), back mixing (i.e. Ohki and Inoue, 1970;
Deckwer et al., 1974; Hikita and Kikukawa, 1974; Joshi, 1980, 1982; Heijnen
and Van’t Riet, 1984; Kawase and Moo-Young, 1986; Zehner, 1986; Westerterp
et al., 1987; Wachi and Nojima, 1990; Majumder, 2008), and pressure drop
(i.e. Carleton et al., 1967; Gharat and Joshi, 1992; Molga and Westerterp, 1997;
Majumder et al., 2006).

1.2 Experimental techniques

In order to study the characteristics of bubble column reactors, i.e. those


introduced above, a variety of experimental techniques have been developed
and utilized. Those experimental techniques can be classified in different
ways. One classification distinguishes global and local measurement tech-
niques. For instance, bed expansion technique (i.e. Akita and Yoshida, 1973;
Guy et al., 1986) or static pressure at different points in the column (i.e. Hikita
et al., 1980) were used to measure the overall gas holdup in bubble columns.
Many types of probes, such as optical probes (i.e Abuaf et al., 1978; Cartellier,
1990), resistivity probe (i.e. Herringe and Davis, 1976; Vince et al., 1981), hot
film anemometry (i.e. Wang et al., 1984; Iskandrani and Kojasoy, 2001) and
so on, were used for local gas holdup measurements.
Another classification of the experimental techniques is based on physi-
cal features of the measurement and thus, the experimental techniques are
divided into two categories, that is, invasive and non-invasive measurement
techniques. For instance, those probes used for the local void fraction are cat-
egorized as invasive measuring techniques. Moreover, hot film anemometry
can also be used to obtained liquid-phase characteristics, i.e. mean velocity
and root-mean-square velocity.
There is a large amount of non-invasive techniques which have been uti-
lized to investigate hydrodynamics of bubble column reactors. For instance,

5
1. Introduction

dynamic gas disengagement (DGD) measures the rate at which the liquid
level or the pressure at different levels in the reactor drops after the gas flow
is shut off. In such way, some characteristics, i.e. the gas holdup, bubble
size distribution, etc. can be obtained (Sriram and Mann, 1977; Daly et al.,
1992; Krishna and Ellenberger, 1996). In addition, photographic techniques,
tomographic techniques, radiographic techniques and so on have also been
used in study of multiphase flows. For a detailed review of the measurement
techniques we refer to Chaouki et al. (1997) and Boyer et al. (2002). Some
examples of experimental techniques used in multiphase flows are listed in
Table 1.1.

Table 1.1: Experimental techniques used in multiphase flows.

Global: Bed expansion; Static pressure; Dy-


Gas holdup namic gas disengagement (DGD); Ultrasound
attenuation technique.
Cross-sectional: Wire-mesh sensors; Electri-
cal capacitance tomography (ECT); Electri-
cal resistance tomography (ERT); γ-ray or X-
ray computed tomography; Neutron radiog-
raphy; Magnetic resonance imaging (MRI).
Local: Optical probe; Resistivity probe; Elec-
trochemical probe; Hot film anemometry.
Flow regime transition Visual observation; Pressure fluctuation; Op-
tical probe; Resistivity probe; Heat transfer
probe; Optical transmittance probe; Acous-
tic probe; Electrical capacitance tomogra-
phy (ECT); Electrical resistance tomography
(ERT); γ-ray or X-ray computed tomography;
Neutron radiography; Computer-automated
radioactive particle tracking (CARPT); Laser
Doppler anemometry (LDA); Ultrasound
computed tomography.

6
1.2. Experimental techniques

Bubble properties Photography; Double sensor resistivity


probe; Double optical probe; Four-point opti-
cal fibre probe; Hot film anemometry; Ultra-
sonic Doppler anemometer.
Interfacial area Dynamic absorption; Ultrasound attenua-
tion technique; Light attenuation technique;
Double sensor resistivity probe; Four-sensor
conductivity probe; Four-point optical fibre
probe.
Velocity field Pitot tube; Positron emission particle tracking
(PEPT); Radioactive particle tracking (RPT
or CARPT); Cinematography; Laser Doppler
anemometry (LDA); Particle image velocime-
try (PIV); Particle tracking velocimetry (PTV);
Fluorescent particle image velocimetry; Hot
film anemometry; X-ray based particle track-
ing velocimetry (XPTV); Magnetic resonance
imaging velocimetry.
Heat transfer Thermocouple; Infrared thermography.
Mass transfer Dynamic physical absorption; Chemical ab-
sorption; Physical desorption; Limiting cur-
rent density.
Back mixing Tracers; Computer-automated radioactive
particle tracking (CARPT).
Pressure drop Manometry; Pitot tube.

7
1. Introduction

1.3 Computational Fluid Dynamics

Computational methods for multiphase flows have been developed during


the past decades (Stewart and Wendroff, 1984; Crowe et al., 1996; Kuipers and
van Swaaij, 1997; Marin, 2006; Jakobsen, 2008). In general, there are two major
approaches, that is, the Eulerian-Eulerian model and the Eulerian-Lagrangian
model. The Eulerian-Eulerian model treats both phases as continuous phases
which are inter-penetrating. The Eulerian-Lagrangian model considers the
liquid phase as a continuous phase, while it treats the other phases as a
dispersed phase in form of discrete elements, i.e. particles or bubbles. In
addition, direct numerical simulations (DNS) that are capable of predicting
the interface as well as the flow field of the two phases are also frequently
used in two-phase flow modelling. DNS can be used to obtain closures for
forces acting on discrete elements, such as the drag, lift and virtual mass
(Dijkhuizen, 2008).
For multiphase isothermal systems, the conservation equations for mass
and momentum in the Eulerian-Eulerian model are respectively given by:
∂  
αk ρk + ∇ · αk ρk uk = Rk (1.1)
∂t
∂  
αk ρk uk + ∇ · αk ρk uk uk = −αk ∇p − ∇ · (αk τk ) + M kl + Rk uk + Sk + αk ρk g (1.2)
∂t
where ρk , uk , αk and τk represent, respectively, the macroscopic density, veloc-
ity, volume fraction and viscous stress tensor of the kth phase, p the pressure,
Rk a source term describing mass exchange between phase k and the other
phase, M kl the interphase momentum exchange term between phase k and
phase l and Sk a momentum source term of phase k due to phase changes and
external forces other than gravity.
The Eulerian-Eulerian model is typically used to study large-scale flow
structures and dense dispersed systems due to its lower computational de-
mand compared to the Eulerian-Lagrangian model. A disadvantage of the
model is the need for appropriate closure laws for the interphase transport of
mass, momentum and heat for nonisothermal multiphase systems.
For dispersed multiphase flows, a Lagrangian description of the dispersed
phase generally treats the individual elements as rigid spheres (i.e., neglecting

8
1.3. Computational Fluid Dynamics

particle deformation and internal flows). The translational motion of the


particle is governed by Newton’s second law:

d
(mv) = ΣF (1.3)
dt
where m = ρV is the mass of the element. The term on the right hand side ΣF
denotes net force acting on the particle including surface and body forces, i.e.
gravity, pressure, drag, lift, virtual mass, wall force, etc.
The element trajectory is then calculated as:

dr
=v (1.4)
dt
The interfacial coupling between the phases is considered in different
ways. A simple way is only to consider the impact of the local characteristics
of the continuous phase on the individual element and neglect any effects
that the presence of the dispersed phase may have on the continuous phase.
This is usually referred to as one way coupling and is only valid in systems
with a very low fraction of the dispersed phase. When the fraction of the
dispersed phase is relatively high and the effects of the individual elements
on the continuous phase cannot be ignored, two way coupling is required.
For dense systems, four way coupling is necessary to take into account the
additional collision effects between elements.
The Eulerian-Lagrangian model has the advantage of considering the mi-
croscopic transport phenomena by taking into account the direct collision
and hydrodynamic interaction between neighboring elements. Moreover, in
bubble column reactors, the residence time of the gas phase in the form of
bubbles can be easily obtained with the Eulerian-Lagrangian model. Hence,
it is possible to study the back mixing of the gas phase with the theory of
residence time distribution (RTD) and then evaluate the performance of the
bubble column reactor.
A disadvantage of the Eulerian-Lagrangian approach is its relatively high
computational cost for the dispersed phase, particularly in very dense systems
which are common in industrial multiphase chemical reactors. Even though
efforts can be made to improve the model, i.e. improving the efficiency of the

9
1. Introduction

algorithm for dealing with collisions (Hoomans et al., 1996; Wu et al., 2010), it
may not be feasible for the Eulerian-Lagrangian model to keep track of a high
number of discrete elements of the dispersed phase.

1.4 Objectives & outline

The objective of this thesis is to study the bubble properties, i.e. bubble
velocity, bubble size and local void fraction, in heterogeneous flow regime
with a four-point optical fibre probe and to further develop and improve the
Eulerian-Lagrangian model in the study of fluid dynamics, back mixing of
both the gas and the liquid phase and performance of bubble column reactors.
In the experimental work, the accuracy of a four-point optical fibre probe is
investigated with a high speed camera in single bubble experiments. Bubble
properties in the heterogeneous flow regime are then measured with the four-
point optical probe. Effects of the superficial gas velocity and the column
height on the bubble properties are studied. In addition, effect of the gas
distributor on the hydrodynamics and back mixing of the gas phase in bubble
column reactors is investigated by using the Eulerian-Lagrangian model and
the theory of residence time distribution (RTD). The swarm effects on the drag
force in bubble column are studied by comparing simulations using various
drag coefficient correlations from literature and measurements using particle
image velocimetry (PIV). Breakup models in literature are implemented in
the Eulerian-Lagrangian model to consider the coalescence and breakup of
bubbles in turbulent flows. Finally, massless tracer particles which move
with the liquid phase are introduced into the model. Moreover, overall gas
holdup and back mixing of both the gas and the liquid phases are studied and
compared with correlations in literature.
The contributions to these topics are organized in chapters as follows:
Chapter 2 investigates the performance of a four-point optical fibre probe
for the determination of the bubble velocity. Single bubble experiments are
carried out and five liquids with different physical properties are used. The
results obtained from the four-point optical fibre probe are compared with
those obtained by image processing using a high speed camera. The accu-

10
Nomenclature

racy of the bubble velocity determination from the optical probe is discussed
according to the physical properties of the liquid phase.
Chapter 3 studies the bubble properties in the heterogeneous flow regime
in a square bubble column with the four-point optical fibre probe. The per-
formance of the four-point optical fibre probe is further discussed. Bubble
size and specific interfacial area are estimated according to the measurements
from the optical probe. Furthermore, the effects of the column height and the
gas flow rate on the bubble properties are investigated.
Chapter 4 discusses the effect of the gas distributor on the hydrodynamics
and back mixing of the gas phase in a square bubble column by using an
Eulerian-Lagrangian model. Theory of residence time distribution (RTD) is
adopted to study the back mixing of the gas phase.
Chapter 5 studies the drag coefficient correlations for bubble swarms in
literature with the aid of the Eulerian-Lagrangian model. The simulation
results are compared with the measurements of particle image velocimetry
(PIV).
Chapter 6 implements breakup models reported in literature in the
Eulerian-Lagrangian model. Furthermore, critical Weber numbers related
to bubble breakup in turbulent flows are also considered and combined with
a coalescence model. Simulation results of the Eulerian-Lagrangian model are
compared with PIV measurements. Moreover, the bubble size distributions
from the numerical simulations are compared with those obtained from the
four-point optical fibre probe.
Chapter 7 discusses the performance of a square bubble column. Massless
tracer particles are introduced into the Eulerian-Lagrangian model which
move with the liquid phase in the bubble column. Gas holdup and back
mixing of both the gas and liquid phases are investigated and compared with
correlations reported in literature.

Nomenclature

F force acting on a discrete element, [N]


g gravitational acceleration, [m s−2 ]

11
1. Introduction

m mass of a discrete element, [kg]


Mkl momentum source term due to interaction between phase k and
phase l, [kg m−2 s−2 ]
p pressure, [N m−2 ]
r position vector, [m]
Rk mass source term in phase k, [kg m−3 s−1 ]
Sk momentum source term due to phase changes and external forces
other than gravity in phase k, [kg m−2 s−2 ]
t time, [s]
u liquid velocity, [m s−1 ]
v velocity of a discrete element, [m s−1 ]

Greek letters

α volume fraction, [-]


ρ density, [kg m−3 ]
τ stress tensor, [N m−2 ]

References

N. Abuaf, O. C. Jones Jr., and G. A. Zimmer. Optical probe for local void frac-
tion and interface velocity measurements. Review of Scientific Instruments,
49(8):1090–1094, 1978.

K. Akita and F. Yoshida. Gas holdup and volumetric mass transfer coefficient
in bubble columns: Effects of liquid properties. Ind. Eng. Chem. Process Des.
Develop., 12(1):76–80, 1973.

A. A. Avdeev, B. F. Balunov, and V. I. Kiselev. Heat transfer in bubble layers at


high pressures. Experimental Thermal and Fluid Science, 5(6):728–735, 1992.

C. Boyer, A. -M. Duquenne, and G. Wild. Measuring techniques in gas-


liquid and gas-liquid-solid reactors. Chemical Engineering Science, 57(16):
3185–3215, 2002.

12
References

A. J. Carleton, R. J. Flain, J. Rennie, and F. H. H. Valentin. Some properties


of a packed bubble column. Chemical Engineering Science, 22(12):1839–1845,
1967.

A. Cartellier. Optical probes for local void fraction measurements: Charac-


terization of performance. Review of Scientific Instruments, 61(2):874–886,
1990.

J. Chaouki, F. Larachi, and M. P. Duduković. Noninvasive tomographic and


velocimetric monitoring of multiphase flows. Industrial and Engineering
Chemistry Research, 36(11):4476–4503, 1997.

C. T. Crowe, T. R. Troutt, and J. J. N. Chung. Numerical models for two-phase


turbulent flows. Annual Review of Fluid Mechanics, 28:11–43, 1996.

J. G. Daly, S. A. Patel, and D. B. Bukur. Measurement of gas holdups and


sauter mean bubble diameters in bubble column reactors by dynamics gas
disengagement method. Chemical Engineering Science, 47(13-14):3647–3654,
1992.

W. -D. Deckwer. Bubble column reactors. Chichester [etc.] : Wiley and sons,
1992.

W. -D. Deckwer, R. Burckhart, and G. Zoll. Mixing and mass transfer in tall
bubble columns. Chemical Engineering Science, 29(11):2177–2188, 1974.

J. M. Delhaye and P. Bricard. Interfacial area in bubbly flow: Experimental


data and correlations. Nuclear Engineering and Design, 151(1):65–77, 1994.

W. Dijkhuizen. Derivation closures for bubbly flows using direct numerical simu-
lations. PhD thesis, Enschede, April 2008.

J. Dudley. Mass transfer in bubble columns: A comparison of correlations.


Water Research, 29(4):1129–1138, 1995.

J. R. Fair, A. J. Lambright, and J. W. Andersen. Heat transfer and gas holdup


in a sparged contactor. IandEC Process Design and Development, 1(1):33–36,
1962.

13
1. Introduction

S. D. Gharat and J. B. Joshi. Transport phenomena in bubble colunm reactor


ii: Pressure drop. The Chemical Engineering Journal, 48(3):153–166, 1992.

S. Guet, R. V. Fortunati, R. F. Mudde, and G. Ooms. Bubble velocity and


size measurement with a four-point optical fiber probe. Particle and Particle
Systems Characterization, 20(3):219–230, 2003.

C. Guy, P. J. Carreau, and J. Paris. Mixing characteristics and gas hold-up of a


bubble column. Canadian Journal of Chemical Engineering, 64(1):23–35, 1986.

J. J. Heijnen and K. Van’t Riet. Mass transfer, mixing and heat transfer phe-
nomena in low viscosity bubble column reactors. The Chemical Engineering
Journal, 28(2):B21–B42, 1984.

R. A. Herringe and M. R. Davis. Structural development of gas-liquid mixture


flows. Journal of Fluid Mechanics Digital Archive, 73(01):97–123, 1976.

H. Hikita and H. Kikukawa. Liquid-phase mixing in bubble columns: Effect


of liquid properties. The Chemical Engineering Journal, 8(3):191–197, 1974.

H. Hikita, S. Asai, K. Tanigawa, K. Segawa, and M. Kitao. Gas hold-up in


bubble columns. The Chemical Engineering Journal, 20(1):59–67, 1980.

B. P. B. Hoomans, J. A. M. Kuipers, W. J. Briels, and W. P. M. van Swaaij. Dis-


crete particle simulation of bubble and slug formation in a two-dimensional
gas-fluidised bed: A hard-sphere approach. Chemical Engineering Science,
51(1):99–118, 1996.

A. Iskandrani and G. Kojasoy. Local void fraction and velocity field descrip-
tion in horizontal bubbly flow. Nuclear Engineering and Design, 204(1-3):
117–128, 2001.

H. A. Jakobsen. Chemical Reactor Modeling: Multiphase Reactive Flows. Springer,


1st edition, July 2008.

J. B. Joshi. Axial mixing in multiphase contactors - A unified correlation.


Transactions of the Institution of Chemical Engineers, 58(3):155–165, 1980.

14
References

J. B. Joshi. Gas phase dispersion in bubble columns. The Chemical Engineering


Journal, 24(2):213–216, 1982.

N. Kantarci, F. Borak, and K. O. Ulgen. Bubble column reactors. Process


Biochemistry, 40(7):2263–2283, 2005.

I. Kataoka, M. Ishii, and A. Serizawa. Local formulation and measurements


of interfacial area concentration in two-phase flow. International Journal of
Multiphase Flow, 12(4):505–529, 1986.

Y. Kawase and M. Moo-Young. Liquid phase mixing in bubble columns with


newtonian and non-newtonian fluids. Chemical Engineering Science, 41(8):
1969–1977, 1986.

Y. Kawase and M. Moo-Young. Theoretical prediction of gas hold-up in


bubble columns with newtonian and non-newtonian fluids. Industrial and
Engineering Chemistry Research, 26(5):933–937, 1987.

S. L. Kiambi, A. M. Duquenne, A. Bascoul, and H. Delmas. Measurements


of local interfacial area: Application of bi-optical fibre technique. Chemical
Engineering Science, 56(21-22):6447–6453, 2001.

R. Krishna and J. Ellenberger. Gas holdup in bubble column reactors operating


in the churn-turbulent flow regime. AIChE Journal, 42(9):2627–2634, 1996.

R. Krishna, P.M. Wilkinson, and L.L. Van Dierendonck. A model for gas
holdup in bubble columns incorporating the influence of gas density on flow
regime transitions. Chemical Engineering Science, 46(10):2491–2496, 1991.

J. A. M. Kuipers and W. P. M. van Swaaij. Application of computational fluid


dynamics to chemical reaction engineering. Reviews in Chemical Engineering,
13(3):1–118, 1997.

T. -J. Lin and L. -S. Fan. Heat transfer and bubble characteristics from a
nozzle in high-pressure bubble columns. Chemical Engineering Science, 54
(21):4853–4859, 1999.

15
1. Introduction

S. Luther, J. Rensen, and S. Guet. Bubble aspect ratio and velocity measure-
ment using a four-point fiber-optical probe. Experiments in Fluids, 36(2):
326–333, 2004.

S. K. Majumder. Analysis of dispersion coefficient of bubble motion and


velocity characteristic factor in down and upflow bubble column reactor.
Chemical Engineering Science, 63(12):3160–3170, 2008.

S. K. Majumder, G. Kundu, and D. Mukherjee. Prediction of pressure drop


in a modified gas-liquid downflow bubble column. Chemical Engineering
Science, 61(12):4060–4070, 2006.

A. Manera, B. Ozar, S. Paranjape, M. Ishii, and H. M. Prasser. Comparison be-


tween wire-mesh sensors and conductive needle-probes for measurements
of two-phase flow parameters. Nuclear Engineering and Design, 239(9):1718–
1724, 2009. 15th International Conference on Nuclear Engineering (ICONE
15).

G. B. Marin, editor. Advances in Chemical Engineering, volume 31. Elsevier Inc.,


2006.

J. Merchuk, S. Ben-Zvi, and K. Niranjan. Why use bubble-column bioreactors?


Trends in Biotechnology, 12(12):501–511, 1994.

E. J. Molga and K. R. Westerterp. Experimental study of a cocurrent upflow


packed bed bubble column reactor: pressure drop, holdup and interfacial
area. Chemical Engineering and Processing, 36(6):489–495, 1997.

Y. Ohki and H. Inoue. Longitudinal mixing of the liquid phase in bubble


columns. Chemical Engineering Science, 25(1):1–16, 1970.

S. T. Revankar and M. Ishii. Theory and measurement of local interfacial area


using a four sensor probe in two-phase flow. International Journal of Heat
and Mass Transfer, 36(12):2997–3007, 1993.

M. C. Ruzicka, J. Zahradnı́k, J. Drahos, and N. H. Thomas. Homogeneous-


heterogeneous regime transition in bubble columns. Chemical Engineering
Science, 56(15):4609–4626, 2001.

16
References

S. C. Saxena and N. S. Rao. Heat transfer and gas holdup in a two-phase


bubble column: Air-water system - review and new data. Experimental
Thermal and Fluid Science, 4(2):139–151, 1991.

S. Schlüter, A. Steiff, and P. M. Weinspach. Heat transfer in two- and three-


phase bubble column reactors with internals. Chemical Engineering and
Processing, 34(3):157–172, 1995.

Y. T. Shah, B. G. Kelkar, S. P. Godbole, and W. -D. Deckwer. Design parameters


estimations for bubble column reactors. AIChE Journal, 28(3):353–379, 1982.

A. Shaikh and M. H. Al-Dahhan. A review on flow regime transition in bubble


columns. International Journal of Chemical Reactor Engineering, 5, 2007.

K. Sriram and R. Mann. Dynamic gas disengagement: A new technique for


assessing the behaviour of bubble columns. Chemical Engineering Science, 32
(6):571–580, 1977.

H. B. Stewart and B. Wendroff. Two-phase flow: Models and methods. Journal


of Computational Physics, 56(3):363–409, 1984.

M. J. Tan and M. Ishii. A method for measurement of local specific interfacial


area. International Journal of Multiphase Flow, 16(2):353–358, 1990.

A. K. Verma. Heat transfer mechanism in bubble columns. The Chemical


Engineering Journal, 42(3):205–208, 1989.

M. A. Vince, G. Krycuk, and R. T. Lahey Jr. Development of a radio frequency


excited local impedance probe. Nuclear Engineering and Design, 67(1):125–
136, 1981.

S. Wachi and Y. Nojima. Gas-phase dispersion in bubble columns. Chemical


Engineering Science, 45(4):901–905, 1990.

K. B. Wang and L. T. Fan. Mass transfer in bubble columns packed with


motionless mixers. Chemical Engineering Science, 33(7):945–952, 1978.

17
1. Introduction

S. K. Wang, S. J. Lee, O. C. Jones Jr., and R. T. Lahey Jr. Local void frac-
tion measurement in techniques in two-phase bubbly flows using hot-film
anemometry. volume 31, pages 39–45, 1984.

K. R. Westerterp, W. P. M. van Swaaij, and A. A. C. M. Beenackers. Chemical


Reactor Design and Operation. John Wiley & Sons Ltd., 2nd edition, 1987.

C. L. Wu, A. S. Berrouk, and K. Nandakumar. An efficient chained-hash-table


strategy for collision handling in hard-sphere discrete particle modeling.
Powder Technology, 197(1-2):58–67, 2010.

P. Zehner. Momentum, mass and heat transfer in bubble columns. part 2. axial
blending and heat transfer. International chemical engineering, 26(1):29–35,
1986.

18
Chapter
2
Single bubble experiment
For the understanding of high void fraction flows in bubble columns, quantitative
information on bubble size, bubble velocity as well as gas void fraction is of crucial
importance. Optical probes have the advantage of low cost, simplicity of operation and
easy interpretation of the results. Moreover optical probes are well-suited to obtain
bubble properties in bubbly flows at high void fraction.
In this chapter, the performance of a four-point optical fibre probe was investigated
by comparing probe data with results obtained with photography. Five liquids with
different material properties were used in combination with air sparged from the dis-
tributor at the bottom a flat bubble column.
It is found that the liquid properties have a significant influence on the bubble ve-
locity obtained from the optical probe. In viscous liquids, the bubble deforms and is
decelerated by the optical probe during the approach process and thus the inaccuracy
of velocity determination results. In low viscosity liquids, the bubble’s wobbling
behavior also results in inaccuracy.

19
2. Single bubble experiment

2.1 Introduction

Two-phase gas-liquid flows are frequently encounted in variety of industrial


processes, such as bubble column reactors, stirred tank reactors, waste water
treatment units and so on. Reliable reactor design and scale-up require data
on gas holdup, pressure drop, flow regime, heat and mass transfer rates and
flow patterns of the phases.
Direct optical techniques i.e. particle image velocimetry (Lindken et al.,
1999; Deen, 2001; Bröder and Sommerfeld, 2002) which are commonly used
in gas-liquid flows, are only applicable in bubbly flows at relatively low gas
hold-up.
Application of optical probes as an alternative technique for measurements
in multiphase flow offers the advantages of low cost, simplicity of operation
and relative straightforward interpretation of results. Optical probes can be
used to measure local void fraction, bubble frequency, bubble velocity, time-
average local interfacial area and mean bubble chord length. During the past
decades, optical probes have been intensively utilized in multiphase flows
(Jones and Delhaye, 1976; Boyer et al., 2002).
Multiple point probes have been proposed. Since such probes offer, in
principle, the possibility to determine all three components of the bubble ve-
locity. For instance, four-point optical probes were used to determine bubble
velocity and size (Guet et al., 2003; Luther et al., 2004).
Besides the ability of measuring bubble velocity, multiple point optical
probes are able to recognize those bubbles with a direction of motion deviating
from the axial direction and thus, leading to reduced errors in axial component
of the bubble velocity.
The objective of the present study is to investigate performance of a four-
point optical fibre probe with respect to the bubble velocity measurement.
Furthermore, the influence of material properties of the liquid phase on the
accuracy of the velocity determination is also investigated.
For this purpose, single bubble experiments were conducted. A high speed
camera was used to record images of the approach of a bubble rising in a flat
bubble column towards a four-point optical fibre probe and the subsequent

20
2.2. Experimental setup

piercing of the bubble by the tips of the probe. Meanwhile, signals generated
by the four-point optical fibre probe were recorded simultaneously. Images
taken by the camera and signals generated by the four-point optical fibre probe
were processed afterwards separately. Velocities from the image processing
were treated as reference velocities to evaluate the velocities obtained from the
four-point optical fibre probe. Five liquids with different material properties,
including, viscosity, surface tension, density were used during measurements.

2.2 Experimental setup

The experimental setup mainly consists of a small flat bubble column (10 ×
110 × 500 mm), an Imager Pro HS CMOS camera with 12 bit, 1024 × 1280
pixel resolution, a four-point optical probe and a light source, which are
schematically shown in Figure 2.1.

Figure 2.1: Sketch of experimental setup.

During the experiments, an air bubble was released from a small hole in
the center of the bottom plate with a diameter of approximately 1 mm. The
range of produced bubble sizes ranged from 3 to 5 mm.

21
2. Single bubble experiment

The four-point optical probe was positioned at the top of the bubble col-
umn, a few centimeters below the liquid surface. The signals of the four-point
optical probe were read with a LabView program and stored onto a hard disk.
The high speed camera is mounted in front of the bubble column and
a light source illuminated the bubble column from behind, thus employing
a shadowgraphy technique. Shadow images of bubble and the four-point
optical probe were recorded by the camera and stored on the computer. The
image acquisition program is DaVis from LaVision. The applied field of view
of the camera is 700 × 1024 pixels.
The four-point optical probe has three tips of the same length that form
an equivalent angle. The fourth tip is positioned in the center between the
other three tips. The radial distance d = 0.5 mm and the vertical distance is
∆s = 1.4 mm. The geometrical configuration of the probe tips is shown in
Figure 2.2.

Figure 2.2: Geometry of four-point optical probe.

The principle of the optical probe is based on different refractive indices


of light in both the gas and the liquid phase. For instance, the signal of the
optical probe remains low when the probe is immersed in the liquid phase,
whereas the signal increases to a high level in the gas phase due to total
reflection of light. With sufficient signal to noise ratio, the two phases can be
distinguished. Consequently, by using a certain separation method, the gas
phase could be easily retrieved from the output signal.
Three different liquids and two aqueous mixtures with different mass frac-

22
2.3. Data processing

tion of glycerol were used during the measurements. The physical properties
(at room temperature) of each of the selected liquids are listed in Table 2.1.

Table 2.1: Physical properties of employed liquids at room temperature.

Density Viscosity Surface tension


[kg/m3 ] [mPa·s] [mN/m]
Decane 730 0.838 23.37
Diethylene glycol 1120 30.2 44.7
Glycerol(60%) 1153 10.7 66.9
Glycerol(72%) 1187 27.6 66.5
Water 998 1.00 72.75

2.3 Data processing

2.3.1 Digital image processing


The intensity images taken by the camera were processed with MatLab. In
the image, a small region containing the bubble was cropped and then, an
image segmentation method was employed within this small region to dis-
tinguish the bubble contours. Image segmentation was based on the method
of Otsu (1979), which is a nonparametric and unsupervised method of auto-
matic threshold selection for picture segmentation. An optimal threshold was
selected by the discriminant criterion, namely, maximizing the separability of
the resultant classes in gray levels. A brief description of the method is given
below:
An image consists of L gray levels [1, 2, ..., L] and the number of pixels at
gray level i are denoted by ni . In addition, the total number of pixels N is
obtained by summing ni over all levels, N = n1 + n2 + ... + nL . The fraction of
pixels at level i is given by:

pi = ni /N, pi ≤ 0 (2.1)

X
L
pi = 1 (2.2)
i=1

23
2. Single bubble experiment

Suppose the pixels are dichotomized into two classes C0 and C1 (back-
ground and objects, or vice versa) by a threshold at level k. C0 denotes pixels
with levels [1, ..., k], and C1 denotes pixels with levels [k + 1, ..., L]. The prob-
abilities of class occurrence, ω and the class mean levels, µ, respectively are
given by
X
k
ω0 = P(C0 ) = pi = ω(k) (2.3)
i=1

X
L
ω1 = P(C1 ) = pi = 1 − ω(k) (2.4)
i=k+1

and
X
k X
k
µ(k)
µ0 = iP(i|C0 ) = ipi /ω0 = (2.5)
ω(k)
i=1 i=1

X
L X
L
µT − µ(k)
µ1 = iP(i|C0 ) = ipi /ω1 = (2.6)
1 − ω(k)
i=k+1 i=k+1

The total mean level of the original picture is given by:

X
L
µT = µ(L) = ipi (2.7)
i=1

The between-class variance is given by:

σ2B = ω0 (µ0 − µT )2 + ω1 (µ1 − µT )2 = ω0 ω1 (µ1 − µ0 )2 (2.8)

The threshold k∗ that maximizes σ2B is treated as optimal threshold to segment


the objects from the background in the image.
Sezgin and Sankur (2004) compared some image thresholding techniques
and evaluated their performance quantitatively. By using several thresh-
olding performance criteria, Otsu’s method performed excellent for image
segmentation within 40 image thresholding methods. Sezgin and Sankur
(2004) also reported that Otsu’s method can give satisfactory results when the
number of pixels in each class are close to each other.
Properties of the bubble, such as area, centre of mass, perimeter, equivalent
diameter, orientation, major axis length, minor axis length and so on, can be

24
2.3. Data processing

calculated through manipulating the pixels representing the bubble in the


image. After this step, the co-ordinates of the centre of mass obtained from
the cropped region were transformed back to the coordinates system of the
original image.
Subsequently, the vertical component of the bubble velocity was calculated
from the following equation in a sequence of images:

zi+1 − zi
vi = (2.9)
∆t
where z is the vertical co-ordinate of the centre of mass of the bubble and ∆t
is the time difference between two subsequent images.
Three characteristic velocities were calculated separately. One is the aver-
age velocity before the bubble hits the probe, which will be termed the terminal
velocity. The second characteristic velocity is the velocity at the moment of
interaction between the bubble and the longest tip of the optical probe. The
third one is the average velocity during the bubble-probe contact, which will
be termed the piercing velocity. These three velocities will be used as a basis
for comparison with the velocity obtained from the four-point optical probe.

2.3.2 Signal processing of optical probe


The main task of the signal processing of the optical probe data is to identify
the parts representing the bubble in the raw signal. That is, the moments of
probe entry and exit through each bubble need to be found in each pulse. The
idea to find the relevant moments is based on different levels in the raw data
signal, i.e. the liquid level, the gas level and the noise level (Cartellier, 1992;
Barrau et al., 1999; Harteveld, 2005). “Pre-signals” were observed in some
bubble piercing events, i.e. as shown in Figure 2.6(a). In viscous liquids,
such as glycerol solutions, the possible reason for these pre-signals is that
the bubble surface is perpendicular as it approaches the tip. This leads to
detectable reflections by the surface prior to piercing. In low viscosity liquids,
wobbling behavior of bubbles with a moderate size may also result in the
occurrence of a pre-signal. Therefore, a criterion based on the noise level to
determine the entry time is debatable. Harteveld (2005) suggested that the

25
2. Single bubble experiment

threshold should be set to 10% of the bubble plateau level. However, it is hard
to remove pre-signals in this way if the signal-to-noise ratio of the probe is
low.
In the present study, the entry and exit moments were determined on basis
of the following equations:

Ventry = VL + 0.1(VG − VL ) (2.10)

Vexit = VL − 0.1(VG − VL ) (2.11)

and thus identical to the threshold used by Harteveld (2005). In some cases
where a pre-signal occurs, additional processing was employed. The entry
point was selected accordingly at the very beginning of the ascending ramp
after the primary peak.

Figure 2.3: Signals and corresponding time differences.

Based on above steps, the moments of probe entry and exit can be deter-
mined for each bubble. With such information, the time difference between
the upper surface of a bubble hitting the longest tip and the other three tips
can be derived as show in Figure 2.3.
The velocity of bubble was calculated from the following equation:

∆s ∆s
u= = P , i = 1, 2, 3 (2.12)
∆t 1
3 ∆ti

26
2.3. Data processing

where ∆s is the vertical distance between the longest tip and the other three
tips, whereas ∆ti is the time difference between hitting of the longest tip and
short tip i. Due to the dynamics of a bubble during the interaction with the
probe, these three time intervals may vary from one another. This is amongst
others due to the bubble hitting the probe off-centre and irregular deformation
of the upper bubble surface during the interaction. To reduce the error in the
bubble velocity determination, the following selection criterion was used:

∆ti − ∆t
∆t < β, i = 1, 2, 3 (2.13)

Mudde and Saito (2001) and Fortunati et al. (2002) studied the influence of
this criterion on the accuracy of the bubble velocity both numerically and ex-
perimentally. In Mudde’s simulations, in case 8% < β < 20% the influence on
the accuracy is negligible in comparison to other error sources. In Fortunati’s
experiment, no significant effect was observed on the average rise velocity
and its standard deviation in case β < 25%. In the present single bubble
experiments, β = 30% was adopted in order to allow more bubbles to be con-
sidered. The choice of this selection criterion will be discussed in more detail
in chapter 3. Moreover, the chord length of bubble can then be determined
from the bubble velocity u and the time span for the bubble travelling through
the optical probe t, which will also be introduced in the following chapter.

2.3.3 Uncertainty
During image processing, the uncertainty in the instantaneous bubble veloc-
ity is related to the image acquisition rate, the determination of geometrical
properties (i.e. center of mass of the bubble) and applied magnification factor.
The magnification factor is obtained by taking the diameter of the probe sup-
port as a reference length. The DaVis software is used to acquire images and
the time interval between sequential images is very accurate. The accuracy
of geometrical properties of the bubble, such as, centre of mass of the bubble,
bubble diameter and so on, mainly depends on the performance of the image
segmentation method provided that the acquired image has good contrast
between bubble and its background.

27
2. Single bubble experiment

Due to inaccuracies in the manufacturing of the probe, the length of the


three short tips of the probe may not be exactly the same. The distances
between the longest tip and the other three short tips differ slightly. Fur-
thermore, the three time differences used to calculate the bubble velocity also
varied because of unpredictable interaction between the bubble and the four-
point optical probe. These two influences were considered to analyze the
uncertainty in the bubble velocity obtained from the four-point optical probe.

2.4 Results and discussion

2.4.1 Interaction between bubble and probe


Due to different physical properties of the liquids, the behavior of the bubble
when approaching the tips of the probe differs. Particularly, the viscosity of
the liquid influences the bubble shape significantly.
There are two main different interactions between the bubble and the four-
point optical probe in the liquid. For instance, a rising bubble in a viscous
liquid keeps its shape after being released from the bottom of the column and
only deforms when it gets pierced by the probe . On the contrary, however, it
is hard to determine if the deformation of the bubble is due to the interaction
with the probe in liquids with low viscosity, such as decane and water. In such
liquids, bubbles start wobbling as soon as they are injected into the liquid and
the shape of the bubbles changes continuously.
A set of images depicting an air bubble rising in an aqueous mixture with
72% of glycerol are shown in Figure 2.4. The bubble retains an ellipsoidal
shape before encountering the probe. It can clearly be seen that the shape of
the rising bubble changes during interaction with the tips of the optical probe.
The roof of the bubble becomes flatter as it hits the probe. The retarding effect
of the probe is more obvious for bubbles using in viscous liquids.
In Figure 2.5, continuous movements of a bubble rising in water are
recorded. The shape of the bubble changes continuously during the entire
path in the column and it appears that the probe has almost no effect on the
bubble shape. Furthermore, the retarding effect of the probe on the bubble
is negligible. This retarding effect of liquids with different viscosity on rising

28
2.4. Results and discussion

(a) t = 0.000 s (b) t = 0.008 s (c) t = 0.016 s


Figure 2.4: Rising bubble in a glycerol solution (72%).

bubbles will be discussed in the following sections.


Furthermore, the signals of the probe change when the surface of the
bubble passes through the tips of the probe and the processing program
finds the bubble according to these changes of the signals. Therefore, the
wobbling behavior of bubbles in low viscous liquid may result in inaccurate
determination of bubble velocities.

(a) t = 0.000 s (b) t = 0.006 s (c) t = 0.012 s


Figure 2.5: Rising bubble in water.

29
2. Single bubble experiment

2.4.2 Bubble rise velocity


The signals of the probe corresponding to the above two bubble-probe inter-
actions (Figure 2.4 & Figure 2.5) are plotted in Figure 2.6. The times corre-
sponding to tips of the probe entering the bubble can be extracted from the
signals according to the signal processing method mentioned above. Thus,
the average time difference between the central tip and the other three tips
and thus the velocity of the bubble can be determined by knowing the vertical
distance between the central tip and other tips. In the meantime, since the
trajectory of the bubble is recorded by the high speed camera, the bubble
velocity can also be calculated from the image processing method.
Figures 2.7 and 2.8 show the instantaneous positions and velocities that
were recorded with the camera and velocity determined from the optical
probe. Note that in the plots, the origin of coordinates is defined at the bot-
tom of the central tip. Due to effects of physical properties, such as viscosity,
surface tension and the size of the bubble, the trajectories of the bubble in
the low viscosity liquid are different from those of the bubble in the viscous
liquid. In the low viscosity liquids, the bubble positions show more fluctua-
tions compared to bubbles rising in viscous liquids because of the wobbling
behaviour, which is reflected in the corresponding velocity plots. According
to both figures, however, it can be found that the vertical component of the
bubble position increases approximately linearly with time. Therefore, the
instantaneous velocities prior to piercing can be used to calculate the average
terminal bubble velocity in liquids of both low and high viscosity.
In viscous liquid, the ellipsoidal shape of the bubble is affected during
the interaction with the probe. As a result, the roof of the bubble starts to
become flat slowly. Meanwhile, the bubble is decelerated which is reflected
by the short curved part of plots of the relative position of the bubble in
Figure 2.7(a). Consequently, the vertical velocity of the bubble decreases.
The retarding effect of the probe in viscous liquid is obvious. The bubble
velocity from the optical probe is much smaller than the average bubble
velocity before approaching the probe, which means that the measurement of
the bubble velocity with the optical probe may induce considerable errors in
viscous liquids.

30
2.4. Results and discussion

0 0
Tip0 Tip0
Tip1 Tip1
Tip2 Tip2
Signal [Voltage]

Signal [Voltage]
Tip3 Tip3
-2 -2

-4 -4

-6 -6
-0.05 0 0.05 0.1 -0.02 -0.01 0 0.01 0.02 0.03 0.04
Relative time [s] Relative time [s]

(a) Glycerol solution (72%) (b) Water

Figure 2.6: Example signals of the optical probe.

0.2
Relative position [m]

0
Velocity [m/s]

0.15
-0.01

-0.02 0.1 From image


From probe

0 0.05 0.1 0.15 -0.02 -0.01 0 0.01


Time [s] Relative position [m]

(a) Relative position vs. time (b) Vertical velocity vs. relative position

Figure 2.7: Air-glycerol solution (72%) system.

However, the optical probe has little retarding effect on bubbles rising
in low viscosity liquid (i.e. water). It is hard to distinguish changes of the
instantaneous bubble velocity due to the presence of the probe from those
due to the wobbling behavior of the bubble. It is also found that the bubble
velocity from the optical probe is within the range of fluctuation of the bubble
velocity.

31
2. Single bubble experiment

Relative position [m]

Velocity [m/s]
0.25

-0.01

0.2

-0.02 From image


0.15 From probe

0 0.05 0.1 -0.02 -0.01 0


Time [s] Relative position [m]

(a) Relative position vs. time (b) Vertical velocity vs. relative position

Figure 2.8: Air-water system.

2.4.3 Comparison of bubble velocities

The bubble velocities obtained from both the high speed camera and the
optical probe are compared with the use of parity plots. The terminal bubble
velocities obtained from image processing are used to compare with those
obtained from the optical probe (Figure 2.9). The terminal bubble velocity
is determined by averaging the instantaneous bubble velocities before the
bubble approached the optical probe. Since the field of view of the camera is
far away from the bubble injector, the bubble is believed to rise at its terminal
velocity in the region of interest. This can also be motivated from the plots of
the bubble displacement versus time (i.e. Figure 2.7(a) & 2.8(a)). However, the
presence of the optical probe in the path of the rising bubble seems to obstruct
the rising bubble. For instance, it is obvious that the bubble velocity reduces
slightly when approaching the probe in Figure 2.7(b). Therefore, in order
to evaluate the performance of the optical probe, the bubble velocity at the
moment of interaction with the longest tip of the optical probe is also used as a
basis for comparison with results obtained from the optical probe (Figure 2.10).
In addition, the velocity obtained from the four-point optical fibre probe is the
one which is neither the terminal velocity nor the instantaneous velocity of the
bubble at the moment that the roof of the bubble touches the longest tip of the
probe. This velocity is actually measured during the deceleration of the bubble

32
2.4. Results and discussion

during the contact with the probe. Hence, a comparison between the velocity
obtained from the probe and an average velocity during the interaction with
the probe obtained from the high speed camera is presented as well, as shown
in Figure 2.11.

Diethylene glycol Decane


Glycerol (60%) Water
0.25 Glycerol (72%)
0.25
Optical probe

Optical probe
0.2
0.2
0.15

0.15
0.1

0.05 0.1
0.05 0.1 0.15 0.2 0.25 0.1 0.15 0.2 0.25
HS camera HS camera

(a) Viscous liquids (b) Low viscosity liquids

Figure 2.9: Parity plots of terminal bubble velocities.

The straight lines in the parity plots are identity. In Figure 2.9, it can be seen
that the presence of the four-point optical probe has a considerable retarding
effect on the bubble velocity in viscous liquids. A large viscosity of the liquid
causes deceleration of the bubble movement along the tip. Accordingly, inac-
curacies arise during bubble velocity measurements using the optical probe.
For instance, the average deviation between the terminal bubble velocity and
velocity measured with the probe is about 45% in diethylene glycol. In the
glycerol solution, the deviation is around 33%. In the glycerol solution with
60% glycerol, the deviation is about 29%.
However, the terminal bubble velocities in low viscosity liquids are much
closer to those obtained with the camera. Meanwhile, one can find that
there is significant scatter of the bubble velocity measured by the optical
probe. The wobbling behavior of the air bubble in low viscosity liquids
induces some difficulties for measuring the bubble velocity. The shape of the
bubble is already changing continuously when it hits the tips. The velocity
measurement by the probe is strongly influenced by the dynamics of the

33
2. Single bubble experiment

bubble surface . On the contrary, the obstructing effect of the probe on the
rising bubble is not so significant. The average deviation between the terminal
bubble velocity and the one measured with the probe is about 18% in decane
and 20% in water.

Diethylene glycol Decane


Glycerol (60%) Water
0.25 Glycerol (72%) 0.25
Optical probe

Optical probe
0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05
0.05 0.1 0.15 0.2 0.25 0.05 0.1 0.15 0.2 0.25
HS camera HS camera

(a) Viscous liquids (b) Low viscosity liquids

Figure 2.10: Parity plots of instantaneous bubble velocities.

In Figure 2.10(a), it can be seen that points in the plot are much closer
to the identity compared with the points in Figure 2.9(a). This can also be
seen from the average deviations. The average deviation between the instant
bubble velocity and that obtained from the optical probe in diethylene glycol
reduces from 45% to 40%. In the two glycerol solutions, the average deviation
also decreases. In glycerol solution (72%), it is about 25%, whereas in the
glycerol solution with 60% glycerol, the deviation is 25%. However, there is
no difference in the average deviation in the two low viscosity liquids.
According to the comparison shown in Figure 2.11(a), one can find that
the discrepancy between the velocity obtained from the optical probe and
the average velocity during the interaction determined from the high speed
camera gets smaller. For instance, the average deviation between the bubble
velocity obtained from the optical probe and the average velocity during the
interaction from image processing in diethylene glycol is 34%. In the glycerol
solutions, the average deviations for the two velocities are 18% in the glycerol
solution (60%) and 19% in the glycerol solution (72%) respectively.

34
2.4. Results and discussion

Diethylene glycol Decane


Glycerol (60%) Water
0.25 Glycerol (72%) 0.25
Optical probe

Optical probe
0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05
0.05 0.1 0.15 0.2 0.25 0.05 0.1 0.15 0.2 0.25
HS camera HS camera

(a) Viscous liquids (b) Low viscosity liquids

Figure 2.11: Parity plots of bubble velocities during the interaction.

2.4.4 Analysis of the probe deviation


In order to characterize the performance of the four-point optical fibre probe,
the average deviations between the high speed camera and the probe are
compared in term of the Morton number (Table 2.2). Note that σ1 denotes the
average deviation between the terminal bubble velocity and the velocity de-
termined from the optical probe, σ2 represents the average deviation between
the instantaneous bubble velocity and that from the probe and σ3 represents
the deviation between the average velocity during the interaction from the
high speed camera and that obtained from the optical probe.

Table 2.2: Morton numbers and average deviations.

Morton number σ1 [%] σ2 [%] σ3 [%]


Decane 5.18 × 10−10 18 18 16
Diethylene glycol 8.11 × 10−5 45 40 34
Glycerol (60%) 3.69 × 10−7 29 25 18
Glycerol (72%) 1.64 × 10−5 33 25 19
Water 2.57 × 10−11 20 20 18

Velocities measured in diethylene glycol with the probe have the highest
deviation within these five liquids. Whereas, the deviation measured in

35
2. Single bubble experiment

decane is the lowest one. The order of magnitude of the Reynolds number
is about 103 in water and about 102 in decane. In these cases, the influence
of viscous forces is much smaller than the inertial force acting on the bubble.
Therefore, viscous forces do not dominate in the range of the investigated
Reynolds numbers. As a result the probe has little influence on the motion of
the bubble in these cases. The deviation in water is slightly larger than that
in decane. This could be due to the larger surface tension of water compared
to that of decane. A large surface tension tends to prevent the bubble from
being pierced.
In measurements with viscous liquids, the order of magnitude of the
Reynolds number is about 10. In this situation, viscous forces have a sig-
nificant influence on the motion of the bubble. Therefore, the deviations of
bubble velocity between high speed camera and the optical probe are larger.

2.5 Conclusions

This chapter was dedicated to investigate the performance of a four-point


optical probe with respect to the determination of bubble velocity in a flat
bubble column and the influence of physical properties on the accuracy of the
four-point optical probe. For this purpose, a combined experiment was de-
signed. A CMOS camera was used to verify the performance of the four-point
optical probe. Five liquids with different physical properties were employed.
The motion of the bubble and the interaction with the optical probe were
recorded via the high speed camera. The signals of the probe were acquired
meanwhile. Both terminal bubble velocity and instantaneous bubble velocity
at the moment the bubble hits the central tip of the probe were compared
with that from signal processing of the probe. The bubble shape remains
nearly constant before the bubble approaches the tips of the probe in viscous
liquids, and the bubble only deforms when it gets pierced by the probe. In
low viscosity liquids, however, the bubble starts wobbling and changes its
shape continuously during its rise in the column. Discrepancies of bubble
velocity between image processing and signal processing is more obvious in
viscous liquids.

36
Nomenclature

Physical properties of the liquid, such as viscosity and surface tension,


influence the performance of the four-point optical probe. The Reynolds
number and Morton number were combined to describe the extent of the
discrepancy between the bubble velocity obtained from image processing
and that from the four-point optical probe. In the viscous dominant regime
(low Re), the average deviation is large when the Morton number increases.
When the viscosity does not dominate the motion of the bubble (high Re), the
Morton number is low and the surface tension will influence the accuracy in
the velocity measurement of the probe.
In short, by applying a four-point optical fibre probe in a bubble column,
inaccuracies with respect to bubble velocities cannot be avoided. Particularly,
the discrepancy becomes large when the viscosity of the liquid increases.
However, based on the fact of the stable behavior of bubbles rising in a viscous
liquid, correction for velocity obtained from the four-point optical fibre probe
may be possible.
Furthermore, in low viscosity liquid systems (i.e. high Reynolds flows),
this inaccuracy is lowest. For high void fraction bubbly flow in a bubble
column at low liquid velocities, where PIV or LDA may not work ideally, the
four-point optical probe would also be an appropriate option.

Nomenclature

p fraction of pixels, [-]


P probability, [-]
∆s distance between the central tip and the other three tips, [m]
∆t time difference, [s]
t time of individual bubble traveling through tip of the probe, [s]
u bubble velocity from four-point optical fibre probe, [m s−1 ]
v bubble velocity from image processing, [m s−1 ]
V voltage, [V]
z vertical position of bubble, [m]

37
2. Single bubble experiment

Greek letters

β bubble selection criterion, [-]


µ class mean, [-]
ω class occurence, [-]
σ variance, [-]

Indices

0, 1 class, [-]
B between-class, [-]
T total, [-]
entry probe entering bubble, [-]
exit probe exiting bubble, [-]

References

E. Barrau, N. Rivière, Ch. Poupot, and A. Cartellier. Single and double optical
probes in air-water two-phase flows: Real time signal processing and sensor
performance. International Journal of Multiphase Flow, 25(2):229–256, 1999.

C. Boyer, A. -M. Duquenne, and G. Wild. Measuring techniques in gas-


liquid and gas-liquid-solid reactors. Chemical Engineering Science, 57(16):
3185–3215, 2002.

D. Bröder and M. Sommerfeld. An advanced LIF-PLV system for analysing


the hydrodynamics in a laboratory bubble column at higher void fractions.
Experiments in Fluids, 33(6):826–837, 2002.

A. Cartellier. Simultaneous void fraction measurement, bubble velocity, and


size estimate using a single optical probe in gas-liquid two-phase flows.
Review of Scientific Instruments, 63(11):5442–5453, 1992.

N. G. Deen. An experimental and computational study of fluid dynamics in gas-


liquid chemical reactors. PhD thesis, Aalborg University, Denmark, 2001.

38
References

R. Fortunati, S. Guet, G. Ooms, R. V. A. Oliemans, and R. F. Mudde. Accuracy


and feasibility of bubble dynamic measurement with four point optical fiber
probes. Lisbon, July 2002. 11th symposium on application of laser technique
to fluid mechanics.

S. Guet, R. V. Fortunati, R. F. Mudde, and G. Ooms. Bubble velocity and


size measurement with a four-point optical fiber probe. Particle and Particle
Systems Characterization, 20(3):219–230, 2003.

W. K. Harteveld. Bubble columns: Structure or Stability? PhD thesis, Delft


University of Technology, The Netherlands, 2005.

O. C. Jones and J. -M. Delhaye. Transient and statistical measurement tech-


niques for two-phase flows: A critical review. International Journal of Multi-
phase Flow, 3(2):89–116, 1976.

R. Lindken, L. Gui, and W. Merzkirch. Velocity measurements in multiphase


flow by means of particle image velocimetry. Chemical Engineering and
Technology, 22(3):202–206, 1999.

S. Luther, J. Rensen, and S. Guet. Bubble aspect ratio and velocity measure-
ment using a four-point fiber-optical probe. Experiments in Fluids, 36(2):
326–333, 2004.

R. F. Mudde and T. Saito. Hydrodynamical similarities between bubble col-


umn and bubbly pipe flow. Journal of Fluid Mechanics, 437:203–228, 2001.

N. Otsu. Threshold selection method from gray-level histograms. IEEE Trans


Syst Man Cybern, SMC-9(1):62–66, 1979.

M. Sezgin and B. Sankur. Survey over image thresholding techniques and


quantitative performance evaluation. Journal of Electronic Imaging, 13(1):
146–168, 2004.

39
Chapter
3
Bubble properties of heterogeneous
bubbly flows in bubble column
This chapter focuses on the measurements of bubble properties in heterogeneous bubbly
flows in a square bubble column. A four-point optical fibre probe was used for this
purpose. The accuracy and intrusive effect of the optical probe was investigated first.
The results show that the optical probe underestimates bubble properties, such as,
bubble velocity and local void fraction. The presence of the probe in the bubble column
influences the local flow conditions. Particularly, when the probe is placed close to the
liquid surface, this influence is more pronounced. Furthermore, two methods for the
determination of the interfacial area were compared. The results from both methods
agree quite well at low superficial gas velocity, whereas significant discrepancies were
obtained at high superficial gas velocity. Finally, the effect of (initial) liquid height
on bubble properties was studied. No significant difference was found for the three
investigated (initial) liquid heights.

41
3. Bubble properties of heterogeneous bubbly flows

3.1 Introduction

Bubble column reactors are intensively utilized as multiphase contactors and


reactors in chemical, petrochemical, biochemical and metallurgical industries.
Typically these reactors have a rather simple geometry: a container filled with
liquid where gas is introduced via a sparger at the bottom in the form of
bubbles which comes in contact with the liquid. Simple construction and
lack of any mechanically operated parts are two characteristic aspects of these
reactors. Hence, little maintenance is required at low operating costs.
Bubble column reactors have excellent heat and mass transfer character-
istics. In gas-liquid reactors, mass transfer between the gas phase and liquid
phase constitutes an important goal of the process. The volumetric mass
transfer coefficient is a key parameter in the characterization and design of
industrial reactors. Moreover, a high heat transfer coefficient is of special
significance when reactions are involved with large heat effects.
Although the construction of the bubble column reactor is simple, accu-
rate and successful design and scale-up of bubble column reactors requires
detailed understanding of multiphase fluid dynamics and its impact on trans-
port of mass and/or heat. Bubble characteristics, such as bubble size, void
fraction and rise velocity, have significant impact on the hydrodynamics, as
well as on heat and mass transfer coefficients in a bubble column reactor.
Optical fibre probes (Cartellier, 1990, 1992; Mudde and Saito, 2001) offer the
advantage of low cost, simplicity of setup and relative ease of interpretation
of the probe signals. With a single optical fibre probe, the measurement of the
local void fraction can be conducted. Furthermore, using a four-point optical
fibre probe, bubble characteristics, such as bubble velocity and bubble size,
can also be obtained. Moreover, it is also possible to determine the specific
interfacial area using a four-point optical fibre probe.
In this chapter, bubble properties in a square bubble column were mea-
sured with a four-point optical fibre probe over a relatively large range of
superficial gas velocities (0.005 m/s − 0.1 m/s). The flow regime changes
from homogeneous to heterogeneous. Measurements were taken using three
columns with different (initial) liquid heights (H/W = 3, 6 and 9).

42
3.2. Experimental setup

3.2 Experimental setup

The experiments were carried out in a square column with a cross-sectional


area (W × D) of 0.15 × 0.15 m2 . A perforated plate with 49 holes of 1 mm in
diameter was used as a gas sparger at the bottom of the column. These holes
were distributed uniformly at a square pitch of 6.25 mm in the central region
of the perforated plate. Demineralized water was used as liquid phase and
air as the gas phase.
A sketch of the experimental setup is shown in Figure 3.1. A four-point
optical probe was inserted through one of the holes in the side walls. The
optical probe could be moved in the horizontal direction to obtain measure-
ments at different lateral positions (at fixed height). A differential pressure
transducer (DP) was used to measure the differential pressure in the column.

Figure 3.1: Sketch of experimental setup.

During the measurements, three different liquid heights (H/W = 3, 6and9)


were investigated. The probe was moved in the middle plane horizontally.

43
3. Bubble properties of heterogeneous bubbly flows

For each measurement a sampling time exceeding thirty minutes was taken.

3.3 Data processing

3.3.1 Interpretation of the probe


Interpretation of signals to measure bubble velocities from the probe has been
introduced in Chapter 2. In addition, the local void fraction can be extracted
from the signals of the probe. It is expressed as the ratio between accumulated
P
time that the probe tip is inside individual bubbles t and the total sampling
time T (Cartellier, 1990): P
t
α= (3.1)
T
As reported in the previous chapter, the bubble velocity u is determined
from the distance between the central tip and the other three tips and the
average time difference between the instants that the bubble hits the central
tip and the other three tips. Once the bubble velocity u and the time span for
the bubble travelling through the optical probe t are known, it is possible to
calculate the chord length of the bubble dc .

dc = ut (3.2)

Due to the fact that many bubbles can be detected at each position within
the sampling period, many bubble velocities and chord lengths will be ob-
tained at that position for one measurement. For instance, in a small col-
umn (H/W = 3), the probability density functions of bubble velocity and
chord length at position x/W = 0.5, z/H = 0.75 for superficial gas velocities
u g = 0.005 m/s and u g = 0.04 m/s are respectively shown in Figure 3.2.
In order to obtain the average bubble properties at a certain position, the
probability density functions, such as f (u) and f (dc ), are used to determine
the corresponding mean quantities:
Z ∞
u= u f (u) du (3.3)
0
Z ∞
dc = dc f (dc ) ddc (3.4)
0

44
3.3. Data processing

0.05 0.02

0.04
0.015

0.03

f(dc)
f(u)

0.01
0.02

0.005
0.01

0 0
0 0.5 1 1.5 0 0.002 0.004 0.006 0.008 0.01 0.012
u [m/s] dc [m]

(a) Bubble velocity (u g = 0.005 m/s) (b) Chord length (u g = 0.005 m/s)
0.2 0.25

0.2
0.15

0.15
f(dc)
f(u)

0.1
0.1

0.05
0.05

0 0
0 1 2 3 4 5 0 0.05 0.1
u [m/s] dc [m]

(c) Bubble velocity (u g = 0.04 m/s) (d) Chord length (u g = 0.04 m/s)

Figure 3.2: Probability density functions.

The selection criterion has a strong effect on the fraction of “accepted“


bubbles and on the accuracy of the results. As discussed in Chapter 2, the
probability density function of the bubble velocity does not change consid-
erably by increasing β up to 25%. In this chapter, in order to obtain more
”accepted” bubbles, a small test was done before adopting the selection crite-
rion. When increasing β from 25% to 40%, the percentage of accepted bubbles
increases from about 20% to 37% accordingly. However, the resulting devia-
tion of the mean bubble velocity and the mean bubble chord length is about
2% and 1% respectively, which is acceptable for the present study.
During the measurements, at least 1500 bubbles selected from the criterion

45
3. Bubble properties of heterogeneous bubbly flows

mentioned above were used to calculate the average bubble velocity at each
location. Particularly, at high superficial gas velocity, more than 3000 bubbles
were selected, i.e. above 10 000 bubbles were used in case of superficial gas
velocity u g = 0.6 m/s.

3.3.2 Differential pressure transducer


By measuring the differential pressure ∆P over a certain height ∆z before and
after injecting gas into column, the integral gas holdup  can be obtained from
the following equation:
∆P
=1− (3.5)
ρl g∆z
It is assumed that the density of the gas phase is negligible compared to that
of the liquid phase.

3.3.3 Determination of bubble size


Bubble sizes cannot be determined from the optical probe directly since the
optical probe does not always intersect a bubble at its center. That is, a
chord length smaller than the largest vertical bubble dimension is typically
measured. In literature, some studies have been carried out to estimate the
equivalent bubble size from the chord length distribution (Uga, 1972; Herringe
and Davis, 1976; Thang and Davis, 1979). Based on some assumptions, such
as i) all bubbles are spherical and ii) all bubbles travel in the same direction
with the same average velocity, the equivalent bubble diameter de can be
determined with the aid of geometrical probability analysis:

3
de = dc (3.6)
2

where dc is the mean chord length.


Some studies also focused on inverse transformation from the chord length
distribution to bubble size distribution (Turton and Clark, 1988; Liu and Clark,
1995; Liu et al., 1998) where an ellipsoidal or truncated ellipsoidal bubble
shape were under assumed. By assuming that bubbles rise vertically and

46
3.3. Data processing

with a zero angle of attack, the mean horizontal diameter of a bubble with
aspect ratio E (Liu and Clark, 1995) can be obtained from:
3 −1
dH = E dc (3.7)
2
Therefore, the equivalent diameter of the bubble is:
3 −2/3
de = E dc (3.8)
2
Once the aspect ratio of the bubble E is known, one is able to determine the
equivalent diameter of the bubble. However, it may not be easy to determine
the aspect ratio in high void fraction flows. Due to the phenomena of coa-
lescence and breakup of bubbles, the bubble size varies and thus, the bubble
shape varies. Hence, assumptions with respect to the shape of the bubbles
may lead to (considerable) inaccuracies in bubble size.
The local specific interfacial area is related to the local void fraction and
the local equivalent bubble size:

de = (3.9)
a
where α is the local void fraction.
This provides an alternative means of estimating the equivalent bubble
size if the specific interfacial area is known. In the next section, we will
introduce techniques for measuring the specific interfacial area using a four-
point optical fibre probe.

3.3.4 Specific interfacial area


Delhaye and Bricard (1994) pointed out that the structure of a bubbly two-
phase flow can be characterized by two out of the three following parameters:
the interfacial area, the void fraction, and the Sauter mean diameter. In addi-
tion, in two-fluid models for two-phase flows, transport of mass, momentum
and energy across the phase boundaries is represented by interfacial transfer
terms. In general, all interfacial transfer rate can be expressed as the product
of a driving force and the inverse of a length scale L at the interface (Ishii,
1975):

47
3. Bubble properties of heterogeneous bubbly flows

Interfacial transfer rate = 1


L × driving force

The driving forces are characterized by local transport mechanisms such as


molecular and turbulent diffusion, whereas the term 1/L represents the time-
average of the interfacial area per unit of volume, i.e. the specific interfacial
area a, and is related to the structure of the two-phase.
Methods of measuring local specific interfacial area have been proposed
for both a two-point and four-point probe (Kataoka et al., 1986; Tan and Ishii,
1990; Revankar and Ishii, 1993; Shen et al., 2005). For a detailed study of the
interfacial area definition we refer to Ishii (1975). A brief description of the
method for a four-point probe is as follows (Tan and Ishii, 1990).
The time-average of the specific interfacial area at certain position x0 is
given by:
1X
N
1
a= (3.10)
T
j=1 v0j · n j

where N is the number of times over the sampling time T that an interface
passes through x0 where v0 and n represent the interface velocity and unit
normal vector of the interface respectively at x0 .
Let f (x, t) = 0 represent an interface. The jth interface passes through
x0 at time t = t0 j and can be represented as where f j (x0 , t0j ) is assumed to
sufficiently differentiable, f j (x0 , t0j ) = 0.

N
1 X O f j (x0 , t0j )
a= (3.11)
T ∂f
j=1
∂t (x0 , t0 j )

Suppose that the interface passes through three adjacent positions x1 , x2


and x3 at times t1 , t2 and t3 respectively. That is, f j (xk , tk j ) = 0, k = 1, 2, 3.
If the distances sk = |xk − x0 |, k = 1, 2, 3 and the time differences ∆tk j = tk j −
t0j , k = 1, 2, 3 are small in comparison with the length scale and the time scale,
respectively, of the system under consideration, then each of f j (x j , tk j ), k =
1, 2, 3 can be written as a Taylor series expansion about x = x0 and t = t0 j :

∂f
f j (xk , tk j ) = f j (x0 , t0 j ) + sk O f j (x0 , t0 j ) · ξk + ∆tk j (x0 , t0j ) k = 1, 2, 3 (3.12)
∂t

48
3.3. Data processing

where O f j (x0 , t0 j ) · ξk denotes the directional derivative of f j in the direction of


the unit vector ξk which is parallel to the line passing through x0 and xk .
Assume the unit vector ξk , k = 1, 2, 3 is linearly independent and in terms
of the Cartesian components ξkx , ξky and ξkz of ξk the direction cosines cosα j ,
cosβ j and cosγ j of nj can be obtained:

∂ fj
(x0 , t0 ) A
cosα j = − ∂t
1j
(3.13)
O f j (x0 , t0 j ) A0

∂ fj
(x0 , t0 ) A
cosβ j = − ∂t
2j
(3.14)
O f j (x0 , t0 j ) A0

and
∂ fj
(x0 , t0 ) A
cosγ j = − ∂t
3j
(3.15)
O f j (x0 , t0j ) A0

where

ξ1x ξ1y ξ1z

A0 = ξ2x ξ2y ξ2z (3.16)

ξ3x ξ3y ξ3z


∆t1 j ξ1y ξ1z
s1
∆t2 j
A1 j = ξ2y ξ2z (3.17)
s2
∆t3 j
s ξ3y ξ3z
3


ξ ∆t1j
1x ξ1z

s1
∆t2j
A2j = ξ2x ξ2z (3.18)
s2
∆t3j
ξ3x ξ3z
s3

49
3. Bubble properties of heterogeneous bubbly flows

and
ξ ∆t1j
1x ξ1y
s1
∆t2j
A3j = ξ2x ξ2y (3.19)
s2
∆t3j
ξ3x ξ3y
s3
It follows that the identity cos2 α + cos2 β + cos2 γ = 1 and from equation 3.11
that a is given by:
1 Xq 2
N
a= A1j + A22j + A23 j (3.20)
T|A0 |
j=1

The above method makes it possible to measure the specific interfacial


area with the four-point optical fibre probe with minor (severe) assumptions.

3.4 Results and discussion

3.4.1 Intrusive effect of the probe


There is no doubt that the optical probe, as an intrusive instrument, locally
disturbs the flow field and thus, the bubble properties obtained from the
probe. To investigate this effect and verify the feasibility of using the probe,
an extra measurement was performed. The bubble properties at three different
depths (y/D = 0.5, 0.75, and 0.875) were measured with the four-point optical
probe. The small column (H/W = 3) was used. The optical probe was inserted
through the side wall at nearly half height (z/H = 0.55) of the column. Two
cases with different superficial gas velocities (u g = 0.02 m/s and u g = 0.1 m/s)
were considered.
When the optical probe is used to obtain values for local void fraction
α and mean bubble velocity u at position (x, y) in one horizontal plane, the
overall gas flow rate Q can be computed with the following formula and thus,
compared with that from the flow meter to verify the accuracy of the optical
probe.
Z
Q= α(x, y)u(x, y) dA (3.21)
A

50
3.4. Results and discussion

where A is cross-sectional area of the column.


By assuming symmetry of the flow, the total gas flow rate is obtained from
calculating the gas flow rate over half the area of the cross section and/or a
quarter of the cross sectional area.
Furthermore, the surface-averaged void fraction is calculated as follows:

Z
1
α= α(x, y) dA (3.22)
A A

The surface-averaged void fractions are compared with the gas holdups
obtained from the measurements using the differential pressure transducer.
The comparison for both gas flow rate and gas holdup is given in Table 3.1.

Table 3.1: Comparisons of both gas flow rate and gas holdup.

u g = 0.02 m/s u g = 0.1 m/s


Method
Q [m /s]
3
α [-] Q [m3 /s] α [-]
−4 −3
Measured 4.49 × 10 0.026 2.25 × 10 0.11
Probe, A/2 2.98 × 10−4 0.023 1.53 × 10−3 0.08
Probe, A/4 3.13 × 10−4 0.024 1.58 × 10−3 0.09

It can be seen that the gas flow rates obtained from the optical probe
are less than those from the flow meter. The gas flow rate from the optical
probe differs about 30% from that obtained with the flow meter. The surface-
averaged void fraction from the optical probe has a much smaller deviation
compared with results from the differential pressure transducer.
The possible reasons for the observed discrepancies may be various. Be-
sides the assumption of symmetry of flows inside column, the processing
method of the optical probe rules out many bubbles that are rising inside the
column at a large angle of attack and that do not hit the four tips of the probe.
Furthermore, the intrusive effect of the probe is also one of the reasons. In
spite of the discrepancies, the optical fibre probe can still be used in two-phase
bubbly flows to provide meaningful data.

51
3. Bubble properties of heterogeneous bubbly flows

3.4.2 Bubble size

The bubble size is calculated on basis of the two methods described in sec-
tion 3.3.3. For the purpose of simplicity, equation 3.6 is used to calculate
bubble size from the chord length distribution (method A), since the bubble
shape is difficult to determine. Furthermore, equation 3.9 is used to determine
the bubble size from the measurement of the specific interfacial area (equa-
tion 3.20) (method B). Comparisons are made in a tall column (H/W = 6) with
four different superficial gas velocities (u g = 0.005, 0.04, 0.06, 0.08and0.1 m/s).
The results are shown in Figure 3.3.

0.02
ug =0.005m/s, Eqn. 3.6 ug=0.005m/s, Eqn. 3.9
ug=0.04m/s, Eqn. 3.6 ug=0.04m/s, Eqn. 3.9
ug=0.06m/s, Eqn. 3.6 ug=0.06m/s, Eqn. 3.9
ug=0.08m/s, Eqn. 3.6 ug=0.08m/s, Eqn. 3.9
ug=0.1m/s, Eqn. 3.6 ug=0.1m/s, Eqn. 3.9
0.015
de [m]

0.01

0 0.2 0.4 0.6 0.8 1


x/W [-]

Figure 3.3: Bubble size distribution (z/H = 0.68, H/W = 6).

It can be seen that the method A agrees well with the method B at low
superficial gas velocity (u g = 0.005 m/s). However, discrepancies arise. The
bubble sizes from method A are much larger than those from method B. That
is reasonable, since the shape of the bubble may no longer be spherical due to
strong coalescence of bubbles. Therefore, method B will be used for further
discussions.

52
3.4. Results and discussion

3.4.3 Effect of initial liquid height


The effect of initial liquid height on the bubble properties, such as bubble
velocity, local void fraction, interfacial area and equivalent diameter, will now
be discussed. Measurements were taken at height z = 0.34 m in the middle
plane (y/D = 0.5) for three liquid heights (H/W = 3, 6 and 9) at superficial gas
velocity u g = 0.005 m/s.

(a) Mean bubble velocity [m/s] (b) Local void fraction [-]

(c) Specific interfacial area [1/m] (d) Equivalent bubble size [m]

Figure 3.4: Distributions of bubble properties (u g = 0.005 m/s).

From Figure 3.4, one can find that there are only small differences in
the bubble properties among these three liquid heights. In Figure 3.5, bub-
ble properties at higher superficial gas velocity are compared for two liquid
heights (H/W = 3 and 6).
One can see that the bubble properties show little difference. Hence,

53
3. Bubble properties of heterogeneous bubbly flows

(a) Mean bubble velocity [m/s] (b) Local void fraction [-]

(c) Specific interfacial area [1/m] (d) Equivalent bubble size [m]

Figure 3.5: Distributions of bubble properties at higher superficial gas veloc-


ities.

one may conclude that the initial liquid height barely influences the bubble
properties. However, it is more obvious that the presence of the optical probe
in column disturbs the local flow. The optical probe is inserted through the
wall at the right side of the column. The bubble properties at the right half part
of the distributions, such as the mean bubble velocity, the local void fraction
and the specific interfacial area, are smaller than those at the left half part of
the distributions. The closer to the side wall the measurement point is, the
larger the difference is.
In addition, it can be seen that the profile of the mean bubble velocity at
superficial gas velocity u g = 0.005 m/s is flat (see Figure 3.4). That is due to
meandering behavior of the bubble plume in the column. However, bubbles
have more probability to rise in the middle of the column than at the sides,
which can be seen from the profiles of the local void fraction. The local

54
3.4. Results and discussion

void fraction has a peak in the middle of the column. Furthermore, one can
also observe that the specific interfacial area has peak in the middle. The
equivalent bubble size is uniform throughout the horizontal direction in the
column.
As the superficial gas velocity increases, the meandering motion of the
bubble plume changes to a circulation pattern. Most of the bubbles rise in
the middle of the column and some move downwards at the wall region
due to liquid circulation. The profiles of the mean bubble velocity become
parabolic. Moreover the mean bubble velocity, the local void fraction and the
specific interfacial area increase accordingly. For instance, the mean bubble
velocity in the center increases from about 0.43 m/s at a superficial gas velocity
u g = 0.005 m/s to around 0.90 m/s at a superficial gas velocity u g = 0.1 m/s.
The local void fraction in the center increases about ten times (from 2.3% to
around 23%). The specific interfacial area increases from around 23 to 170.
However, the equivalent bubble diameter does not increase so much. The
equivalent bubble diameter ranges from 0.0064 m at a superficial gas velocity
u g = 0.005 m/s while the equivalent bubble diameter becomes more or less
constant (around 0.009 m) at higher superficial gas velocity.

3.4.4 Axial (or vertical) evolution of bubble properties


In this section, we will report the evolution of bubble properties along column
height. Bubble properties in the center of the column (x/W = 0.5, y/D = 0.5)
are considered. For the small liquid height (H/W = 3), the bubble properties
are taken at three heights (z/H = 0.35, 0.55 and 0.75). For the relatively large
liquid height (H/W = 6), three heights (z/H = 0.38, 0.68 and 0.88) were taken.

In Figure 3.6, one can find that the bubble properties, such as the local void
fraction and the specific interfacial area, remain constant along the height at
superficial gas velocity u g = 0.005 m/s. The mean bubble velocity and the
equivalent bubble diameter possess a little variation along the vertical co-
ordinate. At higher the superficial gas velocity, the mean bubble velocity
decreases slightly along the column height which may be due to the spread-
ing of the bubble plume over the cross sectional area. The local void fraction

55
3. Bubble properties of heterogeneous bubbly flows

(a) Mean bubble velocity [m/s] (b) Local void fraction [-]

(c) Specific interfacial area [1/m] (d) Mean bubble size [m]
Figure 3.6: Axial (or vertical) bubble properties (H/W = 3).

and the specific interfacial area first decrease and then increase slightly. Fur-
thermore, the equivalent bubble size increases a little along the height up to
half the height of the column where the increase slows down in the upper
part of the column. One may notice that the equivalent bubble size increases
with the superficial gas velocity up to u g = 0.06 m/s and then, decreases with
the superficial gas velocity.
The axial evolution of the bubble properties at large initial liquid height
(H/W = 6) are plotted in Figure 3.7. It can be seen that the mean bubble
velocity decreases slightly along the height for all superficial gas velocities.
For u g = 0.005 m/s, the local void fraction remains constant and the specific
interfacial area increases slightly. As a result, the equivalent bubble diameter
shows a small decrease in the upper part of the column. Beyond that super-

56
3.5. Conclusions

(a) Mean bubble velocity [m/s] (b) Local void fraction [-]

(c) Specific interfacial area [1/m] (d) Mean bubble size [m]

Figure 3.7: Axial (or vertical) evolution bubble properties (H/W = 6).

ficial gas velocity, the local void fraction and the specific interfacial area first
decrease and then increase slightly. There are similar observations with those
found in the small column. The increase of the equivalent bubble diameter
slows down along the column height.

3.5 Conclusions

In the present work, bubble properties in bubble columns with different initial
liquid heights have been investigated experimentally. In addition the effect of
the superficial gas velocity has been investigated. A four-point optical fibre
probe was used to measure the relevant quantities, such as bubble velocity,
local void fraction, chord length and specific interfacial area. Bubble size

57
3. Bubble properties of heterogeneous bubbly flows

has been determined by using two different methods. One method is based
on an inverse transform of the chord length distribution using geometrical
probability theory. The equivalent bubble diameter is then related to the mean
chord length. Another method is to determine the equivalent bubble size from
the specific interfacial area which is deduced from the data processing of the
four-point optical probe. Comparison of the two methods reveals that results
from the two methods agree with each other quite well at low superficial gas
velocity. However, differences become apparent as the superficial gas velocity
increases. The former method assumes that bubbles have a spherical shape.
That may be true at low superficial gas velocity due to lack of coalescence of
bubbles. The shape of bubble deviates from spherical when the bubble size
increases and deforms. The assumption of spherical bubbles is not valid any
more. Therefore, the latter method was used for further analysis.
The accuracy of the four-point optical probe was also studied. Mean
bubble velocities and local void fractions at three different depths were used
to calculate area-averaged properties, such as gas flow rate and average void
fraction. The area-averaged gas flow rate was compared with that from a
flow meter whereas the area-averaged void fraction was compared with the
gas holdup obtained from differential pressure measurements. Discrepancies
were found for both methods. Possible reasons for the observed discrepancies
may be due to lack of symmetry of flow inside the column, the processing
method of the optical probe and the intrusive effect of the optical probe.
In addition, the effect of initial liquid height on the bubble properties was
investigated. Bubble properties were compared at three (initial) liquid heights
(H/W = 3, 6 and 9) at a superficial gas velocity u g = 0.005 m/s and for two
liquid heights (H/W = 3 and 6) at higher superficial gas velocities. Results
show that the initial liquid height barely influences the bubble properties at
the same height. However, the axial evolution of bubble properties differs
between the systems with different initial liquid heights. Particularly, the
mean bubble velocity decreases continuously along the column height at
small liquid height while it keeps nearly constant at higher superficial gas
velocities.

58
Nomenclature

Nomenclature

a specific interfacial area, [m−1 ]


A cross-sectional area, [m2 ]
D column depth, [m]
d bubble size, [m]
dc chord length, [m]
dc mean chord length, [m]
de equivalent bubble size, [m]
E aspect ratio of bubble, [-]
f function, [-]
g gravitational acceleration, [m s−2 ]
H column height, [m]
L length scale, [m]
n unit normal vector of interface, [-]
N number of times that an interface passes through a certain posi-
tion, [-]
∆P differential pressure, [N m−2 ]
Q volume flow rate, [m3 s−1 ]
t time; time of individual bubble traveling through tip of the probe,
[s]
T sampling time, [s]
u bubble velocity, [m s−1 ]
ug superficial gas velocity, [m s−1 ]
v interface velocity vector, [m s−1 ]
W column width, [m]
x distance in x direction, [m]
y distance in y direction, [m]
z distance in z direction, [m]

Greek letters

α local void fraction, [-]

59
3. Bubble properties of heterogeneous bubbly flows

α surface-averaged void fraction, [-]


β Bubble selection criterion, [-]
 gas holdup, [-]
ρ density, [kg m−3 ]
ξ unit vector, [-]

Indices

0−3 probe tip


H horizontal
j interface
l liquid

References

A. Cartellier. Optical probes for local void fraction measurements: Charac-


terization of performance. Review of Scientific Instruments, 61(2):874–886,
1990.

A. Cartellier. Simultaneous void fraction measurement, bubble velocity, and


size estimate using a single optical probe in gas-liquid two-phase flows.
Review of Scientific Instruments, 63(11):5442–5453, 1992.

J. M. Delhaye and P. Bricard. Interfacial area in bubbly flow: Experimental


data and correlations. Nuclear Engineering and Design, 151(1):65–77, 1994.

R. A. Herringe and M. R. Davis. Structural development of gas-liquid mixture


flows. Journal of Fluid Mechanics Digital Archive, 73(01):97–123, 1976.

M. Ishii. Thermo-fluid dynamic theory of two-phase flow. Collection de la Direction


des Études et Recherches d’Électricité de France, ISSN 0399-4198; 22. [Paris]
: Eyrolles, 1975.

I. Kataoka, M. Ishii, and A. Serizawa. Local formulation and measurements


of interfacial area concentration in two-phase flow. International Journal of
Multiphase Flow, 12(4):505–529, 1986.

60
References

W. Liu and N. N. Clark. Relationships between distributions of chord lengths


and distributions of bubble sizes including their statistical parameters. In-
ternational Journal of Multiphase Flow, 21(6):1073–1089, 1995.

W. Liu, N. N. Clark, and A. I. Karamavruç. Relationship between bubble size


distributions and chord-length distribution in heterogeneously bubbling
systems. Chemical Engineering Science, 53(6):1267–1276, 1998.

R. F. Mudde and T. Saito. Hydrodynamical similarities between bubble col-


umn and bubbly pipe flow. Journal of Fluid Mechanics, 437:203–228, 2001.

S. T. Revankar and M. Ishii. Theory and measurement of local interfacial area


using a four sensor probe in two-phase flow. International Journal of Heat
and Mass Transfer, 36(12):2997–3007, 1993.

X. Shen, Y. Saito, K. Mishima, and H. Nakamura. Methodological improve-


ment of an intrusive four-sensor probe for the multi-dimensional two-phase
flow measurement. International Journal of Multiphase Flow, 31(5):593–617,
2005.

M. J. Tan and M. Ishii. A method for measurement of local specific interfacial


area. International Journal of Multiphase Flow, 16(2):353–358, 1990.

N. T. Thang and M. R. Davis. The structure of bubbly flow through venturis.


International Journal of Multiphase Flow, 5(1):17–37, 1979.

N. N. Turton and R. Clark. Chord length distributions related to bubble size


distributions in multiphase flows. International Journal of Multiphase Flow,
14(4):413–424, 1988.

T. Uga. Determination of bubble-size distribution in a BWR. Nuclear Engineer-


ing and Design, 22(2):252–261, 1972.

61
Chapter
4
Numerical analysis of the effect of
gas sparging on bubble column
hydrodynamics
A discrete bubble model (DBM) has been used to study the effect of gas sparger prop-
erties on the hydrodynamics in a bubble column. As a first step the performance of
the model was evaluated by comparison with experimental data. Subsequently, four
different perforated plates with different sparged areas were used as a gas sparger.
Distributions of liquid velocity, turbulent kinetic energy and void fraction in the cen-
tral plane were compared for the four different systems. Furthermore, the effect of the
sparger location was also investigated. It was found that the liquid phase circulation
becomes more pronounced as the sparged area location is more distant from the center
of the bottom plate.
Finally, gas phase Residence Time Distributions (RTD) were obtained from the sim-
ulations. By employing standard axial dispersion model, the gas phase mixing in the
bubble column was characterized. Results show that the extent of mixing increased
when the sparged area decreased. The axial dispersion coefficient increased as the
sparged area was shifted to the edge of the bottom plate.

63
4. Effect of gas sparging on bubble column hydrodynamics

4.1 Introduction

Bubble column reactors are intensively utilized in chemical, petrochemical,


biochemical and metallurgical industries. These reactors are often preferred
because of simplicity of operation, low operating costs and ease with which the
liquid residence time can be varied. This type of gas-liquid contactor has been
studied intensively during the past decades. Parameters, such as gas holdup,
gas-liquid interfacial area, interfacial mass transfer coefficients and dispersion
coefficients and heat transfer coefficient and so on, have been investigated for
the purpose of scale-up and design of bubble column reactors. Furthermore,
the influence of the gas sparger on the hydrodynamical parameters has been
indicated by Shah et al. (1982).
Hebrard and Bastoul (1996) studied flow regime transition, critical gas
flow rates corresponding to the transition from one regime to the other, gas
holdup, bubble size distribution and axial liquid dispersion coefficients in
columns with three types of gas spargers: a perforated plate, a porous sin-
tered plate and a flexible membrane disc. Their results revealed that the hy-
drodynamical parameters strongly depend on the type of gas sparger used.
Thorat (1998) and Veera and Joshi (1999) studied the combined effect of gas
sparger and height on the gas holdup by using perforated plates. They re-
ported different gas holdup profiles from single point spargers and multipoint
spargers. Bhole et al. (2006) measured the hydrodynamics of a bubble column
with two different gas spargers using LDA. Their results revealed that the
porous plate sparger and the perforated plate sparger result in different col-
umn hydrodynamics. To summarize, the effect of distributor characteristics
on bubble column hydrodynamics, i.e. the details about the distributions of
liquid velocity, turbulent kinetic energy and local void fraction, is important
for reactor design. However, the effect of sparged area of the perforated plate,
such as the size of the sparged area, location of the sparged area and so on, is
interesting as well and will be examined in this chapter.
Recently, the reaction engineering community has been active in explor-
ing the possibilities to utilize Computational Fluid Dynamics (CFD) in the
modeling of multiphase reactors. Eulerian-Lagrangian models are one of the

64
4.2. Discrete bubble model

available Computational Fluid Dynamics tools in the field of bubble column


modeling. Delnoij et al. (1997) used the Eulerian-Lagrangian approach to
calculate the time-dependent two-dimensional motion of small, spherical gas
bubbles in a bubble column operating in the homogeneous regime. Delnoij
et al. (1999) extended their two-dimensional Eulerian-Lagrangian model to a
three-dimensional model. Sommerfeld et al. (2003) developed a bubble coa-
lescence model on the basis of the Lagrangian approach, where bubbles are
traced through the turbulent flow field along their trajectories. Darmana et al.
(2006) applied a parallel algorithm to the Eulerian-Lagrangian model em-
bedding four-way coupling. Hu and Celik (2008) used Eulerian-Lagrangian
approach to study gas-liquid bubbly flow in a flat bubble column by means
of large-eddy simulation (LES) with two-way coupling. Due to additional
efforts in detailed studies of bubbly flows, the Eulerian-Lagrangian approach
has been extended recently to study liquid and gas phase mixing in bubble
columns.
The present work studies the effect of the gas sparger on the hydrody-
namics and gas phase mixing of a square bubble column using several types
of gas spargers (perforated plates) employing the Eulerian-Lagrangian ap-
proach. In addition the influence of the size and the location of sparger area is
investigated. Time-averaged quantities, such as liquid velocity, void fraction
and turbulent kinetic energy, are considered. Furthermore, the residence time
distribution of the gas phase obtained from the Eulerian-Lagrangian model
is used to characterize the gas phase mixing in the bubble column in terms of
an axial dispersion model.

4.2 Discrete bubble model

Our discrete bubble model (DBM) was originally developed by Delnoij et al.
(1997) and Delnoij et al. (1999). The liquid phase hydrodynamics is repre-
sented by volume-averaged continuity and Navier-Stokes equations, while
the motion of each individual bubble is tracked in a Lagrangian way.

65
4. Effect of gas sparging on bubble column hydrodynamics

4.2.1 Bubble dynamics


The motion of each individual bubble is computed from Newton’s second
law. The interaction with liquid phase is accounted for by an interphase
momentum transfer term. For an individual bubble, the equation of motion
and the bubble trajectory equation can be respectively written as:

dv
ρg V = ΣF (4.1)
dt

dr
=v (4.2)
dt
The net force acting on each individual bubble is calculated by considering
all the relevant forces. It is assumed that the net force is composed of separate,
uncoupled contributions due to gravity, pressure, drag, lift, virtual mass and
wall force respectively:

ΣF = FG + FP + FD + FL + FVM + FW (4.3)

The gravity force acting on a bubble in a liquid is given by:

FG = ρ g Vg (4.4)

The far field pressure force incorporating contributions of the Archimedes


buoyancy force, inertial forces and viscous strain is given by:

FP = −V∇P (4.5)

The drag force exerted on a bubble rising in a liquid is expressed as:

1
FD = − CD ρl πd2 |v − u|(v − u) (4.6)
8
where the drag coefficient is taken from Tomiyama et al. (1998):
   
16 48 8 Eo
CD = max min (1 + 0.15Re0.687 ), , (4.7)
Re Re 3 Eo + 4
where Re and Eo are the bubble Reynolds number and the Eötvös number
respectively.

66
4.2. Discrete bubble model

A bubble rising in a non-uniform flow field experiences a lift force due to


vorticity or shear. The shear induced lift force acting on a bubble is usually
written as (Auton, 1987):

FL = −CL ρl V(v − u) × ∇ × u (4.8)

Zhang et al. (2006) studied the influence of the lift coefficient on the bubble
column dynamics using a two-fluid model. It was found that the results with
CL = 0.5 fit best with PIV experimental data. Therefore, CL = 0.5 is used in
the present study as well.
Accelerating bubbles experience a resistance, which is termed the virtual
mass force (Auton, 1987):
 
Dv Du
FVM = −CVM ρl V − (4.9)
Dt Dt
where the D/Dt operators denote the material derivatives pertaining to the
respective phase. In the present work, bubbles are assumed to have a spherical
shape and a virtual mass coefficient of CVM = 0.5 is used.
Bubbles in the vicinity of a solid wall experience a force referred to as the
wall force (Tomiyama et al., 1995):
" #
1 1 1
FW = − C W d 2 − ρl |(v − u)· nz |2 nW (4.10)
2 y (L − y)2

where nz and nW , respectively, are the normal unit vectors in the vertical
and wall normal direction, L is the dimension of the system in the normal
direction, and y is the distance to the wall in that direction. Finally, the wall
force coefficient CW is given by:



exp(−0.933Eo + 0.179) 1 ≤ Eo ≤ 5,
CW =
 (4.11)
0.007Eo + 0.04 5 < Eo ≤ 33.

4.2.2 Liquid phase dynamics


The liquid phase hydrodynamics is described by a set of volume-averaged
conservation equations, which consists of the continuity and Navier-Stokes

67
4. Effect of gas sparging on bubble column hydrodynamics

equations. The presence of the bubbles is reflected by the liquid phase volume
fraction αl and the interphase momentum transfer rate Φ:

∂  
αl ρl + ∇ · αl ρl u = 0 (4.12)
∂t

∂  
αl ρl u + ∇ · αl ρl uu = −αl ∇p − ∇ · (αl τl ) + αl ρl g + Φ (4.13)
∂t
The liquid phase is assumed to be Newtonian, thus the stress tensor τl can be
expressed as:
 
2
τl = −µeff,l (∇u) + (∇u)T − I(∇ · u) (4.14)
3
where µeff,l is the effective shear viscosity. In the present model, the effective
viscosity is composed of two contributions, the molecular viscosity and the
turbulent viscosity:
µeff,l = µL,l + µT,l (4.15)

where µT,l is the turbulent viscosity(or eddy viscosity), which is determined


from turbulence modeling of the liquid phase.
In the present work, a sub-grid scale model is used to simulate the turbu-
lence induced by movements of bubbles. This means that the above conti-
nuity and Navier-Stokes equations are resolved for the field representing the
“large” eddies, while the effect of the subgrid part of the velocity representing
the ”small scales” on the resolved field is included through an eddy-viscosity
subgrid-scale model.
In the present work, two eddy-viscosity models are adopted. The first
eddy-viscosity model was proposed originally by Smagorinsky (1963):

µT,l = ρl (CS ∆)2 |S| (4.16)

where CS is the Smagorinsky constant with a typical value of 0.1 which is


also adopted in the present work. ∆ = (Vcell )1/3 the filter width and S̄ the
characteristic filtered rate of strain which is calculated from:
q
S = 2Sij Sij (4.17)

68
4.2. Discrete bubble model

where Si j is the filtered rate of strain tensor:


!
1 ∂ui ∂u j
Sij = + (4.18)
2 ∂x j ∂xi

Another eddy-viscosity model used to calculate the eddy viscosity was


proposed by Vreman (2004):
s

µT,l = 2.5ρl C2S (4.19)
αij αij

where Bβ = β11 β22 − β212 + β11 β33 − β213 + β22 β33 − β223 , βij = ∆2m αmi αm j and
αij = ∂u j /∂xi . ∆i is the filter width in the i direction.

4.2.3 Collision model


In the present work, a hard sphere collision model (Hoomans et al., 1996) is
adopted to describe the bouncing of bubbles. It mainly consists of two parts.
One is processing the sequence of collisions and another one is dealing with
the collision dynamics. The former is described in detail by Darmana et al.
(2006).
A brief description of the collision model will be presented subsequently:
consider a set of N bubbles with index B = {0, 1, ..., N − 1} and a set of obstacles
(i.e. walls) O. For each bubble l ∈ B, a set of possible collision partners of l,
N(l), can be found within all the other bubbles and obstacles.
According to Allen and Tildesley (1989), the time required for the bubble
l to collide with a collision partner m ∈ N(l) can be determined from their
present positions and velocities:
q
−rlm · vlm − (rlm · vlm )2 − v2lm (r2lm − (Rl + Rm )2 )
tlm = (4.20)
v2lm

where rlm = rl − rm and vlm = vl − vm . Note that if rlm · vlm > 0 the bubbles are
moving away from each other and will not collide. In case of a collision with
a wall the collision time follows simply from the distance to the wall and the

69
4. Effect of gas sparging on bubble column hydrodynamics

normal velocity component toward that wall, which leads for a vertical wall
(i.e. xwall = 0) to the following expression:
rl,x − (xwall + Rl )
tl,wall = (4.21)
vl,x
For each bubble, the minimum collision time is determined by scanning
all collision partners for a possible collision. In order to prevent calculating
the collision time for a certain bubble and its partner twice, only those bubbles
with index larger than the bubble’s index (i.e. l) and obstacles are considered
in the possible collision partners N(l).
Once the minimum collision time for each bubble is determined, a global
minimum collision time (i.e. tab ) and the corresponding collision pair (i.e. a
and b) can be found. First, all bubble positions are updated to the instant of
the collision using a simple explicit integration:

rl (t + tab ) = rl (t) + vl tab ∀l ∈ B (4.22)

Following the movement of all bubbles, collision pairs a and b are touching
and the collision dynamics is applied to process the collision event. The post-
collision velocities of a and b can be calculated according to Hoomans et al.
(1996).

4.3 Simulation details

The bubble column studied here is shown in Figure 4.1. The cross-sectional
area of the column is 0.15 m × 0.15 m (W × D). The column is initially filled
with water to a height of 0.45 m (H). Air is used as the dispersed phase and
introduced into the column through a perforated plate at the bottom of the
column. The material properties of both phases are taken according to room
temperature. The bubble column is operating at atmospheric pressure.
The effects of different configurations of gas spargers on the column hy-
drodynamics are investigated. Perforated plates with 9 holes (3 × 3), 49 holes
(7 × 7), 225 holes (15 × 15) and 484 holes (22 × 22) as shown in Figure 4.2 are
used. All the holes have a diameter of 1 mm and are located in the central
region of the plate with a square pitch of 6.25 mm.

70
4.3. Simulation details

Figure 4.1: Sketch of the bubble column.

Figure 4.2: Schematic representation of different perforated plates.

71
4. Effect of gas sparging on bubble column hydrodynamics

Furthermore, the effect of the position of the sparged area on the perfor-
mance of the bubble column is studied by using perforated plates with 49
holes. That is, besides the perforated plate with 49 holes used above, another
two perforated plates with the same number of holes are adopted for the pur-
pose, as shown in Figure 4.3. The centers of the sparger regions are located at
one-fourth, one-third and a half of the width of the column.

Figure 4.3: Schematic representation of the location of the sparged area of


perforated plates (49 holes).

A computational grid with (20×20×60) cells is adopted in the simulations.


The superficial gas velocity is 0.005 m/s. The bubble diameter in the simula-
tions is kept as a constant d = 0.005 m, which is obtained from experimental
observations of a bubble column operated at the applied superficial gas ve-
locity. The computational time step for the liquid phase is 1.0 × 10−3 s and the
collisions among bubbles and the movements of the bubbles are processed
several times within each time step. The total simulation time is set as 120 s.
The grid size effect was checked by comparing simulation results with those
obtained from a finer grid (30 × 30 × 90). There was no significant difference
between the results of these simulations.
No bubble coalescence and breakup was considered in the present work.
Since the superficial gas velocity adopted in the present work is small, the
homogeneous flow regime prevails where bubble coalescence and breakup
are not significant. Therefore, the bubble size is assumed to be uniform in the
flow.

72
4.4. Results and discussion

4.4 Results and discussion

4.4.1 Comparisons with experimental data

In order to validate the discrete bubble model and the adopted closures, the
simulation results are compared with the experimental data reported by Deen
(2001). Information at three different heights (z/H = 0.35, 0.55, and0.75) in the
column was extracted from the simulation using a distributor with 49 holes.
The two eddy-viscosity models (Smagorinsky, 1963; Vreman, 2004) were used
for the simulations.
The profiles of the time-averaged vertical velocity of the liquid phase are
presented in the Figures 4.4(a)–4.4(c). It can be seen that the results obtained
with the eddy-viscosity model proposed by Vreman (2004) agree with the
experimental data better than those obtained with the eddy-viscosity model
of Smagorinsky (1963). This difference is much more obvious in the lower
part of the column(i.e. z/H < 0.5), whereas in the higher part of the column,
the simulation data from both two eddy-viscosity models are close to the
experimental data.
Furthermore, the profiles of turbulent kinetic energy (tke) from PIV mea-
surements and simulations are also compared (Figures 4.4(d)–4.4(f)). Note
that there are only two components of the liquid velocity available from PIV
measurements. Therefore, the reported turbulent kinetic energy obtained
from experimental data is calculated by assuming that the two horizontal
0 0
components of the liquid velocity are equal, i.e. k ≈ 12 (2ux2 + uz2 ). The turbu-
lent kinetic energy from the simulations, however, is obtained from all three
0 0 0
components of the liquid velocity, i.e. k = 21 (ux2 + u y2 + uz2 ).
It can be seen that the turbulent kinetic energy from the simulations is much
smaller in comparison with the experimental data at the lower height (z/H =
0.35). However, the difference again becomes smaller along the column height.
Meanwhile, it can clearly be seen that the simulations with the eddy-
viscosity model of Vreman (2004) are in better agreement with the experiments
than those with the eddy-viscosity model proposed by Smagorinsky (1963).
The above comparisons demonstrate that the discrete bubble model, with
supplemented proper force closures and turbulence eddy-viscosity model,

73
4. Effect of gas sparging on bubble column hydrodynamics

0.3
0.25
0.2
0.2
0.15
uz [m/s]

uz [m/s]
0.1
0.1
0.05

0.0 0
PIV data PIV data
Vreman (2004) -0.05 Vreman (2004)
Smagorinsky (1963) Smagorinsky (1963)
-0.1 -0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(a) Vertical liquid velocity (z/H = 0.35) (b) Vertical liquid velocity (z/H = 0.55)
0.2 0.03

0.15

0.1
0.02
tke [m2/s2]
uz [m/s]

0.05

0
0.01
-0.05
PIV data PIV data
-0.1 Vreman (2004) Vreman (2004)
Smagorinsky (1963) Smagorinsky (1963)
0.00
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(c) Vertical liquid velocity (z/H = 0.75) (d) Turbulent kinetic energy (z/H = 0.35)
0.03

0.02

0.02
tke [m2/s2]

tke [m2/s2]

0.01
0.01

PIV data PIV data


Vreman (2004) Vreman (2004)
Smagorinsky (1963) Smagorinsky (1963)
0.00 0.00
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(e) Turbulent kinetic energy (z/H = 0.55) (f) Turbulent kinetic energy (z/H = 0.75)

Figure 4.4: Comparisons between simulations and experimental data.

74
4.4. Results and discussion

is suited to investigate the hydrodynamical performance of bubble column


reactors. All subsequent simulation results are based on the discrete bubble
model utilizing the eddy-viscosity model proposed by Vreman (2004).

4.4.2 Effect of different gas sparger


This section discusses the performance of the columns with four different
perforated plates (3 × 3, 7 × 7, 15 × 15 and 22 × 22 holes). The time-averaged
liquid velocity profiles in the center plane of the column (y/D = 0.5) with the
four different perforated plates are shown in Figure 4.5(a)–4.5(b).
It can be seen that the vertical liquid velocity in the column with the
perforated plate of 3 × 3 has a relatively large center velocity. Since the
superficial gas velocities of all the columns are kept the same, the columns
with less holes on the plate have larger gas inlet velocity. Therefore, the profile
of vertical liquid velocity is getting flatter when the sparger area increases.
Profiles of the turbulent kinetic energy in the midplane are shown in the
Figure 4.5(c)–4.5(d). It is found that the turbulent kinetic energy in the central
region of the column with the perforated plate with 9 holes is the highest. The
turbulent kinetic energy is decreasing and becomes more flat with increasing
sparged area. Furthermore, the profile of the turbulent kinetic energy also
becomes flat with increasing height.
The distributions of the void fraction in the different systems are shown
in Figure 4.5(e)–4.5(f). The column with the perforated plate (7 × 7) has a very
high void fraction in the central region of the column (z/H = 0.35 and 0.55).
Remarkably, the column with the perforated plate (3 × 3) has a lower void
fraction in that region. A reason for this can be checked by visualization of
“bubbles“. Consequently, the void fraction in the column with the perforated
plate (3 × 3) is lower. Furthermore, the void fraction becomes flat with height
and the void fraction in the column with more holes in the perforated plate is
rather flat from the beginning and does not change too much with height.
With constant superficial gas velocity, bubbles released from the plate have
lower velocity with increasing number of holes. Furthermore, flows become
more uniform when the sparged area increases. This is consistent with the
above findings. In addition, bubbles injected from small sparged area move

75
4. Effect of gas sparging on bubble column hydrodynamics

9 holes 9 holes
0.3 49 holes 49 holes
225 holes 225 holes
484 holes 0.2 484 holes
0.2

uz [m/s]
uz [m/s]

0.1
0.1

0 0

-0.1 -0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(a) Vertical liquid velocity (z/H = 0.35) (b) Vertical liquid velocity (z/H = 0.55)
0.04 0.04
9 holes 9 holes
49 holes 49 holes
225 holes 225 holes
0.03 484 holes 0.03 484 holes
tke [m2/s2]

tke [m2/s2]
0.02 0.02

0.01 0.01

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(c) Turbulent kinetic energy (z/H = 0.35) (d) Turbulent kinetic energy (z/H = 0.55)
0.08 0.06
9 holes 9 holes
49 holes 49 holes
225 holes 225 holes
0.06 484 holes 484 holes
0.04
αg [-]

αg [-]

0.04

0.02
0.02

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(e) Void fraction (z/H = 0.35) (f) Void fraction (z/H = 0.55)
Figure 4.5: Comparisons of different gas spargers.

76
4.4. Results and discussion

to the sides more easily and thus, the local void fraction becomes smaller.

4.4.3 Effect of location of sparger area

The effect of different location of the same sparged area on a perforated plate
has also been investigated. The locations of the sparged area on the per-
forated plates are shown in Figure 4.3. For each of the configurations, the
time-averaged vertical liquid velocity, turbulent kinetic energy and local void
fraction at two different heights are plotted in Figure 4.6(a)–4.6(f).
It is quite clear that the profiles of time-averaged vertical liquid velocity
are considerably different. Due to the deviation of the sparged area from the
center of the plate, the peak of the vertical liquid velocity distribution also
moves from the center towards the wall. Meanwhile, one can also observe
that the highest vertical liquid velocity for the plate with 9 holes is slightly
larger than that of the others. This originates from the relatively intensive
liquid circulation in the bubble column.
In Figure 4.6(c)–4.6(d), the distributions of turbulent kinetic energy for
the three different systems are plotted. On the right half of the column, the
turbulent kinetic energy decreases with increasing distance of the sparged
area from the center of the column. However, in the left half of the column,
the distributions of the turbulent kinetic energy in the columns are somewhat
more complicated. In the lower part of the column, the turbulent kinetic
energy is the highest for the W/3 case. whereas, at the intermediate height,
the three turbulent kinetic energy distributions are approaching each other.
The distributions of void fraction in the three cases are also quite different
(Figure 4.6(e)–4.6(f)). However, the trends follow those of the vertical liquid
velocity. That is, the peak of the void fraction distribution shifts from the
center to the edge with increasing sparger asymmetry. Furthermore, the void
fraction is highest when the sparger is located nearest to the wall.
It can be concluded that the location of the sparged area on the perforated
plate influences the bubble column hydrodynamics significantly. It also in-
fluences the gas phase mixing in the bubble column, which will be discussed
in the following section.

77
4. Effect of gas sparging on bubble column hydrodynamics

0.4
W/4 W/4
W/3 W/3
W/2 0.2 W/2

0.2
uz [m/s]

uz [m/s]
0.1

0
0

-0.1

-0.2 -0.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(a) Vertical velocity (z/H = 0.35) (b) Vertical velocity (z/H = 0.55)
0.02 0.025

0.02
0.015
tke [m2/s2]

tke [m2/s2] 0.015


0.01
0.01

0.005
W/4 0.005 W/4
W/3 W/3
W/2 W/2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(c) Turbulent kinetic energy (z/H = 0.35) (d) Turbulent kinetic energy (z/H = 0.55)
0.1 0.05
W/4 W/4
W/3 W/3
0.08 W/2 0.04 W/2
Void fraction [-]

0.06 0.03
αg [-]

0.04 0.02

0.02 0.01

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/W [-] x/W [-]

(e) Void fraction (z/H = 0.35) (f) Void fraction (z/H = 0.55)
Figure 4.6: Comparisons of locations of gas spargers.

78
4.4. Results and discussion

4.4.4 Comparisons of residence time distributions


The concept of residence time distributions (RTD), indicated by E(t), is a very
important concept in the analysis of chemical reactors. The idea of RTD’s has
been introduced by Danckwerts (1953).
Residence time distribution theory produces valuable insight with respect
to consequences of non-ideal flow for the macro-mixing in process equip-
ment. The distribution of residence times provides considerable information
about homogeneous, isothermal reactions. For single, first-order reactions,
knowledge of the residence time distribution allows the yield to be calculated
exactly, even in flow systems of arbitrary complexity. For other reaction or-
ders, it is usually possible to calculate tight limits, within which the yield must
lie. Even if the system is nonisothermal and heterogeneous, knowledge of the
residence time distribution provides substantial insight regarding the flow
processes occurring within it (Levenspiel, 1999; Nauman, 2002; Scott Fogler,
2005).
The residence time distribution E(t) is defined in such way that the area
under the curve is unity: Z ∞
E(t) dt = 1 (4.23)
0

The mean residence time tm and the variance σ2 of the residence time
distribution are calculated from:
Z ∞
tm = tE(t) dt (4.24)
0
Z ∞
σ2 = (t − tm )2 E(t) dt (4.25)
0

The axial dispersion model has been used extensively to characterize phase
mixing in chemical reactors. The residence time distribution can be used to
evaluate the dispersion coefficient Da in the axial dispersion model. According
to the axial dispersion model, the mass balance of a species in an unsteady-
state is given by (Levenspiel, 1999):

∂C ∂2 C ∂C
= Da 2 − U (4.26)
∂t ∂z ∂z

79
4. Effect of gas sparging on bubble column hydrodynamics

where the velocity U represents the mean velocity (for the gas phase U =
U g / g ).
In a so-called closed-closed system, a Peclet number Pe can be calculated
(Levenspiel, 1999) by using tm and σ2 determined from RTD data:
σ2 2 2
2
= − 2 (1 − e−Pe )
σ2θ = (4.27)
tm Pe Pe
where σθ is the dimensionless variance in the residence time. The dispersion
coefficient Da can now be determined from the Peclet number:
UH
Da = (4.28)
Pe
In the discrete bubble model, the determination of the residence time of the
gas phase is straightforward. Each bubble traveling in the reactor is tracked
exactly by the discrete bubble model. Therefore, the residence time of each
bubble can be determined from the DBM simulations directly and thus the
residence time distribution of the gas phase can be obtained.
The resulting residence time distributions of the columns with the four
different perforated plates are shown in Figure 4.7.
0.05
9 holes
49 holes
0.04 225 holes
484 holes
E(t) [1/s]

0.03

0.02

0.01

0
0 1 2 3 4 5 6 7
t [s]

Figure 4.7: Bubble residence time distributions

One can find that the residence time distributions obtained from the DBM
simulations agree well with characteristics of typical residence time distribu-

80
4.4. Results and discussion

tion of the bubble phase reported by Molerus (1986). That is, (i) even the
fastest bubbles rise with a finite absolute velocity, i.e. the RTD shows a mini-
mum time; (ii) a steep slope of the RTD is observed for short residence times;
(iii) partial recirculation of small bubbles results in a long tail of the RTD.
The mean residence time, variance, Peclet number and dispersion coeffi-
cient in each column have been calculated from the residence time distribu-
tions shown in Figure 4.7 and are listed in Table 4.1.
According to the Table 4.1, the gas phase dispersion coefficient Da of
the column decreases with decreasing sparger area. This implies that the
degree of backmixing of the gas phase in the bubble column is increasing
with decreasing sparger area.

Table 4.1: Gas holdup and calculations from RTDs of the columns.

7×7
Perforated plate 3×3 15 × 15 22 × 22
W W W
4 3 2
 [-] 0.016 0.014 0.015 0.016 0.020 0.024
tm [s] 1.37 1.23 1.31 1.35 1.61 1.92
σ2 [s2 ] 0.34 0.25 0.25 0.25 0.32 0.18
σ2θ [-] 0.18 0.17 0.15 0.14 0.12 0.05
Pe [-] 9.93 13.50 11.00 12.64 15.13 39.93
Da [m2 /s] 0.014 0.015 0.012 0.011 0.008 0.002

In addition to the comparisons of the columns with different sparger area,


the mixing in the columns with the same sparger area but with different loca-
tions on the bottom plate is also studied. The perforated plates with 49 holes
as shown in Figure 4.3 are used for this purpose. The residence time distri-
butions of these three cases are shown in Figure 4.8. The relevant parameters
characterizing the residence time distribution and the phase mixing in the
columns are calculated and listed in Table 4.1.
The residence time distributions of the three columns have the same vari-
ance but the mean residence time decreases as the location of the sparged area
deviates more from the center of the bottom plate. The axial dispersion coef-
ficient, however, increases meanwhile, which means that the extent of mixing

81
4. Effect of gas sparging on bubble column hydrodynamics

W/4
0.06 W/3
W/2

E(t) [1/s]
0.04

0.02

0
0 1 2 3 4 5 6 7
t [s]

Figure 4.8: Gas phase residence time distributions (49 holes)

becomes more intensive. That is quite reasonable because the flow field in the
column with a asymmetrical gas sparger is much more irregular than that in
a column with a symmetrical gas sparger.

4.5 Conclusions

A discrete bubble model (DBM) has been utilized to investigate the perfor-
mance of a laboratory scale bubble column with a square cross-section. Firstly,
the model including its closures has been validated by comparing the simula-
tion results with experimental data. Two eddy-viscosity models were adopted
for the comparison purpose. Computed time-averaged profiles of vertical liq-
uid velocity and turbulent kinetic energy of the liquid phase at three different
heights (z/H = 0.35, 0.55 and 0.75) were compared with PIV data. The results
show that the eddy-viscosity model proposed by Vreman (2004) performs
better than the model proposed by Smagorinsky (1963). Therefore, the for-
mer eddy-viscosity model was adopted for further investigations. Several
columns with different perforated plates were simulated to study the effect
of the gas sparger on the bubble column hydrodynamics. It was found that
the distributions of liquid velocity, turbulent kinetic energy and void fraction

82
Nomenclature

become uniform as the sparger area increases.


Furthermore, the effect of the sparger location was investigated. The liquid
phase circulation becomes more pronounced as the sparger is moved towards
the wall.
The gas phase residence time distributions of the columns were also ob-
tained from the discrete bubble model and were used to characterize the
macro-mixing of the gas phase in the column in terms of an axial dispersion
model. It is found that the extent of mixing increases when the sparger area
decreases. For instance, the dispersion coefficient of the column with the per-
forated plate of 3 × 3 is seven times larger than that of the column with the
perforated plate of 22 × 22. Furthermore, the effect of the sparger location on
the phase mixing is also investigated by comparing columns with three dif-
ferent sparger locations. In terms of the gas phase residence time distribution,
the mean residence time decreases as the sparger area deviates more from the
center of the plate. The variance of the distribution, however, does not change
with the location of the sparger area. That is, the width of the distribution
remains the same. It also turns out that the location of the sparged area on
the perforated plate influences the phase mixing. The dispersion coefficient
increases as the sparger is moved towards the wall.
The present work focused on a small scale bubble column operating at
low superficial gas velocity without consideration of bubble coalescence and
breakup. However, bubbly flows in large scale reactors at high superficial gas
velocity are more common in industrial applications. Therefore, discrete bub-
ble modeling with bubble coalescence and breakup model is necessary for the
simulation of large scale bubble column reactors operating at superficial gas
velocity. Future work will focus on these aspects. In addition to experimental
studies, discrete bubble simulation offers us an alternative method to study
phase macro-mixing in bubble column reactors.

Nomenclature

B bubbles
C model coefficient, [-]; concentration of a species, [kg m−3 ]

83
4. Effect of gas sparging on bubble column hydrodynamics

d bubble diameter, [m]


D column depth, [m]
Da dispersion coefficient, [m2 s−1 ]
(ρl − ρ g )gd2
Eo Eötvös number, Eo = , [-]
σ
−1
E(t) residence time distribution, [s ]
F force vector, [N]
g gravitational acceleration, [m s−2 ]
H column height, [m]
I unit tensor, [-]
n unit normal vector, [-]
N collision partners
O obstacles
p pressure, [N m−2 ]
Pe Peclet number, [-]
r position vector, [m]
R bubble radius, [m]
ρl |v − u|d
Re bubble Reynolds number, Re = , [-]
µl
S characteristic filtered strain rate, [s−1 ]
t time, [s]
u liquid velocity vector, [m s−1 ]
U mean bubble velocity, [m s−1 ]
v bubble velocity vector, [m s−1 ]
V bubble volume, [m3 ]
W column width, [m]
y distance to the wall, [m]

Greek letters

α void fraction, [-]


∆ subgrid length scale, [m]
 gas holdup, [-]
µ viscosity, [kg m−1 s−1 ]

84
References

Φ volume averaged momentum transfer due to interphase forces,


[kg m−2 s−2 ]
ρ density, [kg m−3 ]
σ surface tension, [N m−1 ]
σ2 variance, [s2 ]
τ stress tensor, [N m−2 ]

Indices

cell computational cell


D drag
eff effective
g gas phase
G gravity
i i direction
j j direction
l liquid phase
L lift; molecular viscosity
m mean; bubble index of possible collision partner
P pressure
S subgrid
T turbulent
VM virtual mass
W wall
z vertical direction

References

M. P. Allen and D. J. Tildesley. Computer Simulation of Liquids. Oxford Univer-


sity Press, USA, June 1989.

T. R. Auton. The lift force on a spherical body in a rotational flow. J. Fluid


Mech., 197:241–257, 1987.

85
4. Effect of gas sparging on bubble column hydrodynamics

M. R. Bhole, S. Roy, and J. B. Joshi. Laser Doppler anemometer measurements


in bubble column: Effect of sparger. Industrial and Engineering Chemistry
Research, 45(26):9201–9207, 2006.

P. V. Danckwerts. Continuous flow systems : Distribution of residence times.


Chemical Engineering Science, 2(1):1–13, February 1953.

D. Darmana, N. G. Deen, and J. A. M. Kuipers. Parallelization of an Euler-


Lagrange model using mixed domain decomposition and a mirror domain
technique: Application to dispersed gas-liquid two-phase flow. Journal of
Computational Physics, 220(1):216–248, 2006.

N. G. Deen. An experimental and computational study of fluid dynamics in gas-


liquid chemical reactors. PhD thesis, Aalborg University, Denmark, 2001.

E. Delnoij, F. A. Lammers, J. A. M. Kuipers, and W. P. M. van Swaaij. Dynamic


simulation of dispersed gas-liquid two-phase flow using a discrete bubble
model. Chemical Engineering Science, 52(9):1429–1458, 1997.

E. Delnoij, J. A. M. Kuipers, and W. P. M. van Swaaij. A three-dimensional


CFD model for gas-liquid bubble columns. Chemical Engineering Science, 54
(13-14):2217–2226, 1999.

G. Hebrard and D. Bastoul. Influence of the gas sparger on the hydrodynamic


behaviour of bubble columns. Chemical Engineering Research and Design, 74
(3):406–414, 1996.

B. P. B. Hoomans, J. A. M. Kuipers, W. J. Briels, and W. P. M. van Swaaij. Dis-


crete particle simulation of bubble and slug formation in a two-dimensional
gas-fluidised bed: A hard-sphere approach. Chemical Engineering Science,
51(1):99–118, 1996.

G. Hu and I. Celik. Eulerian-Lagrangian based large-eddy simulation of a


partially aerated flat bubble column. Chemical Engineering Science, 63(1):
253–271, 2008.

O. Levenspiel. Chemical Reaction Engineering. John Wiley & Sons, 3rd edition,
1999.

86
References

O. Molerus. Modelling of residence time distributions of the gas phase in bub-


ble columns in the liquid circulation regime. Chemical Engineering Science,
41(10):2693–2698, 1986.

E. B. Nauman. Chemical Reactor Design, Optimization, and Scaleup. McGraw-


Hill, 2002.

H. Scott Fogler. Elements of Chemical Reaction Engineering. Prentice Hall PTR,


4th edition, August 2005.

Y. T. Shah, B. G. Kelkar, S. P. Godbole, and W. -D. Deckwer. Design parameters


estimations for bubble column reactors. AIChE Journal, 28(3):353–379, 1982.

J. Smagorinsky. General circulation experiment with the primitive equation.


Monthly Weather Review, 91:99–165, 1963.

M. Sommerfeld, E. Bourloutski, and D. Bröder. Euler/Lagrange calculations


of bubbly flows with consideration of bubble coalescence. Canadian Journal
of Chemical Engineering, 81(3-4):508–518, 2003.

B. N. Thorat. Effect of sparger design and height to diameter ratio on fractional


gas hold-up in bubble columns. Chemical Engineering Research and Design,
76(A7):823–834, 1998.

A. Tomiyama, T. Matsuoka, T. Fukuda, and T. Sakaguchi. A simple numerical


method for solving an incompressible two-fluid model in a general curvi-
linear coordinate system. Advances in Multiphase Flow, pages 241–252,
Amsterdam, 1995. Society of Petroleum Engineers Inc.

A. Tomiyama, I. Kataoka, I. Zun, and T. Sakaguchi. Drag coefficients of single


bubbles under normal and micro gravity conditions. JSME international
journal. Ser. B, Fluids and thermal engineering, 41(2):472–479, 1998.

U. P. Veera and J. B. Joshi. Measurement of gas hold-up profiles by gamma ray


tomography: Effect of sparger design and height of dispersion in bubble
columns. Chemical Engineering Research and Design, 77(4):303–317, 1999.

87
4. Effect of gas sparging on bubble column hydrodynamics

A. W. Vreman. An eddy-viscosity subgrid-scale model for turbulent shear


flow: Algebraic theory and applications. Physics of Fluids, 16(10):3670–3681,
2004.

D. Zhang, N. G. Deen, and J. A. M. Kuipers. Numerical simulation of the dy-


namic flow behavior in a bubble column: A study of closures for turbulence
and interface forces. Chemical Engineering Science, 61(23):7593–7608, 2006.

88
Chapter
5
Discrete bubble modeling of
bubbly flows: Swarm effects
The performance of several drag correlations reported in literature for bubble swarms
has been investigated. A discrete bubble model (DBM) based on the Eulerian-
Lagrangian approach was adopted for this purpose. Numerical simulations for a
square bubble column were performed and the results were compared with PIV mea-
surements. The drag model reported by Lima Neto et al. (2008) predicts the vertical
liquid velocity and the relative velocity better than the other drag models at a superfi-
cial gas velocity u g = 0.0024 m/s. As the superficial gas velocity increases, however,
Lima Neto’s drag model and Wen & Yu’s model tend to overestimate the relative
velocity between the two phases. Among the other models, Rusche’s model gives a
better prediction of the vertical liquid velocity in the lower part of the bubble column.
Furthermore, most of the models can predict the liquid velocity in higher parts of the
column well. However, there are still some aspects that need to be considered and
improved to advance the accurate simulation of bubbly flows at high void fraction.

89
5. Discrete bubble modeling of bubbly flows: Swarm effects

5.1 Introduction

Two-phase gas-liquid flows are encountered in several large scale processes,


in the chemical, biochemical, metallurgical and petrochemical industries. De-
tailed knowledge on hydrodynamics of two-phase gas-liquid flows is essential
for design and scale-up of chemical reactors. Computational Fluid Dynamics
(CFD) offers the possibility to investigate hydrodynamics of these complex
systems and reduce the experimental effort required for design and scale-up
of related process equipments. During the past decades, numerical simulation
of multiphase flows has received considerable attention (Jakobsen, 2008).
Eulerian-Lagrangian modeling, as one of the two major numerical ap-
proaches, has been adopted to study bubbly flows in recent years (Trapp,
1993; Lapin, 1994; Delnoij et al., 1997, 1999; Sommerfeld et al., 2003; Darmana
et al., 2006; Hu and Celik, 2008).
For numerical simulation of bubbly flows, reliable closures are required to
represent the interfacial momentum transfer rate (i.e. the effective drag acting
on bubbles). A number of theoretical and experimental studies have been
conducted to evaluate the drag coefficient of single bubbles rising in quies-
cent liquids (Clift et al., 1978; Zun and Grošelj, 1996; Tomiyama et al., 1998).
However, industrial processes are generally operated at high void fractions.
Ishii and Zuber (1979) and Ishii and Hibiki (2005) have developed constitu-
tive relations for the drag force and the relative velocity in bubbly, droplet
and particulate flows for a wide range of dispersed volume fraction. They
identified four flow regimes for bubbles: viscous regime, distorted regime,
churn-turbulent regime and slug regime. According to them, the dispersed
phase fraction has a considerable influence on the drag coefficient of the dis-
persed phase. Holland and Bragg (1995) reported that the relative velocity
between the dispersed phase and the continuous phase is expressed as the
product of the terminal velocity of a single dispersed element and a correction
factor (1 − α g )n−1 . For bubble columns, the expression shows that the relative
velocity of a bubble rising in a swarm is lower than the terminal velocity of
an isolated bubble in the liquid. It follows that the presence of neighboring
bubbles increases the drag on a bubble. Moreover, the value of n can be ap-

90
5.1. Introduction

proximated as n = 2 in practice. Garnier et al. (2002) measured the relative


velocity between the gas and the liquid phase in an air-water system at high
void fractions (0 − 0.3) using a double optical probe and hot-film anemometry.
They found that the average relative velocity decreases when the void frac-
tion increases. Their measurements implied that the ratio between the drag
coefficient of a bubble in a swarm and that of an isolated bubble in an infinite
−2
liquid can be expressed as (1 − Cα1/3 g ) . where C is a constant and about
unity. Rusche (2002) reviewed the correlations to determine the drag force
in dispersed two-phase flows at low and high void fractions and proposed
his own correlation. By comparing those correlations with experimental data
in literature, Rusche found that the new correlation gives the best prediction
for the terminal velocity of bubbles rising in swarms. Meanwhile, he also
pointed out that the correlation of Wen and Yu (1966) can give reasonable
results, although it was originally developed for fluid-particle systems. The
correlations of Rusche have been implemented in a two-fluid model and im-
provement of the quality of the predictions was reported (Behzadi et al., 2004).
By means of the volume of fluid method, Bertola et al. (2004) simulated the
motion of a bubble swarm. They showed that different bubbles of different
size possess a different trend of relative velocity versus gas holdup. Simonnet
et al. (2007) studied the relative velocity in a swarm of bubbles using Laser
Doppler Velocimetry (LDV) and a double optical probe. They found that
below a critical value of the local void fraction, the relative velocity decreases
due to hindrance effects when the void fraction increases. Beyond this crit-
ical value, the relative velocity increases with the local void fraction. They
reported that the presence of surfactants gives a totally different trend com-
pared to that for pure water. Finally, they proposed a new correlation for the
drag coefficient embedding Jamialahmadi et al. (1994)’s correlation for predic-
tion of terminal rise velocity of single bubbles. In their numerical simulations
Gentric et al. (2008) reproduced the characteristic evolution of the gas holdup
against superficial gas velocity and captured some characteristic features of
different flow regimes by using this new drag closure. Lima Neto et al. (2008)
investigated air-water bubbly jets in a large water tank experimentally. They
found that the relative velocities exceeded the terminal velocities of isolated

91
5. Discrete bubble modeling of bubbly flows: Swarm effects

bubbles given by Clift et al. (1978) and proposed a new correlation for the
drag coefficient as function of the bubble Reynolds number.
In the present work, Eulerian-Lagrangian modelling of bubbly flows in a
square column is investigated with emphasis on the performance of different
closures for drag.

5.2 Discrete bubble model

Our discrete bubble model (DBM) was originally developed by Delnoij et al.
(1997, 1999) and is based on volume-averaged continuity and Navier-Stokes
equations for the liquid phase, while the motion of each individual bubble is
tracked in a Lagrangian fashion taking into account bubble-bubble encoun-
ters.

5.2.1 Bubble dynamics


The motion of each individual bubble is computed from Newton’s second
law. For an individual bubble, the equation of motion and bubble trajectory
equation can respectively be written as:

dv
ρg V = ΣF (5.1)
dt

dr
=v (5.2)
dt
The net force acting on each individual bubble is calculated by considering
all the relevant forces. It is assumed that the net force is composed of separate,
uncoupled contributions due to gravity, pressure, drag, lift, virtual mass and
wall force respectively:

ΣF = FG + FP + FD + FL + FVM + FW (5.3)

The gravity force acting on a bubble in a liquid is given by:

FG = ρ g Vg (5.4)

92
5.2. Discrete bubble model

The far field pressure force incorporating contributions of the Archimedes


buoyancy force, inertial forces and viscous strain is given by:

FP = −V∇P (5.5)

The drag force exerted on a bubble rising in a liquid is expressed as:

1
FD = − CD ρl πd2 |v − u|(v − u) (5.6)
8
where CD represents the drag coefficient (see section 5.3).
A bubble rising in a non-uniform liquid flow field experiences a lift force
due to vorticity or shear in the flow field. The shear induced lift force acting
on a bubble is usually written as (Auton, 1987):

FL = −CL ρl V(v − u) × ∇ × u (5.7)

The lift coefficient is calculated according to Tomiyama et al. (2002):




 min[0.288tanh(0.121Re), f (EoH )] EoH < 4,




CL = 
 f (EoH ) 4 ≤ EoH ≤ 10, (5.8)



−0.29 EoH > 10.

where
f (EoH ) = 0.00105EoH 3 − 0.0159EoH 2 + 0.474 (5.9)

EoH is the Eötvös number defined by using the maximum horizontal dimen-
sion of a bubble:
(ρl − ρ g )gd2H
EoH = (5.10)
σ
The maximum horizontal diameter of the bubble is obtained from the
bubble aspect ratio E according to Wellek et al. (1966):

dV 1
E= = (5.11)
dH 1 + 0.163Eo0.757
where dV is the maximum vertical diameter of the bubble and Eo is the Eötvös
(ρl − ρ g )gd2
number, Eo = .
σ

93
5. Discrete bubble modeling of bubbly flows: Swarm effects

The relation between the above two diameters and the diameter of the
bubble d in the discrete bubble modeling is as follows:

d = (dV d2H )1/3 (5.12)

Accelerating bubbles experience a resistance, which is described as the


virtual mass force (Auton, 1987):
 
Dv Du
FVM = −CVM ρl V − (5.13)
Dt Dt
where the D/Dt operators denote the substantiative derivatives pertaining to
the respective phases. In the present work, bubbles are assumed to have a
spherical shape and a virtual mass coefficient of 0.5 is used.
Bubbles in the vicinity of a solid wall experience a force referred to as the
wall force (Tomiyama et al., 1995):
" #
1 1 1
FW = − C W d 2 − ρl |(v − u)· nz |2 nW (5.14)
2 y (L − y)2
where nz and nW , respectively, are the normal unit vectors in the vertical and
wall normal direction, L is the dimension of the system in the wall normal
direction, and y is the distance to the wall in that direction. Finally, the wall
force coefficient CW is given by:



exp(−0.933Eo + 0.179) 1 ≤ Eo ≤ 5,
CW =  (5.15)
0.007Eo + 0.04 5 < Eo ≤ 33.

5.2.2 Liquid phase dynamics


The liquid phase hydrodynamics is described by a set of volume-averaged
conservation equations for mass and momentum. The presence of the bub-
bles is reflected by the liquid phase volume fraction αl and the interphase
momentum transfer rate Φ:
∂  
αl ρl + ∇ · αl ρl u = 0 (5.16)
∂t
∂  
αl ρl u + ∇ · αl ρl uu = −αl ∇p − ∇ · (αl τl ) + αl ρl g + Φ (5.17)
∂t

94
5.2. Discrete bubble model

The liquid phase is assumed to be Newtonian, thus the stress tensor τl can be
expressed as:
 
2
τl = −µeff,l ((∇u) + (∇u)T − I(∇ · u) (5.18)
3

where µeff,l is the effective viscosity. In the present model, the effective viscos-
ity is composed of two contributions, the molecular viscosity and the turbulent
viscosity:

µeff,l = µL,l + µT,l (5.19)

where µT,l is the turbulent viscosity(or eddy viscosity), which is determined


from turbulence modeling of the liquid phase.
In the present work, a sub-grid scale model is used to simulate the tur-
bulence induced by bubble movement. This means that the conservation
equations account for “large eddies”, while the effect of the “subgrid“ eddies
are accounted for through an eddy-viscosity model.
The model proposed by Vreman (2004) was used to calculate the eddy
viscosity:
s

µT,l = 2.5ρl C2S (5.20)
αij αij

where Bβ = β11 β22 − β212 + β11 β33 − β213 + β22 β33 − β223 , βij = ∆2m αmi αm j and
αij = ∂u j /∂xi . ∆i is the filter width in the i direction.

5.2.3 Collision model

In the present paper, a hard sphere collision model (Hoomans et al., 1996)
is adopted to describe the (possible) bouncing of bubbles. It consists of two
main parts. One part is processing the sequence of collisions and another part
is dealing with the collision dynamics. The former is described in detail by
Darmana et al. (2006). More details are given in Chapter 4 of this thesis.

95
5. Discrete bubble modeling of bubbly flows: Swarm effects

5.3 Drag coefficient correlations

5.3.1 Wen & Yu’s model

Although the correlation of Wen and Yu (1966) was proposed for gas-solid sys-
tems, we used this model here for the purpose of comparison. The associated
expression for the drag coefficient is give by:


 24 −1.65
 Re (1 + 0.15Res )(1 − α g ) Res < 1000
0.687

CD = 
 s (5.21)

0.44(1 − α g )−1.65 Res ≥ 1000

where the bubble Reynolds number Res is based on the superficial velocity:

(1 − α g )ρl |v − u|d
Res = (5.22)
µl

5.3.2 Ishii & Zuber’s model

According to Ishii and Zuber (1979) and Ishii and Hibiki (2005), for the drag
correlation of bubbles four flow regimes should be distinguished: the viscous
regime, the distorted flow regime, the churn-turbulent flow regime and the
slug regime.
The drag coefficient correlations are given as follows:
 24

 (1 + 0.1Re0.75

 s ) Viscous regime

 Re

 2 √ " 1 + 17.67(1 − α g )1.3 #2
s




 Eo Distorted regime
CD = 
 3 18.67(1 − α g )1.5 (5.23)



 8

 3 (1 − α g )
2
 Churn-turbulent regime



9.8(1 − α g )3 Slug regime

During the numerical simulations, the flow regime is distinguished ac-


cording to Morud and Hjertager (1996). That is, the flow is in the distorted
regime for α g ≤ 0.3 and in the churn-turbulent regime for 0.3 < α g < 0.7.

96
5.3. Drag coefficient correlations

5.3.3 Holland’s model


According to Holland and Bragg (1995), the drag coefficient of a bubble in a
bubble swarm can be expressed as follows :

CD = CD∞ (1 − α g )−2 (5.24)

where CD∞ is the drag coefficient of an isolated bubble.

5.3.4 Tomiyama’s single bubble model


Tomiyama et al. (1998) developed a simple but accurate drag coefficient model
for single bubbles. The model consists of three correlations, respectively
corresponding to pure, slightly contaminated, and contaminated gas-liquid
systems. For the purpose of comparison of the drag coefficient for the case
with bubble swarms, the correlation for pure gas-liquid systems is adopted:
h h 16 48 i 8 Eo i
CD∞ = max min (1 + 0.15Re0.687 ), , (5.25)
Re Re 3 Eo + 4
ρl |v − u|d
where Re is the bubble Reynolds number, Re = .
µl

5.3.5 Rusche’s model


According to Rusche (2002), the drag coefficient can be expressed as the prod-
uct of the drag coefficient for an isolated bubble in an infinite stagnant liquid
and a correction factor:
CD = CD∞ f (α g ) (5.26)
where f (α g ) represents the effect arising from the presence of other bubbles:

f (α g ) = exp(3.64α g ) + α0.864
g (5.27)

5.3.6 Simonnet’s model


Simonnet et al. (2007) gave the following relation for air-pure water systems:
 ! −2/25
 α g 25 
CD = CD∞ (1 − α g ) (1 − α g ) + 4.8
25
 (5.28)
1 − αg 

97
5. Discrete bubble modeling of bubbly flows: Swarm effects

where CD∞ is the drag coefficient of single bubbles in infinite liquids, which
is deduced from Jamialahmadi et al. (1994)’s correlation:
V1 V2
V∞ = q (5.29)
V12 + V22

where
1 ∆ρ 2 3µl + 3µb
V1 = gd (5.30)
18 µl 2µl + 3µb
and s
2σ gd
V2 = + (5.31)
d(ρl + ρ g ) 2
The drag coefficient for a single bubble in infinite liquids CD∞ is then
determined as follows:
4 (ρl − ρ g )gd
CD∞ = (5.32)
3 V∞ 2
ρl
Due to the fact that Simonnet’s model only provides the drag coefficient
correlation in the range of the local void fraction between 0 and 0.3, we use
the drag coefficient at 0.3 when the local void fraction exceeds 0.3.

5.3.7 Lima Neto’s model


Lima Neto et al. (2008) studied air-water bubbly jets in a large water tank and
proposed the following correlation for the drag coefficient in bubble swarms:

CD = 0.0828lnRe − 0.403 450 < Re < 10000 (5.33)

where Re is the bubble Reynolds number.


Since Lima Neto et al. (2008) only gave the drag coefficient correlation
within the range of Re from 450 to 10, 000, the drag coefficient at Re below 450
is determined from the correlation from Tomiyama et al. (1998).

5.3.8 Comparison of the drag coefficient models


In order to compare the above models for the drag coefficient in an air-water
system, we here assume that the bubble size is equal to 0.005 m and addition-
ally that the relative velocity between the gas phase and the liquid phase is

98
5.3. Drag coefficient correlations

0.2 m/s. And thus, the drag coefficient corresponding to each model can be
calculated as function of the local void fraction, which is shown in Figure 5.1.
Note that the drag coefficients from those models independent of the local
void fraction, such as Tomiyama’s model and Lima Neto’s model, are treated
as constants in the plot. Moreover, for those models without explicit specifi-
cation of the correlation for single bubble, i.e. Holland’s model and Rusche’s
model, the drag coefficient CD∞ is calculated according to the correlation of
Tomiyama et al. (1998).

10
Wen & Yu (1966)
Ishii & Zuber (1979)
8 Holland (1995)
Tomiyama (1998)
6 Rusche (2002)
CD [-]

Simonnet (2007)
Lima Neto (2008)
4

0
0 0.2 0.4 0.6 0.8
αg [-]

Figure 5.1: Comparison of the drag coefficient models at a relative velocity of


0.2 m/s.

From Figure 5.1, one can see that the drag coefficient obtained from Lima
Neto’s model is very small compared to the others and about 8.5 times smaller
than that from Tomiyama’s single bubble model. Based on Wen & Yu’s model,
the drag coefficient increases monotonically and the slope becomes larger with
increasing local void fraction. Furthermore, the drag coefficient obtained
from Ishii & Zuber’s model first increases slightly up to α g = 0.3 and then
decreases. The drag coefficient increases considerably with α g fraction in
Holland’s and Rusche’s models. On the contrary, the drag coefficient obtained
from Simonnet’s model first slightly increases with α g . Subsequently the drag
coefficient decreases rapidly beyond the local void fraction of 0.15. When

99
5. Discrete bubble modeling of bubbly flows: Swarm effects

the local void fraction approaches 0.3, the drag coefficient obtained from
Simonnet’s model approaches the value from Lima Neto’s model.

5.4 Simulation details

The bubble column studied here is shown in Figure 5.2. The cross-sectional
area of the column is 0.15 m × 0.15 m (W × D). The column is initially filled
with water to a height of 0.45 m (H). Air is used as the dispersed phase and
introduced into the column through a perforated plate with 49 holes (7 × 7) at
the bottom of the column. The material properties of both phase are taken at
room temperature (20 °C) and atmospheric pressure.
A computational grid with (20×20×60) cells is adopted in the simulations.
Four superficial gas velocities, i.e. u g = 0.0024 m/s, 0.0049 m/s, 0.0073 m/s and
0.0097 m/s are used. The bubble diameter in the simulations is assumed to
be a constant, 0.005 m. Note that this assumption may not be correct since
the bubble size may vary in the applied range of superficial gas velocities due
to coalescence and breakup of bubbles. For the purpose of comparison of
different drag coefficient correlations for bubbles rising in a swarm, however,
a uniform bubble size is used here. Moreover, the time step in the simulation
for the liquid phase is set as 1.0 × 10−3 s and the collisions among bubbles and
the movements of the bubbles are processed multiple times within each time
step. The total simulation time is 150 s.

5.5 Results and discussion

5.5.1 Instantaneous gas holdup and liquid velocity


The gas holdup in the bubble column was logged at every 0.04 s. An example
of the history of the overall gas holdup obtained with Rusche’s drag model is
shown in Figure 5.3(a) for u g = 0.0024 m/s.
It can be seen that the gas holdup first increases rapidly since the sim-
ulation starts and reaches a maximum value very soon. After reaching the
maximum value, the gas holdup decreases in a very short time and then,
starts fluctuating. The increase of the gas holdup reflects that the bubbles

100
5.5. Results and discussion

Figure 5.2: Sketch of the bubble column.

are continuously injected into the bubble column and start rising in the col-
umn. At the moment when the peak maximum is reached, some bubbles
have already reached the level of the liquid surface and most of them escape
from the column. Hence, there is a rapid decrease in the curve right after the
peak. Once the circulation flow pattern has developed, some bubbles at the
top of the column may be trapped by the circulating liquid and remain in the
column. Some of these bubbles may leave the column soon, whereas others
may take a little more time to escape. These effects produce the fluctuation
of the gas holdup in the curve. In addition, the instantaneous vertical liquid

101
5. Discrete bubble modeling of bubbly flows: Swarm effects

0.01 0.8

0.6
0.008

0.4

uz [m/s]
0.006
ε [-]

0.2
0.004
0

0.002
-0.2

0 -0.4
0 50 100 150 0 50 100 150
t [s] t [s]

(a) Instantaneous gas holdup [-] (b) Instantaneous liquid velocity [m/s]

Figure 5.3: Instantaneous quantities vs simulation time.

velocity at position (x/W = 0.5, y/D = 0.5, z/H = 0.96) is also plotted versus
time (Figure 5.3(b)). The obtained trend is similar to that obtained for the gas
holdup. However, unlike the gas holdup, the liquid velocity shows more fluc-
tuations at the beginning of the simulation. The averaging of hydrodynamic
quantities, i.e. velocities of both phases, is initiated 30 seconds after the start
of the simulation.

5.5.2 Comparisons of different drag models


In this section, the simulation results will be compared with PIV measure-
ments of Deen (2001). The comparison is divided into four parts according to
the superficial gas velocity. For each superficial gas velocity, average quan-
tities, such as vertical liquid velocity and relative velocity, at three different
heights (z/H = 0.3, 0.5 and 0.7) in the central plane (y/D = 0.5) of the bubble
column will be compared using the drag closures reported in section 5.3.

u g = 0.0024 m/s

Profiles of the average vertical liquid velocity and relative velocity between
the gas phase and the liquid phase at three heights are shown in Figure 5.4
and compared with the PIV measurements at corresponding heights. The

102
5.5. Results and discussion

results reveal that Lima Neto’s drag correlation predicts the vertical liquid
velocity better than all other correlations in the central region of the column.
However, the model underestimates the vertical liquid velocity in the region
between the center and the wall, particularly in the right part of the column
(x/W > 0.5). Among the other five drag coefficient correlations, Wen & Yu’s
model exhibits better performance for the prediction of the vertical liquid
velocity compared to the other models. Simonnet’s model has the largest
deviation in the central region in lower parts of the column (z/H ≤ 0.5). Ishii
& Zuber’s correlation and Tomiyama’s single bubble correlation show similar
performance regarding the prediction of the vertical liquid velocity in the
central region. Rusche’s model combined with Tomiyama’s single bubble
correlation shows moderate performance of all the six models. The similar
behavior is also found with Holland’s model combined with Tomiyama’s
single bubble correlation.
In addition, by comparing the relative velocity between the two phases
obtained from both the PIV measurements and the simulations, one can find
that the relative velocities between the gas phase and the liquid phase ob-
tained with Wen & Yu’s model and Lima Neto’s model for the drag coefficient
correlations are higher than those with the other four correlations. This is
due to the fact that the drag forces calculated from these two correlations are
smaller than the others. Furthermore, the other four correlations produce
similar profiles along the x direction at the three heights. In the lower parts
of the column, i.e. z/H = 0.3, Lima Neto’s model overestimates the relative
velocity in the central region. Along the height, however, this model can
predict the relative velocity quite well. This also holds for Wen & Yu’s model.

u g = 0.0049 m/s

In Figure 5.5, the simulation results and PIV measurements at u g = 0.0049 m/s
are compared with each other. It can be seen that all the models again overesti-
mate the vertical liquid velocity in the central region at low part of the column
(z/H ≤ 0.5). Ishii & Zuber’s model and Tomiyama’s model for single bubbles
exhibit similar trends and produce the largest deviation from the PIV data in

103
5. Discrete bubble modeling of bubbly flows: Swarm effects

(a) Vertical liquid velocity (z/H = 0.3) (b) Relative velocity (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Relative velocity (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Relative velocity (z/H = 0.7)

Figure 5.4: Comparisons at superficial gas velocity u g = 0.0024 m/s.

104
5.5. Results and discussion

that region. Moreover, the other four models yield better predictions with
respect to the vertical liquid velocity. At increased heights, i.e. z/H = 0.7,
one can find that Wen & Yu’s model and Simonnet’s model perform well.
Furthermore, Rusche’s model also predicts the vertical liquid velocity quite
well in the central region. However, the profile of the vertical liquid velocity
according to Rusche’s drag model seems to underestimate the velocity in the
left part of the column (x/W < 0.5) and slightly overestimates the velocity
in the right part (x/W > 0.5). On the contrary, Holland’s model seems can
predict the vertical liquid velocity in the left part of the column well, while
the model underestimates the liquid velocity in the right part. In addition, it
is worth to mention that Simonnet’s drag model can predict the vertical liquid
velocity away from the central region very well in the entire bubble column.
Comparisons for the relative velocity are presented in Figure 5.5(b), Fig-
ure 5.5(d) and Figure 5.5(f). It is quite clear that Lima Neto’s drag model
overestimates the relative velocity at all three heights. Wen & Yu’s model can
predict the relative velocity well in the central region of the lower part of the
column (z/H ≤ 0.5) compared with the other models. However, it overesti-
mates the relative velocity at the height z/H = 0.7. Furthermore, Simonnet’s
drag model, Ishii & Zuber’s model as well as Tomiyama’s model are able to
predict the relative velocity quite well. Finally Rusche’s drag model underes-
timates the relative velocity between the two phases slightly, while Holland’s
model has better estimation.

u g = 0.0073 m/s

When the superficial gas velocity increases up to 0.0073 m/s, it can be seen
that Tomiyama’s model for isolated bubbles produces large deviations for the
vertical liquid velocities in the central region of the column. However, the
model predicts the vertical liquid velocity well in the upper part of the column.
It is also clear that all the other models again overestimate the vertical liquid
velocity in the central region in the lower part of the column (z/H ≤ 0.5). The
models due to Ishii & Zuber, Holland, Rusche and Lima Neto yield better
prediction with respect to the vertical liquid velocities. In the upper part of
the column (z/H = 0.7), most of the models can predict the vertical liquid

105
5. Discrete bubble modeling of bubbly flows: Swarm effects

(a) Vertical liquid velocity (z/H = 0.3) (b) Relative velocity (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Relative velocity (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Relative velocity (z/H = 0.7)

Figure 5.5: Comparisons at superficial gas velocity u g = 0.0049 m/s.

106
5.6. Conclusions

velocity quite well except that the models of Wen & Yu, Holland, Simonnet
and Lima Neto underestimate the vertical liquid velocity in either the left or
the right part of the column.
By comparing the profiles of the relative velocity from numerical sim-
ulations with PIV measurements, it can been seen that Wen & Yu’s model
and Lima Neto’s model clearly increasingly overestimate the relative velocity
between gas phase and liquid phase along the column height. In addition,
the other models predict the relative velocity quite well in the central region
compared to the wall region. However, one can find that these models agree
with PIV measurements at height z/H = 0.7 very nicely.

u g = 0.0097 m/s

At a superficial gas velocity u g = 0.0097 m/s, one can find that the vertical
liquid velocity predicted with all the drag models is much larger than the
PIV measurements in the central region in the lower part of the column. Wen
& Yu’s model are much closer to the PIV data than the others. Moreover,
Rusche’s model and Lima Neto’s model show moderate performance. At the
height z/H = 0.7, however, Wen & Yu’s model apparently underestimates the
vertical liquid velocity in the central region of the column. In addition the
results from Holland’s model and Rusche’s model also show slightly different
profiles compared with the PIV data. The others agree with the data quite
well.
When looking at the comparisons of the relative velocity between the two
phases, one can see that Wen & Yu’s model and Lima Neto’s drag model
overestimate the relative velocity at all three heights in the bubble column.
The other models show similar trends for the relative velocity along the x
direction, particularly at the height z/H = 0.7. However, the profile of the
relative velocity in the upper part of the column is reproduced.

5.6 Conclusions

In this chapter, the performance of several drag coefficient correlations pro-


posed in literature has been tested by performing detailed Euler-Lagrange

107
5. Discrete bubble modeling of bubbly flows: Swarm effects

(a) Vertical liquid velocity (z/H = 0.3) (b) Relative velocity (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Relative velocity (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Relative velocity (z/H = 0.7)

Figure 5.6: Comparisons at superficial gas velocity u g = 0.0073 m/s.

108
5.6. Conclusions

(a) Vertical liquid velocity (z/H = 0.3) (b) Relative velocity (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Relative velocity (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Relative velocity (z/H = 0.7)

Figure 5.7: Comparisons at superficial gas velocity u g = 0.0097 m/s.

109
5. Discrete bubble modeling of bubbly flows: Swarm effects

simulations. In our study, we focused on the swarm effect with respect to


the effective drag experienced by a bubble rising in a swarm. The simulation
results were compared with PIV measurements to evaluate these correlations.
Three heights in the bubble column were considered: z/H = 0.3, 0.5 and 0.7.
The results show that Lima Neto’s drag model and Wen & Yu’s drag model
give better prediction of both the vertical liquid velocity and the relative ve-
locity between the gas phase and the liquid phase in the entire column at low
superficial gas velocity (i.e. u g = 0.0024 m/s). As the superficial gas velocity
increases, however, Lima Neto’s model tends to overestimate the relative ve-
locity considerably even though it still can predict the vertical liquid velocity
well. Wen & Yu’s drag model shows a large deviation with the prediction of
both the vertical liquid velocity and the relative velocity at high superficial
gas velocity (i.e. u g > 0.0049 m/s. Among the other models, Rusche’s model
and Holland’s model predict the vertical liquid velocity in the lower part of
the column better than the other three models at the moderate superficial gas
velocity (i.e. 0.0049 m/s ≤ u g ≤ 0.0073 m/s). At u g = 0.0097 m/s, however,
Rusche’s model gives a better performance with the prediction of the vertical
liquid velocity in the lower part of the column. In the higher part of the col-
umn, these five drag models have a similar performance with the prediction
of the vertical liquid velocity particularly at high superficial gas velocity. In
addition, the five models also have a similar trend on predicting the relative
velocity between the gas phase and the liquid phase. Meanwhile, it can be
seen that Rusche’s drag model predicts the relative velocity better at high
superficial gas velocity.
The present work has compared the performance of different drag closures
with the prediction of hydrodynamics of bubbly flows. We found that Lima
Neto’s drag model and Wen & Yu’s model have a better performance at low
superficial gas velocity. As the superficial gas velocity increases, these two
models tend to overestimate the relative velocity between the gas phase and
the liquid phase. It can also be seen that Rusche’s model can predict the hy-
drodynamics of the bubbly flows better compared to the other models at high
superficial gas velocity. However, there is also need for further improvement
of the drag coefficient correlation with swarm effects. Some factors important

110
Nomenclature

in bubbly flows as well need to be considered, i.e. the effect of bubble swarms
on the lift force. Meanwhile, coalescence and breakup of bubbles should also
be taken into account in the modelling in order to simulate the bubbly flows
reasonably at high void fraction flows.

Nomenclature

C model coefficient, [-]


CS Smagorinsky constant, [-]
d bubble diameter, [m]
D column depth, [m]
E bubble aspect ratio, [-]
(ρl − ρ g )gd2
Eo Eötvös number, Eo = , [-]
σ
F force vector, [N]
f function
g gravity acceleration, [m s−2 ]
H column height, [m]; horizontal direction
I unit tensor, [-]
n unit normal vector, [-]
p pressure, [N m−2 ]
r position vector, [m]
R bubble radius, [m]
ρl |v − u|d
Re bubble Reynolds number, Re = , [-]
µl
t time, [s]
u liquid velocity vector, [m s−1 ]
v bubble velocity vector, [m s−1 ]
V volume, m3 ; terminal velocity, [m s−1 ]
W column width, [m]

Greek letters

α void fraction, [-]


111
5. Discrete bubble modeling of bubbly flows: Swarm effects

∆ subgrid length scale, [m]


 gas holdup, [-]
µ viscosity, [kg m−1 s−1 ]
Φ volume averaged momentum transfer due to interphase forces,
[kg m−2 s−2 ]
ρ density, [kg m−3 ]
σ surface tension, [N m−1 ]
τ stress tensor, [N m−2 ]

Indices

b bubble
D drag
eff effective
g gas phase
G gravity
H horizontal direction
i i direction
j j direction
l liquid phase
L lift; molecular viscosity
m mixture
P pressure
s superficial velocity
T turbulence
V vertical direction
VM virtual mass
W wall
∞ infinite medium

112
References

Acknowledgement

The authors would like to gratefully acknowledge Dr. Lima Neto for sharing
his experimental results.

References

T. R. Auton. The lift force on a spherical body in a rotational flow. J. Fluid


Mech., 197:241–257, 1987.

A. Behzadi, R. I. Issa, and H. Rusche. Modelling of dispersed bubble and


droplet flow at high phase fractions. Chemical Engineering Science, 59(4):
759–770, 2004.

F. Bertola, G. Baldi, D. Marchisio, and M. Vanni. Momentum transfer in a


swarm of bubbles: estimates from fluid-dynamic simulations. Chemical
Engineering Science, 59(22-23):5209–5215, 2004. ISCRE18.

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

D. Darmana, N. G. Deen, and J. A. M. Kuipers. Parallelization of an Euler-


Lagrange model using mixed domain decomposition and a mirror domain
technique: Application to dispersed gas-liquid two-phase flow. Journal of
Computational Physics, 220(1):216–248, 2006.

N. G. Deen. An experimental and computational study of fluid dynamics in gas-


liquid chemical reactors. PhD thesis, Aalborg University, Denmark, 2001.

E. Delnoij, F. A. Lammers, J. A. M. Kuipers, and W. P. M. van Swaaij. Dynamic


simulation of dispersed gas-liquid two-phase flow using a discrete bubble
model. Chemical Engineering Science, 52(9):1429–1458, 1997.

E. Delnoij, J. A. M. Kuipers, and W. P. M. van Swaaij. A three-dimensional


CFD model for gas-liquid bubble columns. Chemical Engineering Science, 54
(13-14):2217–2226, 1999.

113
5. Discrete bubble modeling of bubbly flows: Swarm effects

C. Garnier, M. Lance, and J. L. Marié. Measurement of local flow characteristics


in buoyancy-driven bubbly flow at high void fraction. Experimental Thermal
and Fluid Science, 26(6-7):811–815, 2002.

M. Gentric, C. Olmos, E. Midoux, and N. Simonnet. Cfd simulation of the flow


field in a bubble column reactor: Importance of the drag force formulation
to describe regime transitions. Chemical Engineering and Processing: Process
Intensification, 47(9-10):1726–1737, 2008.

F. A. Holland and R. Bragg. Fluid Flow for Chemical Engineers. Elsevier Ltd.,
second edition, 1995.

B. P. B. Hoomans, J. A. M. Kuipers, W. J. Briels, and W. P. M. van Swaaij. Dis-


crete particle simulation of bubble and slug formation in a two-dimensional
gas-fluidised bed: A hard-sphere approach. Chemical Engineering Science,
51(1):99–118, 1996.

M. Ishii and T. Hibiki. Thermo-fluid dynamics of two-phase flow. Springer, 2005.

M. Ishii and N. Zuber. Drag coefficient and relative velocity in bubbly, droplet
or particulate flows. AIChE Journal, 25(5):843–855, 1979.

H. A. Jakobsen. Chemical Reactor Modeling: Multiphase Reactive Flows. Springer,


1st edition, July 2008.

M. Jamialahmadi, C. Branch, and H. Muller-Steinhagen. Terminal bubble


rise velocity in liquids. Chemical Engineering Research and Design, 72(A1):
119–122, 1994.

A. Lübbert A. Lapin. Numerical simulation of the dynamics of two-phase


gas-liquid flows in bubble columns. Chemical Engineering Science, 49(21):
3661–3674, 1994.

I. E. Lima Neto, D. Z. Zhu, and N. Rajaratnam. Bubbly jets in stagnant water.


International Journal of Multiphase Flow, 34(12):1130–1141, 2008.

K. E. Morud and B. H. Hjertager. LDA measurements and CFD modelling


of gas-liquid flow in a stirred vessel. Chemical Engineering Science, 51(2):
233–249, 1996.

114
References

H. Rusche. Computational fluid dynamics of dispersed two-phase flows at high phase


fractions. PhD thesis, Imperial College, London, 2002.

M. Simonnet, C. Gentric, E. Olmos, and N. Midoux. Experimental determi-


nation of the drag coefficient in a swarm of bubbles. Chemical Engineering
Science, 62(3):858–866, 2007.

M. Sommerfeld, E. Bourloutski, and D. Bröder. Euler/Lagrange calculations


of bubbly flows with consideration of bubble coalescence. Canadian Journal
of Chemical Engineering, 81(3-4):508–518, 2003.

A. Tomiyama, T. Matsuoka, T. Fukuda, and T. Sakaguchi. A simple numerical


method for solving an incompressible two-fluid model in a general curvi-
linear coordinate system. Advances in Multiphase Flow, pages 241–252,
Amsterdam, 1995. Society of Petroleum Engineers Inc.

A. Tomiyama, I. Kataoka, I. Zun, and T. Sakaguchi. Drag coefficients of single


bubbles under normal and micro gravity conditions. JSME international
journal. Ser. B, Fluids and thermal engineering, 41(2):472–479, 1998.

A. Tomiyama, H. Tamai, I. Zun, and S. Hosokawa. Transverse migration of


single bubbles in simple shear flows. Chemical Engineering Science, 57(11):
1849–1858, 2002.

J. A. Trapp. A discrete particle model for bubble-slug two-phase flows. Journal


of Computational Physics, 107(2):367–377, 1993.

A. W. Vreman. An eddy-viscosity subgrid-scale model for turbulent shear


flow: Algebraic theory and applications. Physics of Fluids, 16(10):3670–3681,
2004.

R. M. Wellek, A. K. Agrawal, and A. H. P. Skelland. Shape of liquid drops


moving in liquid media. AIChE Journal, 12(5):854–862, 1966.

C. Y. Wen and Y. H. Yu. Mechanics of fluidization. Chemical Engineering


Progress Symposium Series, 62:100–111, 1966.

115
5. Discrete bubble modeling of bubbly flows: Swarm effects

I. Zun and J. Grošelj. The structure of bubble non-equilibrium movement in


free-rise and agitated-rise conditions. Nuclear Engineering and Design, 163
(1-2):99–115, 1996.

116
Chapter
6
Discrete bubble modeling of
bubbly flows: Implementation of
breakup models
Breakup models developed in literature were implemented in our model, which is
based on an Eulerian-Lagrangian approach. Moreover, the critical Weber number for
bubble breakup utilized by many authors in turbulent flows was also incorporated in
the model. Only binary breakage is considered in this work. For the models utilizing
the critical Weber number, two different daughter size distributions, namely bell shape
and a U shape were used.
First the implementation was verified by simulating bubbly flows in a square bubble
column. The simulated breakup frequency and daughter size distribution were com-
pared with those obtained from the models. Subsequently, the simulation results were
compared with detailed PIV measurements. Finally, the predicted bubble size distri-
butions were compared with chord length distributions obtained from measurements
with a four-point optical fibre probe.

117
6. Implementation of breakup models

6.1 Introduction

In two-phase gas-liquid flows, the properties of the dispersed phase are very
important (i.e. interface topology). The complex interface topology and its
dynamics (due to coalescence and breakup) poses a considerable difficulty in
the study of bubbly flows. Hence, mechanisms of coalescence and breakup
are of special significance for the investigation of two-phase gas-liquid flows.
The deformation and breakup of a bubble can be described with a sim-
ple static force balance, which introduces the ratio between the force that
causes the deformation and the surface tension that tends to counteract the
deformation (Hinze, 1955). Depending on the type of flow, the cause of the
deformation varies. No matter what the nature of the deformation is, how-
ever, the breakup occurs when the ratio exceeds a critical value. The ratio
is expressed as a critical Weber number. Sevik and Park (1973) studied the
critical Weber number for air bubbles in a high Reynolds number water jet
(Kolev, 2007). Walter and Blanch (1986) proposed an expression for the maxi-
mum stable bubble size in solutions and conducted a comprehensive study on
the effect of the presence of surfactants on bubble breakup. Risso and Fabre
(1998) analyzed the breakup mechanism of a bubble in turbulent flows under
microgravity conditions and estimated the critical Weber number. Qian et al.
(2006) studied the critical Weber number of bubbles in homogeneous turbu-
lence under zero gravity conditions using the lattice Boltzmann method. A
detailed review of the mechanisms of deformation and breakup of drops and
bubbles has been given by Risso (2000).
In recent years, CFD has emerged as a powerful tool to study multiphase
flow phenomena and assess their impact on the performance of multiphase
chemical reactors (Jakobsen, 2008). In particular population balance mod-
elling has received considerable attention in the past decade to account for
the size distribution of the dispersed phase. In order to close the population
balance equations, i.e. through expressions for the birth rate and the death
rate due to breakup of bubbles, many efforts on the mathematical formulation
of coalescence and breakup models in turbulent flows have been made since
then. Based on the turbulent nature of liquid-liquid dispersion, a phenomeno-

118
6.2. Discrete bubble model

logical model was developed to describe drop breakup by Coulaloglou and


Tavlarides (1977). The basic premise is that an oscillating deformed drop will
break if the turbulent kinetic energy transmitted to the drop by turbulent ed-
dies exceeds the drop surface energy. The breakup frequency of a drop of size
d was defined as:
!
1 
Ω(d) = fraction of drops breaking (6.1)
Breakup time

Other formulations of breakup mechanisms in turbulent flows have been


presented by Mihail and Straja (1986), Lee et al. (1987), Prince and Blanch
(1990), Tsouris and Tavlarides (1994), Luo and Svendsen (1996), Sathyagal
et al. (1996), Kostoglou et al. (1997), Martı́nez-Bazán et al. (1999a), Martı́nez-
Bazán et al. (1999b), Hagesaether et al. (2002), Lehr et al. (2002), Wang et al.
(2003) and Kostoglou and Karabelas (2005). Moreover, detailed reviews about
breakage models can be found in Lasheras et al. (2002) and Liao and Lucas
(2009).
Population balance modelling coupled to the two-fluid approach has been
frequently used to study the multiphase flows (Lehr and Mewes, 2001; Olmos
et al., 2001; Chen et al., 2005; Wang et al., 2006; Bannari et al., 2008). However
for the Eulerian-Lagrangian approach, there are barely studies on hydrody-
namics of two-phase gas-liquid flows with breakup models integrated into
the model.
In the present work, an attempt has been made to integrate two breakup
models from literature into an Eulerian-Lagrangian (EL) framework. In addi-
tion, an expression for the critical Weber number for bubble breakup reported
in literature has been adopted. The simulation results were compared with
PIV measurements to evaluate the performance of the EL model incorporating
coalescence and breakup.

6.2 Discrete bubble model

Our discrete bubble model (DBM) was originally developed by Delnoij et al.
(1997, 1999). The liquid phase is described by volume-averaged continuity

119
6. Implementation of breakup models

and Navier-Stokes equations, while the motion of each individual bubble is


tracked in a Lagrangian fashion.

6.2.1 Bubble dynamics


The motion of each individual bubble is computed from Newton’s second
law. For an individual bubble, the equation of motion can be written as:
dv
ρg V = ΣF (6.2)
dt

dr
=v (6.3)
dt
The net force acting on each individual bubble is calculated by considering
all the relevant forces. It is assumed that the net force is composed of separate,
uncoupled contributions due to gravity, pressure, drag, lift, virtual mass and
wall force respectively:

ΣF = FG + FP + FD + FL + FVM + FW (6.4)

The gravity force acting on a bubble in a liquid is given by:

FG = ρ g Vg (6.5)

The far field pressure force incorporating contributions of the Archimedes


buoyancy force, inertial forces and viscous strain is given by:

FP = −V∇P (6.6)

The drag force exerted on a bubble rising in a liquid is expressed as:


1
FD = − CD ρl πd2 |v − u|(v − u) (6.7)
8
The drag coefficient CD is determined by taking swarm effect into account.
According to Rusche (2002), the drag coefficient in a swarm of bubble can be
expressed as the product of the drag coefficient for an isolated bubble rising
in an infinite quiescent liquid and a correction coefficient:

CD = CD∞ f (α g ) (6.8)

120
6.2. Discrete bubble model

where f (α g ) represents the effect arising from the presence of other bubbles:

f (α g ) = exp(3.64α g ) + α0.864
g (6.9)

The drag coefficient for an isolated bubble CD∞ is calculated according to


Tomiyama et al. (1998):
   
16 48 8 Eo
CD∞ = max min (1 + 0.15Re0.687 ), , (6.10)
Re Re 3 Eo + 4
A bubble rising in a non-uniform liquid flow field experiences a lift force
due to vorticity or shear in the flow field. The shear induced lift force acting
on a bubble is usually written as (Auton, 1987):

FL = −CL ρl V(v − u) × ∇ × u (6.11)

The lift coefficient is calculated according to Tomiyama et al. (2002):




 min[0.288tanh(0.121Re), f (EoH )] EoH < 4,




CL = 
 f (EoH ) 4 ≤ EoH ≤ 10, (6.12)



−0.29 EoH > 10.

where
f (EoH ) = 0.00105EoH 3 − 0.0159EoH 2 + 0.474 (6.13)

EoH is the Eötvös number defined by using the maximum horizontal dimen-
sion of a bubble:
(ρl − ρ g )gd2H
EoH = (6.14)
σ
The maximum horizontal diameter of the bubble is obtained from the
bubble aspect ratio E according to Wellek et al. (1966):

dV 1
E= = (6.15)
dH 1 + 0.163Eo0.757
where dV is the maximum vertical diameter of the bubble and Eo is the Eötvös
(ρl − ρ g )gd2
number, Eo = .
σ

121
6. Implementation of breakup models

The relation between the above two diameters and the diameter of the
bubble d in the discrete bubble modeling is as follows:

d = (dV d2H )1/3 (6.16)

Accelerating bubbles experience a resistance, which is described as the


virtual mass force (Auton, 1987):
 
Dv Du
FVM = −CVM ρl V − (6.17)
Dt Dt
where the D/Dt operators denote the substantiative derivatives pertaining to
the respective phase. In the present work, bubbles are assumed to have a
spherical shape and a virtual mass coefficient of 0.5 is used.
Bubbles in the vicinity of a solid wall experience a force referred to as the
wall force (Tomiyama et al., 1995):
" #
1 1 1
FW = − C W d 2 − ρl |(v − u)· nz |2 nW (6.18)
2 y (L − y)2

where nz and nW , respectively, are the normal unit vectors in the vertical and
wall normal direction, L is the dimension of the system in the wall normal
direction, and y is the distance to the wall in that direction. Finally, the wall
force coefficient CW is given by:



exp(−0.933Eo + 0.179) 1 ≤ Eo ≤ 5,
CW =
 (6.19)
0.007Eo + 0.04 5 < Eo ≤ 33.

6.2.2 Liquid phase dynamics


The presence of the bubbles is reflected by the liquid phase volume fraction
αl and the interphase momentum transfer rate Φ in the volume-averaged
conservation equations:

∂  
αl ρl + ∇ · αl ρl u = 0 (6.20)
∂t
∂  
αl ρl u + ∇ · αl ρl uu = −αl ∇p − ∇ · (αl τl ) + αl ρl g + Φ (6.21)
∂t

122
6.3. Coalescence model

The liquid phase is assumed to be Newtonian, thus the stress tensor τl can be
expressed as:
 
2
τl = −µeff,l ((∇u) + (∇u)T − I(∇ · u) (6.22)
3
where µeff,l is the effective viscosity. In the present model, the effective viscos-
ity is composed of two contributions, the molecular viscosity and the turbulent
viscosity:
µeff,l = µL,l + µT,l (6.23)

where µT,l is the turbulent viscosity (or eddy viscosity), which is determined
from turbulence modeling of the liquid phase.

6.2.3 Collision model


In the present paper, a hard sphere collision model (Hoomans et al., 1996)
is adopted to describe the bouncing of bubbles. It consists of two main
parts; i) processing the sequence of collisions and ii) dealing with the collision
dynamics. The former is described in detail by Darmana et al. (2006). This
approach is also used in this work.

6.3 Coalescence model

Bubble coalescence is considered by comparing the film drainage time and


the contact time between two bubbles (Prince and Blanch, 1990). The film
drainage time can be obtained with a simple expression:
r
Rij ρl h0
tij = ln (6.24)
16σ h f

where h0 is the initial film thickness which is assumed to be 1 × 10−4 m for an


air-water system. The critical film thickness where rupture occurs is given as
1 × 10−8 m. The equivalent bubble radius for two different sized bubbles is
given by Hofman and Chesters (1982):
!−1
1 1 1
Rij = + (6.25)
2 Ri R j

123
6. Implementation of breakup models

The contact time is dependent on the bubble size and the turbulent intensity.
An estimate of the contact time in turbulent flows is given as:

R2/3
ij
τij = (6.26)
1/3
where  is the turbulent energy dissipation rate.
Sommerfeld et al. (2003) provided an estimate of the contact time by as-
suming that it is proportional to a deformation distance divided by the normal
component of the relative velocity:
CC Rij
τij = (6.27)
|vni − vnj |
where CC is a model constant.
The properties of the new bubble in case of coalescence are obtained from
conservation of mass and momentum.

6.4 Breakup models & implementation

6.4.1 Breakup models


Breakup frequency

According to Martı́nez-Bazán et al. (1999a), the basic premise for a bubble to


break is that its surface has to deform, and enough energy must be provided
by the turbulent stresses of the surrounding liquid. They postulated that the
breakup frequency is proportional to the difference between the turbulent
stresses 21 ρl β(d)2/3 and the surface pressure 6σ/d. In other words, the larger
the difference is, the larger the probability that the bubble will break in a
certain time. On the other hand, the breakup frequency should decrease to
zero in case this difference vanishes. Thus, the breakup frequency is estimated
as: s
0.25 12σ
Ω(d) = β(d)2/3 − (6.28)
d ρl d
where the constant β = 8.2 was given by Batchelor (1953) and the constant
0.25 was found experimentally by Martı́nez-Bazán et al. (1999b).

124
6.4. Breakup models & implementation

The dependence of the breakup frequency on the bubble diameter is shown


in Figure 6.1(a). The breakup frequency possesses a maximum as the bubble
diameter increases and after reaching the maximum, the breakup frequency
decreases monotonically with increasing bubble size.

25 2e+04
ε=0.5 m2/s3 ε=0.5 m2/s3
ε=1.0 m2/s3 ε=1.0 m2/s3
20 ε=2.0 m2/s3 ε=2.0 m2/s3
1.5e+04
Ω(d) [1/s]

Ω(d) [1/s]
15
1e+04
10

5,000
5

0 0
0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08 0.1
d [m] d [m]

(a) Martı́nez-Bazán et al. (1999a) (b) Lehr et al. (2002)


Figure 6.1: Breakup frequency vs bubble diameter.

Lehr et al. (2002) assumed that the breakup of a bubble is determined


by the balance between the interfacial force acting at the bubble surface and
the inertial force of an impinging eddy. The interfacial force depends on the
shape of the bubble and on the size of daughter bubbles. They computed
the breakup probability based on the criterion that the kinetic energy of the
eddy exceeds a critical energy which is obtained from the force balance. The
breakup frequency is then expressed as:

 √ 
d5/3 19/15 ρ7/5  2σ9/5 
Ω(d) = 0.5 l 
exp −  (6.29)
σ7/5 d3 ρ9/5 6/5  l

As shown in the Figure 6.1(b), the breakup frequency increases monoton-


ically with increasing bubble diameter and there is no maximum, which is
different from that of Martı́nez-Bazán et al. (1999a).

125
6. Implementation of breakup models

Daughter size distribution

By assuming that a bubble only breaks into two bubbles, Martı́nez-Bazán et al.
(1999b) postulated that the probability of the formation of a pair of bubbles
of sizes d1 and d2 from a mother bubble of size d, P(d∗ ), is weighted by the
product of two surplus stresses which are associated with the formation of the
two daughter bubbles. Hence, the daughter size distribution can be written
as:
P(d∗ )
β(d∗ ) = R 1 (6.30)
P(d∗ ) dd∗
0

and
!" #
∗ ∗2/3 12σ ∗3 2/9 12σ
P(d ) = d − (1 − d ) − (6.31)
8.2ρl d5/3 2/3 8.2ρl d5/3 2/3

where d∗ = d1 /d. Note that the dimensionless diameter d∗ can easily be


converted to a dimensionless volume V ∗ : V ∗ = d∗3 . Hence, for the sake of
clarity, only the dimensionless volume V ∗ is considered.
The daughter size distribution according to Martı́nez-Bazán et al. (1999b)
is plotted in Figure 6.2(a). One can see that the daughter size distribution has
a bell shape. In other words, it is more likely for a mother bubble to break into
two bubbles with equal volume. Moreover the distribution becomes wider as
the mother bubble size increases.
Lehr et al. (2002) described the daughter size distribution as a lognormal
distribution which gives:
    

   3/5 2/5 2 
 

 9 
ln  22/5
d1 ρl
  
exp  − 
 
  


 4 σ 3/5 


β(V ) =     (6.32)
√ 
  3  21/15 dρ3/5  2/5 
 
 
1 + erf  ln  
l
V∗ π  
 2 σ3/5 

where V ∗ = V1 /V.
As shown in Figure 6.2(b), unequal-size breakage is more likely according
to the above daughter size distribution. And the probability of unequal-sized
breakage increases rapidly with the mother bubble size.

126
6.4. Breakup models & implementation

3 25
d=0.005 m d=0.005 m
d=0.007 m d=0.007 m
d=0.01 m 20 d=0.01 m

2
15
β(V*)

β(V*)
10
1
5

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
*
V [-] V* [-]

(a) Martı́nez-Bazán et al. (1999b) (b) Lehr et al. (2002)


Figure 6.2: Daughter size distribution.

6.4.2 Critical Weber number


When a bubble moves in a turbulent flow, the turbulent eddies deform its
surface and may eventually cause its breakup. According to Hinze (1955),
only the eddies with length scale smaller than the bubble size are capable to
break the bubble while larger eddies merely transport the bubble. This leads
to a quantitative description of turbulence induced bubble breakup. The
turbulent shear stress imposed by the continuous phase deforms the bubble
and breaks it if it overcomes the surface tension force. The relevant Weber
number is thus defined as:
ρl u02 d
We = (6.33)
σ
By assuming that the turbulence is isotropic at the bubble scale and the bubble
diameter belongs to the inertial turbulent subrange, the Weber number is
expressed as a function of the dissipation rate  (Risso, 2000):

2ρl 2/3 d5/3


We = (6.34)
σ
For the breakup of a gas jet in a turbulent liquid stream, the critical Weber
number is given as We = 2.48 (Kolev, 2007). By investigating the behavior
of bubbles of different diameters under microgravity conditions, Risso and
Fabre (1998) found that the critical Weber number ranges between 2.7 and

127
6. Implementation of breakup models

7.8. They also obtained a value of the critical Weber number close to 4.5 by
analyzing how the mean deformation is related to the mean turbulence in-
tensity. Hence, they suggested that a minimal Weber number of about 5 is
necessary for breakup when the turbulence is the only cause of deformation.
Qian et al. (2006) performed lattice Boltzmann simulations for the deforma-
tion and breakup of bubbles in homogeneous turbulence under zero gravity
conditions. The minimum Weber number for bubble breakup was found to
be about 3.0.
The beta distribution can be used to describe the daughter bubble size
when a critical Weber number is involved to determine bubble breakup in the
turbulent flows. The beta distribution is formulated as follows:
Γ(a + b) ∗a−1
f (V ∗ ) = V (1 − V ∗ )b−1 (6.35)
Γ(a)Γ(b)

where Γ is the gamma function and a and b shape parameters.

2.5 a=b=0.5
a=b=2.0
2

1.5
f(V*)

0.5

0
0 0.2 0.4 0.6 0.8 1
V* [-]

Figure 6.3: Beta distributions.

In order to obtain different daughter size distributions, two different sets


of shape parameters of the beta distribution, a = b = 0.5 and a = b = 2.0,
were used. Hence, two different shapes of the beta distributions, such as
bell shape and U shape, can be obtained, as shown in Figure 6.3. By using

128
6.5. Simulation details

different daughter size distributions, one can get equal-sized breakage of a


mother bubble with either the highest probability or the lowest probability.

6.4.3 Implementation
The above breakup models provide quantitative information on dynamics
of bubble breakage in turbulent flows. For instance, the breakup frequency
is related to the time interval in which the bubble breakage occurs. Hence,
one may find a way to integrate the bubble breakage into the discrete bubble
model (DBM) by taking the breakup time into account.
In the discrete bubble model, each bubble is tracked individually in the
turbulent flow field. The bubble properties, such as bubble velocity, bubble
diameter and bubble position in the flow, and local flow properties, i.e. dis-
sipation rate of turbulence kinetic energy, are known. Hence, it is possible
to determine the breakup time for each bubble at its position in the flow. If
the breakup time of a bubble is smaller than the computational time step, the
bubble breaks immediately within that time step. On the other hand, if the
calculated breakup time of the bubble exceeds the computational time step,
the program starts counting the time since that instant. The computation
moves to the next time step and once the counted time exceeds the breakup
time, the bubble breaks. In this approach, the history of turbulent eddies
colliding with the bubble is considered.
The resulting daughter bubble sizes are determined according to the above
introduced daughter size distributions. The velocities of the two daughter
bubbles are assumed to be the same as that for the mother bubble. One of the
two daughter bubbles is located at the same position as the mother bubble
while the other daughter bubble is located around that daughter bubble ran-
domly. After breakage, each bubble in the turbulent flow field is then tracked
by considering the breakage criterion again.

6.5 Simulation details

The bubble column studied here is shown in Figure 6.4. The cross-sectional
area of the column is 0.15 m × 0.15 m (W × D). The column is initially filled

129
6. Implementation of breakup models

with water to a height of 0.45 m (H). Air is used as the dispersed phase and
introduced into the column through a perforated plate with 49 holes (7 × 7) at
the bottom of the column. The material properties of both phases are taken at
room temperature and atmospheric pressure.

Figure 6.4: Sketch of the bubble column.

A computational grid with (20×20×60) cells is adopted in the simulations.


Three superficial gas velocities, i.e. 0.0049, 0.0073 and 0.0097 m/s were used.
The initial (i.e. at the distributor) bubble diameter in the simulations is 0.005 m,
which is consistent with experimental observations in a bubble column at
the applied superficial gas velocity. The computational time step for the

130
6.6. Results and discussion

liquid phase is set at 1.0 × 10−3 s and the collisions among bubbles and the
movements of the bubbles are processed multiple times within each time
step. The total simulation time is 150 s.
The coalescence model and two breakup models described above are used
in the simulations. Moreover, two critical Weber numbers, i.e. We = 2.48 and
We = 5.0, are adopted for the purpose of comparison.
The sub-grid scale model is required to represent the turbulent flow in-
side the bubble column. This means that the continuity and Navier-Stokes
equations are resolved for the “large” eddies, while the ”small scales” of the
resolved filed are included through an eddy-viscosity subgrid-scale model.
The eddy-viscosity model proposed by Vreman (2004) was used to calcu-
late the eddy viscosity:
s

µT,l = 2.5ρl C2S (6.36)
αij αij

where Bβ = β11 β22 − β212 + β11 β33 − β213 + β22 β33 − β223 , βij = ∆2m αmi αm j and
αij = ∂u j /∂xi . ∆i is the filter width in the i direction.
In the large eddy simulations, however, the dissipation rate of turbulent
kinetic energy cannot be obtained directly. Moreover, the dissipation rate only
from the resolved scales is negligible compared with that from the subgrid
scales. By assuming a local equilibrium between turbulent kinetic energy
production and dissipation, the dissipation rate can be estimated as (Jiménez
et al., 2001; Hartmann et al., 2004; Delafosse et al., 2009):
!2
(µL,l + µT,l ) ∂ui ∂u j
= + (6.37)
2ρl ∂x j ∂xi

6.6 Results and discussion

6.6.1 Validation of the implementation


Before discussing the performance of different breakup models with respect
to prediction of bubbly flow characteristics, it is necessary to first verify the
implementation of the models, such as, breakup frequency and daughter size
distribution. In Figure 6.5, the breakup frequency of each breaking bubble

131
6. Implementation of breakup models

with size d at a certain dissipation rate in a numerical simulation is compared


with the breakup frequency curve calculated from the corresponding breakup
model at the same dissipation rate. Three different energy dissipation rates,
 = 0.5, 1.0 and 2.0 m2 /s3 , are chosen for the comparison. One can clearly see
that the breakup frequencies obtained from the numerical simulation agree
very well with the curve calculated from the corresponding breakup model.
This indicates that the breakup frequency models have been implemented
correctly.

40 600
Model (ε=0.5 m2/s3) Model (ε=0.5 m2/s3)
Model (ε=1.0 m2/s3) Model (ε=1.0 m2/s3)
Model (ε=2.0 m2/s3) Model (ε=2.0 m2/s3)
30
DBM (ε=0.5 m2/s3) DBM (ε=0.5 m2/s3)
400
Ω(d) [1/s]

Ω(d) [1/s]
DBM (ε=1.0 m2/s3) DBM (ε=1.0 m2/s3)
DBM (ε=2.0 m2/s3) DBM (ε=2.0 m2/s3)
20

200
10

0 0
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015
d [m] d [m]

(a) Martı́nez-Bazán et al. (1999a) (b) Lehr et al. (2002)


Figure 6.5: Verification of implementation for the breakup frequency.

Moreover, the daughter size distribution obtained from the numerical


simulations is also compared with the probability distributions of bubble
sizes obtained from the breakup model (see Figure 6.6). It can be seen that the
bubble size distribution obtained from the numerical simulations reproduces
the trend of the corresponding model very well. In addition for the cases
with critical Weber numbers, the bell-shaped and U-shaped daughter size
distributions are also reproduced well.

6.6.2 Comparison of different breakup models


In order to evaluate the performance of the above breakup models and the
critical Weber numbers for bubble breakup, the coalescence model in combi-
nation with the hydrodynamic model, averaged quantities, such as vertical

132
6.6. Results and discussion

25
3 Model (d=0.005 m) DBM (d=0.005 m) Model (d=0.005 m)
Model (d=0.007 m) DBM (d=0.007 m) Model (d=0.007 m)
Model (d=0.01 m) DBM (d=0.01 m) 20 Model (d=0.01 m)
DBM (d=0.005 m)
2 DBM (d=0.007 m)
15

β(V*)
β(V*)

DBM (d=0.01 m)

10
1
5

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
*
V [-] V* [-]

(a) Martı́nez-Bazán et al. (1999b) (b) Lehr et al. (2002)


0.08
0.15

0.06

0.1
β(V*)
β(V*)

0.04

0.02
0.05

0
0 0.2 0.4 0.6 0.8 1 0 1
V* [-] V* [-]

(c) Bell-shaped distribution (d) U-shaped distribution

Figure 6.6: Verification of implementation for the daughter size distribution.

liquid velocity and relative velocity between the gas phase and the liquid
phase, will be considered. Comparisons are carried out at three heights in
the column (z/H = 0.3, 0.5 and 0.7). Moreover, simulations assuming uni-
form bubble size with and without considering swarm effects are included for
the purpose of reference. In the following figures, the symbols representing
different conditions of simulations are kept the same.

u g = 0.0049 m/s

The average vertical liquid velocity and the local void fraction obtained from
the simulations with superficial gas velocity u g = 0.0049 m/s are plotted in

133
6. Implementation of breakup models

Figure 6.7. It can be seen that the value of the critical Weber number influences
the vertical liquid velocity considerably, particularly in the central region of
the bubble column. Assuming the same daughter size distribution, i.e. the
bell-shaped distribution, the vertical liquid velocities predicted with We = 5.0
are higher than those predicted with We = 2.48. Furthermore, assuming the
same critical Weber number, i.e. We = 2.48, but with different daughter size
distributions, i.e. bell shape and U shape, the difference between the predicted
and experimental vertical liquid velocities in the central region is getting larger
along column height. Martı́nez-Bazán’s breakup model produces similar per-
formance regarding the prediction of the vertical liquid velocity assuming the
critical Weber number We = 2.48 and a bell-shaped daughter size distribu-
tion. Lehr’s breakup model produces a relatively large deviation. For the
case of uniform bubble size with a single bubble drag closure also a large
discrepancy with the experiments is obtained. However, it can be seen that at
the relatively low superficial gas velocity, i.e. u g = 0.0049 m/s, application of
a uniform bubble size considering the swarm effects on the drag force gives
better predictions of the vertical liquid velocity in the central region of the
column.
Comparison of local void fraction reveals that the local void fractions
predicted with the critical Weber number, We = 5.0, in the center is larger
than those predicted with other models. The predicted local void fractions
assuming an uniform bubble size considering the swarm effect and critical
Weber number, We = 2.48, with bell-shaped daughter size distribution are
close to each other.

u g = 0.0073 m/s

In Figure 6.8, profiles of the vertical liquid velocity and the local void fraction
at three heights in the bubble column are compared. The simulation with a
critical Weber number, We = 5.0, again shows the largest discrepancy with re-
spect to the prediction of the vertical liquid velocity in the center. In the central
region, Martı́nez-Bazán’s model and Lehr’s model have similar performance
compared to the case with the uniform size and single bubble drag closure in
the lower part of the column (z/H ≤ 0.5). Moreover, the models with a critical

134
6.6. Results and discussion

(a) Vertical liquid velocity (z/H = 0.3) (b) Local void fraction (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Local void fraction (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Local void fraction (z/H = 0.7)

Figure 6.7: Comparisons at superficial gas velocity u g = 0.0049 m/s.

135
6. Implementation of breakup models

Weber number, We = 2.48, predict better the liquid velocity in the lower part
of the column. However, they underestimate the liquid velocity in the higher
parts of the column. The model with uniform bubble size considering the
swarm effect predicts the liquid velocity quite well in the bubble column.
Profiles of the local void fraction show that the model with the critical
Weber number, We = 5.0, gives the highest void fraction in the central region
of the column. Furthermore, the local void fractions obtained from the models
with the critical Weber number, We = 2.48, are close to those predicted with
the model with uniform bubble size considering the swarm effect in the lower
part of the column (z/H ≤ 0.5). However, they start to deviate from each other
along the column height.

u g = 0.0097 m/s

At large superficial gas velocity, u g = 0.0097 m/s, profiles of the vertical liquid
velocity and the local void fraction obtained from the above models (Fig-
ure 6.9) show some differences compared with low superficial gas velocity.
It can be seen that the model with the critical Weber number, We = 5.0, can
predict the vertical liquid velocity quite well in the central region of the bubble
column. On the contrary, the model with critical Weber number, We = 2.48
and bell-shaped daughter size distribution and the model with uniform bub-
ble size considering swarm effects produce large deviations in the lower part
of the column. However, the former predicts the profile of the vertical liquid
velocity in the higher part of the column (i.e. z/H = 0.7) well. The model with
uniform bubble size and single bubble drag closure has the largest deviation
however in the higher part of the column this model performs quite well.
Martı́nez-Bazán’s model, Lehr’s model and the model with the critical Weber
number, We = 2.48 and U-shaped daughter size distribution show similar
performance in predicting the liquid velocity in the lower part of the bubble
column (z/H ≤ 0.5). With increasing column height, however, these models
show a worse performance.
In addition, one can see that the local void fractions obtained from the
model with uniform bubble size are slightly higher than those obtained from
the other models in the central region. In the lower part of the column, the

136
6.6. Results and discussion

(a) Vertical liquid velocity (z/H = 0.3) (b) Local void fraction (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Local void fraction (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Local void fraction (z/H = 0.7)

Figure 6.8: Comparisons at superficial gas velocity u g = 0.0073 m/s.

137
6. Implementation of breakup models

model with uniform bubble size and swarm effects predicts higher values of
the local void fraction than the others in the central region. The model that
uses the critical Weber number We = 2.48 and the U-shaped daughter size
distribution has similar performance as Martı́nez-Bazán’s model. Along the
column height, differences between the models are getting more pronounced.

(a) Vertical liquid velocity (z/H = 0.3) (b) Local void fraction (z/H = 0.3)

(c) Vertical liquid velocity (z/H = 0.5) (d) Local void fraction (z/H = 0.5)

(e) Vertical liquid velocity (z/H = 0.7) (f) Local void fraction (z/H = 0.7)

Figure 6.9: Comparisons at superficial gas velocity u g = 0.0097 m/s.

138
6.6. Results and discussion

6.6.3 Bubble size distribution


In order to evaluate the performance of the breakup models in combina-
tion with a coalescence model with respect to the prediction of bubble
size, the bubble size distributions at two positions in the bubble column
(z/H = 0.55, x/W = 0.5 and z/H = 0.75, x/W = 0.5) are compared with those
obtained from a four-point optical fibre probe at corresponding positions. In
Chapter 3, we introduced Uga (1972) and other authors’ work on the deter-
mination of the bubble size from a measured chord length distribution. For
instance, by assuming that all bubbles are spherical and travel in the same di-
rection with the same average velocity, Uga (1972) determined the equivalent
bubble diameter de from the mean chord length as:
Z ∞ Z ∞
3 3
de = d f 0 (d) dd = dc = dc f (dc ) ddc (6.38)
0 2 2 0

where dc is the chord length.


We hereby assume that the probability density function of the bubble size
is the same as that of the chord length, that is, f 0 (d) = f (dc ). Therefore, the
bubble size can be estimated as d = 1.5dc roughly and the bubble size distribu-
tion obtained from the four-point optical fibre probe can then be plotted in the
Figure 6.10. The estimated bubble size distributions are stretched to the right
side from the chord length distributions in the plots. The bubble size distri-
butions also show that the bubble size d = 0.005 m has the highest probability
in the bubble column approximately. Moreover, due to the geometry of the
four-point optical fibre probe, those bubbles with sizes or the chord lengths
pierced by the probe smaller than 0.0015 m can not be detected. Hence, there
are no data points for those bubbles in the plots.
Subsequently, the bubble size distributions determined from the simula-
tion results are then compared with those estimated from the chord length
distributions using the measurements with the four-point optical fibre probe,
as shown in the Figure 6.11.
When comparing the distributions at larges bubble size, it is found that
the bubble size distributions predicted by the above models have very similar
trends with those obtained from the optical probe except for Martı́nez-Bazán’s

139
6. Implementation of breakup models

0.5 0.4
Chord length Chord length
Bubble diameter Bubble diameter
0.4
0.3
Distribution [-]

Distribution [-]
0.3
0.2
0.2

0.1
0.1

0 0
0 0.005 0.01 0.015 0.02 0 0.005 0.01 0.015 0.02
Dimension [m] Dimension [m]

(a) z/H = 0.55, x/W = 0.5 (b) z/H = 0.75, x/W = 0.5
Figure 6.10: Distributions of the chord length and the bubble size.

model. At the central part of the column, the bubble size distribution pre-
dicted by Martı́nez-Bazán’s model has higher probability than others at about
d = 0.007 m. At the higher part of the column, the distributions obtained
from the numerical simulations are slightly shifted to the larger bubble size
compared with that determined from the optical probe. That suggests that
the implemented coalescence and breakup models into the discrete bubble
model are able to predict the bubble size distribution reasonably well. Fur-
thermore, the model that uses a U-shaped daughter size distribution has a
high probability to obtain some very small bubbles, which suggests that the
implementation of the U-shaped daughter size distribution is successful.

6.7 Conclusions

In this work, coalescence and breakup models have been combined with a
detailed Euler-Lagrange (EL) model incorporating four-way coupling. The
implementation of the breakup models in the EL model is based on the con-
cept of breakup frequency (breakup time). By considering the breakup time
for bubbles in turbulent flows and the history of bubble and turbulent eddies
encounter, bubble breakup is incorporated into the discrete bubble model.
Only binary breakup is considered in the breakup models. The daughter size

140
6.7. Conclusions

0.5 0.5
We=2.48, bell shape Martinez-Bazan, 1999
We=2.48, U shape Lehr, 2002
Bubble size distribution

Bubble size distribution


We=5.0, bell shape Optical probe
0.4 Optical probe
0.4

0.3 0.3

0.2 0.2

0.1 0.1

0.0 0.0
0.000 0.005 0.010 0.015 0.020 0.025 0.000 0.005 0.010 0.015 0.020 0.025
d [m] d [m]

(a) z/H = 0.55, x/W = 0.5 (b) z/H = 0.55, x/W = 0.5

0.4 0.4
We=2.48, bell shape Martinez-Bazan, 1999
We=2.48, U shape Lehr, 2002
Bubble size distribution

Bubble size distribution


We=5.0, bell shape Optical probe
0.3 Optical probe 0.3

0.2 0.2

0.1 0.1

0.0 0.0
0.000 0.005 0.010 0.015 0.020 0.000 0.005 0.010 0.015 0.020
d [m] d [m]

(c) z/H = 0.75, x/W = 0.5 (d) z/H = 0.75, x/W = 0.5
Figure 6.11: Bubble size distribution.

distribution after bubble breakage is determined according to models avail-


able in literature. In addition, critical Weber numbers for bubble breakup in
turbulent flow reported in literature have been adopted as well. Two different
daughter size distributions, i.e. bell shaped and U shaped distribution, were
used in this case.
The correct incorporation of the breakup models is first verified by com-
paring the breakup frequency and the daughter size distribution obtained
from both numerical simulations and the underlying breakup models. The
results show that the implementation of the breakup models was correct.

141
6. Implementation of breakup models

Furthermore, the comparison of the simulated and experimental average ver-


tical liquid velocity revealed that the simulations considering coalescence
and breakup of bubbles show quite different results when compared with the
model without considering coalescence and breakup. For instance, the model
that uses the critical Weber number We = 2.48 gives better predictions on the
vertical liquid velocity in the lower part of the bubble column. Meanwhile,
it is also found that the model with uniform bubble size and swarm effect
perform better at low superficial gas velocity (u g = 0.0049 m/s).
In addition, the bubble size distributions obtained from the numerical
simulation were compared with those estimated from the chord length dis-
tributions with a four-point optical fibre probe. The results suggest that the
incorporation of both coalescence and breakup models in the discrete bubble
model produces reasonable results with respect to the evolution of bubble
size in bubbly flows.

Nomenclature

a shape parameter of beta distribution, [-]


b shape parameter of beta distribution, [-]
C model coefficient, [-]
d bubble diameter, [m]
D column depth, [m]
(ρl − ρ g )gd2
Eo Eötvös number, Eo = , [-]
σ
f function, [-]
F force vector, [N]
g gravity acceleration, [m s−2 ]
h0 initial film thickness, [m]
h1 critical film thickness, [m]
H column height, [m]
I unit tensor, [-]
n unit normal vector, [-]
p pressure, [N m−2 ]

142
Nomenclature

r position vector, [m]


R bubble radius, [m]
ρl |v − u|d
Re bubble Reynolds number, Re = , [-]
µl
t time, [s]
u liquid velocity vector, [m s−1 ]
v bubble velocity vector, [m s−1 ]
V bubble volume, [m3 ]
V∗ relative bubble volume, [-]
W column width, [m]
We Weber number, [-]

Greek letters

α void fraction, [-]


β daughter size distribution, [-]
 turbulent kinetic energy dissipation rate, [m2 s−3 ]
Γ gamma distribution, [-]
µ viscosity, [kg m−1 s−1 ]
Φ volume averaged momentum transfer due to interphase forces,
[kg m−2 s−2 ]
Ω breakup frequency, [s−1 ]
ρ density, [kg m−3 ]
σ surface tension, [N m−1 ]
τ stress tensor, [N m−2 ]; contact time, [s]

Indices

∆ subgrid length scale, [m]


D drag
eff effective
g gas phase
G gravity
H horizontal direction
143
6. Implementation of breakup models

i bubble index
j bubble index
l liquid phase
L lift; molecular viscosity
m mixture
P pressure
T turbulence
VM virtual mass
W wall
∞ infinite medium

References

T. R. Auton. The lift force on a spherical body in a rotational flow. J. Fluid


Mech., 197:241–257, 1987.

R. Bannari, F. Kerdouss, B. Selma, A. Bannari, and P. Proulx. Three-


dimensional mathematical modeling of dispersed two-phase flow using
class method of population balance in bubble columns. Computers and
Chemical Engineering, 32(12):3224–3237, 2008.

G. K. Batchelor. The theory of homogeneous turbulence. Cambridge University


Press, 1953.

P. Chen, M. P. Duduković, and J. Sanyal. Three-dimensional simulation of


bubble column flows with bubble coalescence and breakup. AIChE Journal,
51(3):696–712, 2005.

C. A. Coulaloglou and L. L. Tavlarides. Description of interaction processes


in agitated liquid-liquid dispersions. Chemical Engineering Science, 32(11):
1289–1297, 1977.

D. Darmana, N. G. Deen, and J. A. M. Kuipers. Parallelization of an Euler-


Lagrange model using mixed domain decomposition and a mirror domain
technique: Application to dispersed gas-liquid two-phase flow. Journal of
Computational Physics, 220(1):216–248, 2006.

144
References

A. Delafosse, J. Morchain, P. Guiraud, and A. Liné. Trailing vortices generated


by a Rushton turbine: Assessment of URANS and large Eddy simulations.
Chemical Engineering Research and Design, 87(4):401–411, 2009.

E. Delnoij, F. A. Lammers, J. A. M. Kuipers, and W. P. M. van Swaaij. Dynamic


simulation of dispersed gas-liquid two-phase flow using a discrete bubble
model. Chemical Engineering Science, 52(9):1429–1458, 1997.

E. Delnoij, J. A. M. Kuipers, and W. P. M. van Swaaij. A three-dimensional


CFD model for gas-liquid bubble columns. Chemical Engineering Science, 54
(13-14):2217–2226, 1999.

L. Hagesaether, H. A. Jakobsen, and H. F. Svendsen. A model for turbulent


binary breakup of dispersed fluid particles. Chemical Engineering Science, 57
(16):3251–3267, 2002.

H. Hartmann, J. J. Derksen, C. Montavon, J. Pearson, I. S. Hamill, and H. E. A.


van den Akker. Assessment of large eddy and RANS stirred tank simu-
lations by means of LDA. Chemical Engineering Science, 59(12):2419–2432,
2004.

J. O. Hinze. Fundamentals of the hydrodynamic mechanism of splitting in


dispersion processes. AIChE Journal, 1(3):289–295, 1955.

A. K. Hofman and G. Chesters. Bubble coalescence in pure liquids. Applied


Scientific Research, 38(1):353–361, 1982.

B. P. B. Hoomans, J. A. M. Kuipers, W. J. Briels, and W. P. M. van Swaaij. Dis-


crete particle simulation of bubble and slug formation in a two-dimensional
gas-fluidised bed: A hard-sphere approach. Chemical Engineering Science,
51(1):99–118, 1996.

H. A. Jakobsen. Chemical Reactor Modeling: Multiphase Reactive Flows. Springer,


1st edition, July 2008.

C. Jiménez, F. Ducros, B. Cuenot, and B. Bédat. Subgrid scale variance and


dissipation of a scalar field in large eddy simulations. Physics of Fluids, 13
(6):1748–1754, 2001.

145
6. Implementation of breakup models

N. I. Kolev. Multiphase Flow Dynamics 2: Thermal and Mechanical Interactions.


Springer, 3rd edition, 2007.

M. Kostoglou and A. J. Karabelas. Toward a unified framework for the deriva-


tion of breakage functions based on the statistical theory of turbulence.
Chemical Engineering Science, 60(23):6584–6595, 2005.

M. Kostoglou, S. Dovas, and A. J. Karabelas. On the steady-state size distri-


bution of dispersions in breakage processes. Chemical Engineering Science,
52(8):1285–1299, 1997.

J. C. Lasheras, C. Eastwood, C. Martı́nez-Bazán, and J. L. Montaes. A review


of statistical models for the break-up an immiscible fluid immersed into a
fully developed turbulent flow. International Journal of Multiphase Flow, 28
(2):247–278, 2002.

C. H. Lee, L. E. Erickson, and L. A. Glasgow. Bubble breakup and coalescence


in turbulent gas-liquid dispersions. Chemical Engineering Communications,
59(1-6):65–84, 1987.

F. Lehr and D. Mewes. A transport equation for the interfacial area density
applied to bubble columns. Chemical Engineering Science, 56(3):1159–1166,
2001.

F. Lehr, M. Millies, and D. Mewes. Bubble-size distributions and flow fields


in bubble columns. AIChE Journal, 48(11):2426–2443, 2002.

Y. Liao and D. Lucas. A literature review of theoretical models for drop and
bubble breakup in turbulent dispersions. Chemical Engineering Science, 64
(15):3389–3406, 2009.

H. Luo and H. F. Svendsen. Theoretical model for drop and bubble breakup
in turbulent dispersions. AIChE Journal, 42(5):1225–1233, 1996.

C. Martı́nez-Bazán, J. L. Montañés, and J. C. Lasheras. On the breakup of an


air bubble injected into a fully developed turbulent flow. Part 1. Breakup
frequency. Journal of Fluid Mechanics, 401:157–182, 1999a.

146
References

C. Martı́nez-Bazán, J. L. Montañés, and J. C. Lasheras. On the breakup of an


air bubble injected into a fully developed turbulent flow. Part 2. Size PDF
of the resulting daughter bubbles. Journal of Fluid Mechanics, 401:183–207,
1999b.

R. Mihail and S. Straja. A theoretical model concerning bubble size distribu-


tions. The Chemical Engineering Journal, 33(2):71–77, 1986.

E. Olmos, C. Gentric, Ch. Vial, G. Wild, and N. Midoux. Numerical simulation


of multiphase flow in bubble column reactors. Influence of bubble coales-
cence and break-up. Chemical Engineering Science, 56(21-22):6359–6365, 2001.

M. J. Prince and H. W. Blanch. Bubble coalescence and break-up in air-sparged


bubble columns. AIChE Journal, 36(10):1485–1499, 1990.

D. Qian, J. B. McLaughlin, K. Sankaranarayanan, S. Sundaresan, and K. Kon-


tomaris. Simulation of bubble breakup dynamics in homogeneous turbu-
lance. Chemical Engineering Communications, 193(8):1038–1063, 2006.

F. Risso. Mechanisms of deformation and breakup of drops and bubbles.


Multiphase Science and Technology, 12(1):1–50, 2000.

F. Risso and J. Fabre. Oscillations and breakup of a bubble immersed in a


turbulent field. Journal of Fluid Mechanics, 372:323–355, 1998.

H. Rusche. Computational fluid dynamics of dispersed two-phase flows at high phase


fractions. PhD thesis, Imperial College, London, 2002.

A. N. Sathyagal, D. Ramkrishna, and G. Narsimhan. Droplet breakage in


stirred dispersions. breakage functions from experimental drop-size distri-
butions. Chemical Engineering Science, 51(9):1377–1391, 1996.

M. Sommerfeld, E. Bourloutski, and D. Bröder. Euler/Lagrange calculations


of bubbly flows with consideration of bubble coalescence. Canadian Journal
of Chemical Engineering, 81(3-4):508–518, 2003.

147
6. Implementation of breakup models

A. Tomiyama, T. Matsuoka, T. Fukuda, and T. Sakaguchi. A simple numerical


method for solving an incompressible two-fluid model in a general curvi-
linear coordinate system. Advances in Multiphase Flow, pages 241–252,
Amsterdam, 1995. Society of Petroleum Engineers Inc.

A. Tomiyama, I. Kataoka, I. Zun, and T. Sakaguchi. Drag coefficients of single


bubbles under normal and micro gravity conditions. JSME international
journal. Ser. B, Fluids and thermal engineering, 41(2):472–479, 1998.

A. Tomiyama, H. Tamai, I. Zun, and S. Hosokawa. Transverse migration of


single bubbles in simple shear flows. Chemical Engineering Science, 57(11):
1849–1858, 2002.

C. Tsouris and L. L. Tavlarides. Breakage and coalescence models for drops


in turbulent dispersions. AIChE Journal, 40(3):395–406, 1994.

T. Uga. Determination of bubble-size distribution in a BWR. Nuclear Engineer-


ing and Design, 22(2):252–261, 1972.

A. W. Vreman. An eddy-viscosity subgrid-scale model for turbulent shear


flow: Algebraic theory and applications. Physics of Fluids, 16(10):3670–3681,
2004.

J. F. Walter and H. W. Blanch. Bubble break-up in gas-liquid bioreactors:


Break-up in turbulent flows. The Chemical Engineering Journal, 32(1):B7–B17,
1986.

T. Wang, J. Wang, and Y. Jin. A novel theoretical breakup kernel function for
bubbles/droplets in a turbulent flow. Chemical Engineering Science, 58(20):
4629–4637, 2003.

T. Wang, J. Wang, and Y. Jin. A CFD-PBM coupled model for gas-liquid flows.
AIChE Journal, 52(1):125–140, 2006.

R. M. Wellek, A. K. Agrawal, and A. H. P. Skelland. Shape of liquid drops


moving in liquid media. AIChE Journal, 12(5):854–862, 1966.

148
Chapter
7
Numerical investigation of gas
holdup and phase mixing in
bubble column reactors
A discrete bubble model (DBM) has been used to study the overall gas holdup and the
phase mixing in bubble column reactors. The Eulerian-Lagrangian approach has the
advantage that it is possible to study the overall gas holdup and the gas phase mixing
in a direct way. Furthermore, by introducing tracer particles, also the liquid phase
mixing can be studied in a Lagrangian manner.
Comparisons suggest that the overall gas holdups obtained from the discrete bubble
model agree very well with the correlations of Kumar et al. (1976), Heijnen and
Van’t Riet (1984) and Ruzicka et al. (2001) within the applied range of the superficial
gas velocity. The gas phase dispersion coefficients from the simulation results agree
pretty well with Wachi and Nojima (1990)’s correlation, which is derived based on
the recirculation theory. In addition, the turbulent diffusion coefficient calculated
from the tracer particles velocities are very close to those from the literature within
the applied range of the superficial gas velocity.

149
7. Gas holdup and phase mixing in bubble column reactors

7.1 Introduction

Bubble column reactors are frequently utilized in the chemical, petrochemical,


biochemical and metallurgical industries. A bubble column reactor is basically
a liquid filled vessel with a gas distributor at the bottom. The gas is sparged
through the distributor in the form of bubbles and comes into contact with the
liquid. The resulting buoyancy driven flow creates strong liquid recirculation.
Bubble column reactors have the advantage that little maintenance and low
operating costs are required due to lack of moving parts and compactness.
Moreover, they have excellent heat and mass transfer characteristics.
There are some important parameters, such as gas holdup, gas-liquid
interfacial area, interfacial mass transfer coefficient, dispersion coefficient and
heat transfer coefficient and so on, which are essential to characterize, scale up
and design the bubble column reactors. For instance, the gas holdup gives the
volume fraction of the phases present in the reactor. On the other hand, the
gas holdup combined with knowledge of the mean bubble diameter allows
determination of the specific interfacial area and the related volumetric mass
transfer rates between the gas and liquid phase. Moreover, the gas holdup is
usually used to identify the flow regime in bubble column reactors. During
the past decades, extensive studies on the gas holdup have been carried out
in several contexts, such as flow regime identification and factors that may
influence the gas holdup in bubble column reactors, such as gas flow rate,
liquid flow rate, geometry of the bubble column, operating conditions (such
as pressure and temperature), physical properties of both phases, sparger type
and so on. The related work has been reviewed by many authors (Shah et al.,
1982; Heijnen and Van’t Riet, 1984; Deckwer and Schumpe, 1993; Kantarci
et al., 2005; Ribeiro Jr. and Lage, 2005; Shaikh and Al-Dahhan, 2007; Joshi
et al., 2008).
Dispersion coefficients are of significance in chemical reactors as well.
There are two main model reactors and related assumptions regarding pre-
vailing state of mixing, ideally mixed flow and plug flow in either single
phase or multiphase chemical reactors (Westerterp et al., 1987; Levenspiel,
1999; Scott Fogler, 2005). In the ideally mixed flow system, the composition of

150
7.2. Discrete bubble model

the reaction mixture is uniform and is typically used to describe either batch
reactors or continuously mixed tank reactors. On the other hand, mixing or
diffusion of the reacting mixture does not occur in the flow direction in plug
flow reactors. Real chemical reactors, however, may neither exhibit perfect
mixed nor plug-flow behavior. For this reason, axial dispersion models, have
been developed to describe the deviations from these simplified model reac-
tors. In addition, the concept of residence time distribution (RTD) has proven
a powerful tool to diagnose and characterize non-ideal reactors since Danck-
werts (1953). In bubble column reactors, investigations on the mixing of both
the gas and liquid phase have been reviewed by several authors, such as Joshi
(1980, 1982); Shah et al. (1982); Heijnen and Van’t Riet (1984) and Deckwer
and Schumpe (1993).
Computational Fluid Dynamics (CFD) has gained considerable interest in
recent years in the field of chemical engineering (Jakobsen, 2008). In addition
to experimental routes, it provides an alternative way to study and character-
ize chemical reactors in detail. On the other hand, chemical reactor modeling
is also crucial for scaleup and design of chemical reactors.
In this work we will focus on the prediction of gas holdup and mixing of
both the gas and liquid phase using the Eulerian-Lagrangian approach. The
simulation results will be compared with corresponding correlations obtained
from literature.

7.2 Discrete bubble model

The discrete bubble model (DBM) used in this study was originally developed
by Delnoij et al. (1997) and Delnoij et al. (1999). In this model, the liquid
phase hydrodynamics is represented by the volume-averaged continuity and
Navier-Stokes equations while the motion of the each individual bubble is
tracked in a Lagrangian way.

7.2.1 Bubble dynamics


The motion of each individual bubble is computed with Newton’s second
law. The liquid phase contributions are taken into account by the interphase

151
7. Gas holdup and phase mixing in bubble column reactors

momentum transfer experienced by each individual bubble. For an individual


bubble, the equations can be written as:

dv
ρg V = ΣF (7.1)
dt

dr
=v (7.2)
dt
The net force acting on each individual bubble is calculated by considering
all contributing forces. It is assumed that the net force is composed of separate,
uncoupled contributions due to gravity, pressure, drag, lift, virtual mass and
wall force respectively:

ΣF = FG + FP + FD + FL + FVM + FW (7.3)

The gravity force acting on a bubble in a liquid is given by:

FG = ρ g Vg (7.4)

The far field pressure force incorporating contributions of the Archimedes


buoyancy force, inertial forces and viscous strain is given by:

FP = −V∇P (7.5)

The drag force exerted on a bubble rising in a liquid is expressed as:

1
FD = − CD ρl πd2 |v − u|(v − u) (7.6)
8
where the drag coefficient CD is determined according to Rusche (2002) which
takes swarm effects into account:
h i
CD = CD∞ exp(3.64α g ) + α0.864
g (7.7)

where the drag coefficient of a single bubble CD∞ is taken from Tomiyama
et al. (1998):
   
16 48 8 Eo
CD∞ = max min (1 + 0.15Re0.687 ), , (7.8)
Re Re 3 Eo + 4

152
7.2. Discrete bubble model

ρl |v − u|d
where Re is the bubble Reynolds number, Re = .
µl
A bubble rising in a non-uniform liquid flow field experiences a lift force
due to vorticity or shear in the flow field. The shear induced lift force acting
on a bubble is usually written as (Auton, 1987):

FL = −CL ρl V(v − u) × ∇ × u (7.9)

where the lift coefficient is calculated according to Tomiyama et al. (2002):




 min[0.288tanh(0.121Re), f (EoH )] EoH < 4,




CL =  f (EoH ) 4 ≤ EoH ≤ 10, (7.10)



−0.29 EoH > 10.

where
f (EoH ) = 0.00105Eo3H − 0.0159Eo2H − 0.0204EoH + 0.474 (7.11)
The modified Eötvös number, EoH is defined by using the maximum horizon-
tal dimension of a bubble as a characteristic length as follows:
(ρl − ρ g )gd2H
EoH = (7.12)
σ
The maximum horizontal diameter of the bubble is obtained from the
bubble aspect ratio E according to Wellek et al. (1966):
dV 1
E= = (7.13)
dH 1 + 0.163Eo0.757
where dV is the maximum vertical diameter of the bubble and Eo is the Eötvös
(ρl − ρ g )gd2
number, Eo = .
σ
The relation between the above two diameters and the diameter of the
bubble d in the discrete bubble modeling is as follows:

d = (dV d2H )1/3 (7.14)

Accelerating bubbles experience a resistance, which is described as the


virtual mass force (Auton, 1987):
 Dv Du 
FVM = −CVM ρl V − (7.15)
Dt Dt

153
7. Gas holdup and phase mixing in bubble column reactors

where the D/Dt operators denote the substantiative derivatives pertaining to


the respective phases. In the present work, bubbles are assumed to have a
spherical shape and a virtual mass coefficient of 0.5 is used.
Bubbles in the vicinity of a solid wall experience a force referred to as the
wall force (Tomiyama et al., 1995):
" #
1 1 1
FW = − CW d 2 − ρl |(v − u)· nz |2 nW (7.16)
2 y (L − y)2

where nz and nW , respectively, are the normal unit vectors in the vertical and
wall normal direction, L is the dimension of the system in the wall normal
direction, and y is the distance to the wall in that direction. Finally, the wall
force coefficient CW is given by:



exp(−0.933Eo + 0.179) 1 ≤ Eo ≤ 5,
CW =
 (7.17)
0.007Eo + 0.04 5 < Eo ≤ 33.

7.2.2 Liquid phase dynamics


The liquid phase hydrodynamics are described by a set of volume-averaged
equations, which consists of the continuity and the Navier-Stokes equations.
The presence of the bubbles is reflected by the local volume fraction of the
liquid phase αl and the interphase momentum transfer rate Φ:

∂  
αl ρl + ∇ · αl ρl u = 0 (7.18)
∂t
∂  
αl ρl u + ∇ · αl ρl uu = −αl ∇p − ∇ · (αl τl ) + αl ρl g + Φ (7.19)
∂t
The liquid phase is assumed to be Newtonian, thus the stress tensor τl can be
expressed as:
 
2
τl = −µeff,l ((∇u) + (∇u)T − I(∇ · u) (7.20)
3
where µeff,l is the effective viscosity. In the present model, the effective viscos-
ity is composed of two contributions, the molecular viscosity and the turbulent
viscosity:
µeff,l = µL,l + µT,l (7.21)

154
7.3. Correlations of the gas holdup

where µT,l is the turbulent viscosity(or eddy viscosity), which is determined


from turbulence modeling of the liquid phase.
In the present work, large eddy simulation (LES) is used to simulate the
turbulence induced by the rising bubbles. Therefore, the above continuity and
Navier-Stokes equations are resolved for the field representing the “large”
eddies, while the effect of the subgrid part of the velocity representing the
”small scales” is included through an eddy-viscosity subgrid-scale model.
The eddy-viscosity model proposed by Vreman (2004) is used to calculate the
eddy viscosity:
s

µT,l = 2.5ρl CS
2
(7.22)
αij αij

where αij = ∂u j /∂xi , βij = ∆2m αmi αmj and Bβ = β11 β22 −β212 +β11 β33 −β213 +β22 β33 −
β223 . ∆i is the filter width in the i direction.

7.2.3 Collision model


In this work, a hard sphere collision model developed by Hoomans et al.
(1996) is adopted to describe the bubble-bubble interaction. It mainly consists
of two parts. The first part is processing the sequence of collisions and the
second part is dealing with the collision dynamics. A detailed description
about the model is referred to Darmana (2006) and Hoomans (2000).

7.3 Correlations of the gas holdup

In bubble column reactors, the gas holdup is an important parameter which


can be used to characterize the hydrodynamics. It is defined as the overall
volume fraction of the gas phase present in the form of bubbles. The gas
holdup mainly depends on the superficial gas velocity. It can be used to
estimate the interfacial area and is thus required to estimate the volumetric
mass transfer rates between the gas and liquid phase.
Fair et al. (1962) carried out measurements of the gas holdup and heat
transfer in two commercial scale bubble columns with an air-water system.
The diameters of the columns were 0.54 m and 1.26 m respectively. By com-

155
7. Gas holdup and phase mixing in bubble column reactors

paring with results obtained in those columns with smaller diameter, they
found that the gas holdup varies directly with the superficial gas velocity and
is less in the column with a large diameter than in the column with a small
diameter. However, when the diameter is larger than 0.54 m, the gas holdup
is no longer influenced by the column diameter. The experimental data repro-
duced from Fair et al. (1962), a linear fitted curve and the gas holdup curve
from Shulman (1950) for a bubble column with a diameter of 0.12 m are shown
in Figure 7.1.

Φ=0.54 m
0.3 Φ=1.26 m
Fitted curve
Gas holdup [-]

Φ=0.12 m

0.2

0.1

0
0 0.02 0.04 0.06
ug [m/s]

Figure 7.1: Gas holdup curve (reproduced from) Fair et al. (1962).

Kato and Nishiwaki (1972) proposed the following correlation for the gas
holdup in an air-water system (Shah et al., 1982):

2.51u g
= γ
(7.23)
0.78 + βu0.8
g (1 − e )

where
β = 4.5 − 3.5 − 2.548φ1.8 (7.24)

and
γ = 717u1.8
g /β (7.25)

156
7.3. Correlations of the gas holdup

Kumar et al. (1976) correlated 382 gas holdup data points ( < 0.35andu g <
0.15 m/s) with the following equation:

 = 0.728u∗ − 0.485u∗2 + 0.0975u∗3 (7.26)

where u∗ is defined as a dimensionless gas velocity given by:


 2 0.25
 ρ 
u = u g  l 

(7.27)
σ∆ρg

According to the circulation theory and an assumption for the single bub-
ble rise velocity, Heijnen and Van’t Riet (1984) proposed the following expres-
sion for the homogeneous flow regime:

 = 4u g (7.28)

which is consistent with the experimental observations by Krishna et al. (1991).


Based on the concept of a characteristic turbulent kinematic viscosity
in bubble columns, Kawase and Moo-Young (1987) developed a theoretical
model for the gas holdup with Newtonian and non-Newtonian fluids. For
Newtonian fluids, the equation is given as:
 2 1/3
 u g 
 = 1.07   (7.29)
φg

Based on hydrodynamic coupling between both phases, Ruzicka et al.


(2001) developed a simple physical model for the homogeneous to heteroge-
neous regime transition in a bubble column. For the homogeneous regime,
the formula for the gas holdup is given by:

u g + uT − B
= (7.30)
2(1 + a)uT
where
B = u2g + u2T − 2(2a + 1)u g uT (7.31)
For a bubble column with dimensions (H = 1 m, φ = 0.15 m), the two param-
eters, uT and a, are 0.22 m/s and 0.35 respectively according to experimental
data.

157
7. Gas holdup and phase mixing in bubble column reactors

7.4 Phase mixing

There are several factors that may induce longitudinal mixing in a bubble
column, such as turbulent eddies, a nonuniform velocity distribution over the
cross section of the reactor and molecular diffusion. In most of the practical
cases, the former two factors dominate the molecular diffusion (Westerterp
et al., 1987).

7.4.1 Axial dispersion model


Axial dispersion models are often used to characterize the extent of phase mix-
ing of chemical reactors. It assumes a diffusion-like process superimposed on
plug flow. The unsteady state one-dimensional convection diffusion equation
is given by Levenspiel (1999):

∂C ∂2 C ∂(UC)
=D 2 − (7.32)
∂t ∂z ∂z
where C represents the species concentration, D the axial dispersion coefficient
and U a characteristic velocity that depends on the phase. For instance for the
gas phase, the characteristic velocity is defined as U = u g /.
The axial dispersion coefficient D is directly related to the process of mixing
in reactors. For instance, a large D means rapid mixing while a small D implies
slow mixing, whereas D = 0 corresponds to plug flow conditions.
When the boundary conditions are known, i.e. open-open boundary con-
ditions, the above equation can be solved analytically. From a pulse tracer
experiment, the dispersion coefficient can then be obtained from the tracer
curve at the outlet of the reactor, that is, from the mean and variance of the
residence time distribution of the tracer.

7.4.2 Residence time distribution


Since Danckwerts (1953) introduced the concept of residence time distribution,
it has been used as an important tool in the analysis of chemical reactor
performance.

158
7.4. Phase mixing

The residence time distribution E(t) normally has the property that the
area under the curve is unity:
Z ∞
E(t) dt = 1 (7.33)
0

In addition, the mean residence time tm and the variance σ2 of the residence
time distribution can be calculated as follows:
Z ∞
tm = tE(t) dt (7.34)
0
Z ∞
σ2 = (t − tm )2 E(t) dt (7.35)
0
Consequently the axial dispersion coefficient can be obtained with the
mean residence time tm and the variance σ2 . For instance, the Peclet number
Pe can be calculated in a closed-closed system (Levenspiel, 1999):

σ2 2 2
σ2θ = = − (1 − e−Pe ) (7.36)
t2m Pe Pe2

The dispersion coefficient D can hence be determined from the Peclet


number:
UH
D= (7.37)
Pe
where H is the height of the system.

7.4.3 Lagrangian description of the mixing


In a Lagrangian view, the diffusion coefficient can be evaluated from the
change in mean-square displacement of fluid elements z2 with time t (Brodkey,
1972):
1 dz2
D= (7.38)
2 dt
where the mean-square displacement z2 is calculated as:
 2
1 X  1 X 
N N
z2 = (zi (t) − zi (0)) −
 (zi (t) − zi (0)) (7.39)
N N
i=1 i=1

159
7. Gas holdup and phase mixing in bubble column reactors

where N represents the number of fluid elements.


In turbulent flows, a turbulent diffusion coefficient can be expressed as
(Brodkey, 1972):
DT = v02 TL (7.40)

where v0 is the root-mean-square value of the fluctuating part of the instanta-


neous fluid velocity and the Lagrangian time scale TL is defined as:
Z ∞
TL = RL (τ) dτ (7.41)
0

where
v(t)v(t + τ)
RL (τ) = (7.42)
v02

7.4.4 Correlations for the gas phase dispersion coefficient


For the axial dispersion coefficient of the gas phase in a bubble column, Diboun
and Schügerl (1967) suggested (Westerterp et al., 1987):

ur φ
Bo = = 0.2 (7.43)
Dg

where Bo is Bodenstein number and ur is the relative velocity between the gas
and the liquid phase.
Men’shchikov and Aerov (1967) studied the gas phase axial mixing in a
bubble column (φ = 0.3 m) with the superficial gas velocity varying in the
range of 0.008-0.1 m/s and proposed the following correlation (Joshi, 1982):

D g = 1.47u0.72
g (7.44)

Towell and Ackermann (1972) correlated their own experimental data and
those from the literature with a correlation as follows (Joshi, 1982):

D g = 19.7φ2 u g (7.45)

The superficial gas velocity applied for the correlation ranges from 0.009 m/s
to 0.13 m/s, whereas the diameter of the bubble column ranges from 0.09 m to
1 m.

160
7.4. Phase mixing

Joshi (1982) reviewed six experimental investigations for the gas phase
dispersion in bubble columns. Based on this, he proposed the following
correlation which covers all the experimental data presented in the six exper-
imental studies:
u2g
D g = 110φ2 (7.46)

Heijnen and Van’t Riet (1984) suggested that the dispersion coefficient is
constant in the homogeneous flow regime according to experimental data in
the literature. They gave the following dispersion coefficient correlation in
the heterogeneous flow regime for the air-water system:

D g = 78(u g φ)1.5 (7.47)

However, they also mentioned that the transition from the homogeneous
to heterogeneous flow regime is difficult to distinguish due to geometric
parameters, such as sparger location and uniformity of gas distribution.
On the basis of the recirculation theory for a bubble column in the turbulent
flow regime, Wachi and Nojima (1990) derived the following equation for the
axial dispersion coefficient of the gas phase:

D g = 20φ3/2 u g (7.48)

They also measured the gas phase dispersion coefficients in two bubble
columns (φ = 0.2 m and 0.5 m) using pulse experiments at superficial gas
velocities in the range of 0.029-0.456 m/s. Comparison between the theoreti-
cal correlation and the measurements revealed that the derived correlation is
able to reflect the dependencies of the axial dispersion of the gas phase on the
column diameter and the superficial gas velocity.

7.4.5 Correlations for the liquid phase dispersion coefficient


Ohki and Inoue (1970) measured the axial liquid phase dispersion coefficient
in batch type bubble columns of three different diameters (φ = 0.04 m, 0.08 m
and 0.16 m) and established a correlation as follows:

Dl = 0.30φ2 u1.2
g + 170δ (7.49)

161
7. Gas holdup and phase mixing in bubble column reactors

where δ is hole diameter of the perforated plate. Note that the units in the
equation are non-SI units. That is, the unit of dispersion coefficient is cm2 /s,
superficial gas velocity has unit of cm/s and the hole diameter is in cm.
Furthermore, the range of superficial gas velocity in his study ranges 0.02 m/s
and 0.25 m/s.
Towell and Ackermann (1972) developed a correlation for dispersion coef-
ficient as a function of the column diameter and the gas velocity (Majumder,
2008):
Dl = 1.23φ1.5 u0.5
g (7.50)
The diameters of the bubble columns used in this study are 0.04 m and 1.07 m,
while the superficial gas velocity ranged from 0.016 to 0.13 m/s and 0.009 to
0.034 m/s respectively (Joshi, 1982).
By measuring the dispersion coefficient in two tall bubble columns with
(φ = 0.15 m and 0.2 m), Deckwer et al. (1974) correlated his own experimental
data with those in literature and proposed an expression which is similar to
the correlation proposed by Towll and Ackermann (1972):

Dl = 0.678φ1.4 u0.3
g (7.51)

The superficial gas velocity used in the two bubble columns ranged from
0.004 to 0.13 m/s and 0.01 to 0.14 m/s respectively.
Hikita and Kikukawa (1974) studied the liquid dispersion coefficient in two
bubble columns and presented a correlation for liquid dispersion coefficient:

Dl = (0.15 + 0.69u0.77
g )φ
1.25
(7.52)

Baird and Rice (1975) adopted dimensional analysis to derive the disper-
sion coefficient based on isotropic turbulence model:

Dl = 0.35φ4/3 (u g g)1/3 (7.53)

Even though the turbulence in a bubble column is not necessarily isotropic,


the correlation using an isotropic turbulence model as the basis gives a di-
mensionally consistent and simple equation.
Zehner (1986) proposed a correlation based on a vortex cell model:

Dl = 0.368φ4/3 (u g g)1/3 (7.54)

162
7.5. Numerical aspects of phase mixing study

where the constant 0.368 is an adjustable parameter.


Heijnen and Van’t Riet (1984) derived the liquid dispersion coefficient
from either the liquid circulation velocity or the Peclet number according to
Joshi (1980). They found out that the two correlations are in good agreement
with each other except for the constant C. The correlations have the following
general form:
Dl = Cφ4/3 (u g g)1/3 (7.55)
where C equals either 0.33 or 0.36.
By relating the axial dispersion coefficient to the longitudinal displacement
of a fluid element during a certain period, Kawase and Moo-Young (1986)
derived the following correlation valid for Newtonian liquids:

Dl = 0.343φ4/3 (u g g)1/3 (7.56)

Note that most of the above correlations that are either obtained from
experiments or derived theoretically in the literature show proportional de-
pendencies of the liquid phase axial dispersion coefficient on the diameter of
the bubble column and the superficial gas velocity. The only difference is that
the powers of these two parameters and the constant vary slightly from each
other.

7.5 Numerical aspects of phase mixing study

7.5.1 Residence time distribution of the gas phase


In Eulerian-Lagrangian modeling of bubbly flows, each bubble is tracked
individually according to Newton’s second law. Hence, the residence time
of each bubble in the system can be determined directly from the simulation
and from this information, the axial dispersion coefficient can be determined
according to the axial dispersion model.

7.5.2 Tracer particles of the liquid phase


Massless tracer particles have been used to visualize unsteady flows in Com-
putational Fluid Dynamics (CFD) (Hin and Post, 1993; Lane, 1997; Bauer et al.,

163
7. Gas holdup and phase mixing in bubble column reactors

2002). In the present work, tracer particles are used to study the mixing of the
liquid phase.
For the liquid phase, the Navier-Stokes equations are solved on an Eulerian
grid and the liquid velocities are obtained on the faces of each computational
cell. The velocity of a tracer particle v at a certain moment t can then be
interpolated according to its position r(t) in the grid. Accordingly, after a time
∆t, the new position of the tracer particle can be calculated as:
Z t+∆t
r(t + ∆t) = r(t) + v(r(t), t) dt (7.57)
t

Equation 7.57 can be evaluated using a numerical integration scheme.


Lane (1997) has reported that the first-order integration scheme could lead to
erroneous results and suggested a higher-order integration to be used. Hence,
the Runge-Kutta-Fehlberg method is used in the present simulations. Details
of this method are given in appendix 7.A.
Yeung and Pope (1988) have shown that the accuracy of the tracer particle
trajectories depends on the accuracy of the interpolation scheme used to cal-
culate the tracer particle velocities rather than on the time-stepping method.
In the present work, two interpolation methods, i.e. trilinear interpolation
and tricubic interpolation (Lekien and Marsden, 2005), are compared with
each other. More information on these interpolation methods is given in
appendix 7.B.

7.6 Simulation details

The bubble column studied here is shown in Figure 7.2. The cross-sectional
area of the column is 0.15 m × 0.15 m (W × D). The column is initially filled
with water to a height of 0.6 m (H). Air is used as the dispersed phase and
introduced into the column through a perforated plate at the bottom of the
column. The material properties of both phases are taken according to their
room temperature values. The bubble column is operating under atmospheric
pressure.
A perforated plate with 576 holes (24 × 24) is used as gas distributor. All
the holes have a diameter of 1 mm and are positioned in the central region of

164
7.6. Simulation details

Figure 7.2: Sketch of the square bubble column.

the plate with a square pitch of 6.25 mm.


A grid with (20×20×80) cells is adopted in the simulations. Five superficial
gas velocities (u g = 0.005, 0.01, 0.015, 0.02 and 0.025 m/s) are adopted. The
diameter of the bubbles in the simulations is kept as a constant, 0.004 m. Due
to the fact that the perforated plate adopted has many holes and bubbles
injected from each hole have very low velocities, the coalescence and breakup
of bubbles can be neglected. The time step for the liquid phase equations is set
to 1.0 × 10−3 s and for the bubbles it is set to 5.0 × 10−5 s. The total simulation
time is 150 s.
There are two sets of tracer particles being used in the simulations. The

165
7. Gas holdup and phase mixing in bubble column reactors

two sets of tracer particles are released uniformly over the cross-section of the
bubble column. The set of tracer particles released from the bottom of the
bubble column is denoted as “Tracer0” and the other set of tracer particles
released from the top of the bubble column is denoted as “Tracer1”. Moreover,
both sets of tracer particles are released simultaneously at the simulation time
t = 30 s at which the flow in the bubble column is fully developed.

7.7 Results and discussion

7.7.1 Gas holdup


Determination of the gas holdup

The instantaneous gas holdup is recorded at every 0.04 s during the sim-
ulation. Examples of gas holdup histories at a superficial gas velocity
u g = 0.005 m/s and u g = 0.025 m/s are plotted in Figure 7.3.

0.03
0.15

0.02 0.1
ε [-]

ε [-]

0.01 0.05

0 0
0 50 100 150 0 50 100 150
t [s] t [s]

(a) u g = 0.005 m/s (b) u g = 0.025 m/s

Figure 7.3: Instantaneous gas holdup [-].

The plots show that the gas holdup increases monotonically at the begin-
ning of the simulation. After a short time, the gas holdup reaches a peak
and then starts to fluctuate during the simulation. Bubbles are injected into
the bubble column continuously and rise in the vertical direction, which is
reflected in the increasing initial part of the curves. Once the bubbles reach the

166
7.7. Results and discussion

top of the bubble column, most of them escape from the column directly. This
leads to the small decrease of the gas holdup after the first peak. As soon as a
liquid circulation pattern in the bubble column has developed, some bubbles
may be trapped in the circulating liquid and remain in the column for a longer
time. Hence, fluctuations in the gas holdup curves result.
In order to obtain the overall gas holdup in the bubble column for each
superficial gas velocity, the instantaneous gas holdups are averaged after 30
seconds, in a similar fashion as described in Chapter 5. The resulting overall
gas holdup for each superficial gas velocity is listed in Table 7.1. From this
table it can be seen that the standard deviation of the averaged gas holdup σ
is quite small compared to the overall gas holdup. Furthermore, within the
adopted range of the superficial gas velocity, the overall gas holdup is related
to the superficial gas velocity with a multiplication factor of 4 approximately.

Table 7.1: Overall gas holdup.

u g [m/s] 0.005 0.01 0.015 0.02 0.025


 [-] 0.0234 0.0471 0.0671 0.0916 0.1119
σ [-] 0.0004 0.0017 0.0040 0.0056 0.0060

Comparison with literature correlations

In Figure 7.4, the overall gas holdups obtained from the simulations are com-
pared with the literature correlations introduced earlier.
It is seen that the simulation results agree with the correlations of Kumar
et al. (1976), Heijnen and Van’t Riet (1984) and Ruzicka et al. (2001) quite well.
These results indicate that the overall gas holdup increases nearly linearly
with the superficial gas velocity with a factor of 4 in the adopted range of
the superficial gas velocity. It can also be seen that the gas holdups obtained
from the correlation of Shulman (1950) are higher than those obtained from
others, which may be due to the diameter of the bubble column as reported
by Fair et al. (1962). In addition, the correlations of Fair et al. (1962) and Kato
and Nishiwaki (1972) are close to each other when the superficial gas velocity

167
7. Gas holdup and phase mixing in bubble column reactors

0.25
Shulman (1950)
Fair (1962)
0.2 Kato (1972)
Kumar (1976)
0.15 Kawase (1987)
ε [-] Heijnen (1984)
Ruzicka (2001)
0.1 DBM

0.05

0
0 0.005 0.01 0.015 0.02 0.025 0.03
ug [m/s]

Figure 7.4: Comparison of the computed gas holdup with literature correla-
tions.

u g < 0.025 m/s. As the superficial gas velocity increases, the gas holdup
obtained according to Kato and Nishiwaki (1972) increases faster than that
obtained from Fair et al. (1962). Moreover, Kawase and Moo-Young (1987)’s
correlation possesses a more flat trend.

7.7.2 Gas phase dispersion


Residence time distribution

The residence times of all individual bubbles in the bubble column are
recorded during the simulation and from this data set, the residence time
distribution is determined. In Figure 7.5, the residence time distribution for
each superficial gas velocity is shown. The plot shows a single peak in the
residence time distribution, which means that most of the bubbles in the bub-
ble column exit the column at the corresponding time. Moreover, the long tail
of the curve indicates that some bubbles remain in the column for longer res-
idence times due to reasons mentioned before. As the superficial gas velocity
increases, it is found that the peak of the residence time distribution shifts to
smaller residence time. That can be explained by the fact that at higher gas

168
7.7. Results and discussion

loading a stronger liquid circulation is created leading to a faster transport of


the bubbles. In addition, the distribution becomes wider at higher superficial
gas velocity.

0.06
ug=0.005 m/s
ug=0.01 m/s
ug=0.015 m/s
ug=0.02 m/s
0.04 ug=0.025 m/s
E(t) [1/s]

0.02

0
0 2 4 6 8 10
t [s]

Figure 7.5: Residence time distributions of the gas phase.

Relevant moments of the residence time distributions, such as the mean


residence time tm and the variance σ2 are determined from the curves. By
using the solution for a closed-closed system, the Peclet number and the
gas phase dispersion coefficients can be determined. The computed results
are shown in Table 7.2. From this table, it can be seen that the variance σ2
and the dimensionless variance σ2θ of the residence time distribution increase
with increasing superficial gas velocity, which is reflected by the wide spread
of the residence time distribution. Furthermore, the dispersion coefficient
increases with increasing superficial gas velocity. This implies that the extent
of gas phase backmixing becomes more pronounced at higher superficial gas
velocity.

Comparison with literature correlations

The simulated dispersion coefficients of the gas phase will subsequently be


compared with a number of correlations from the literature. Note that for

169
7. Gas holdup and phase mixing in bubble column reactors

Table 7.2: Mean and variance of the simulated residence time distributions.
u g [m/s] 0.005 0.01 0.015 0.02 0.025
tm [s] 2.88 2.94 2.90 2.99 3.01
σ2 [s2 ] 0.81 1.35 2.06 2.77 3.38
σ2θ [-] 0.10 0.16 0.25 0.31 0.37
Pe [-] 19.50 11.74 7.00 5.25 4.05
D g [m2 /s] 0.007 0.011 0.019 0.025 0.033

the earlier introduced correlations, the correlation of Joshi (1982) involves the
relation between the superficial gas velocity and the overall gas holdup in
the bubble column. Due to the fact that the overall gas holdups from the
simulations agree with the correlations in the literature quite well, we used
the overall gas holdup data obtained from the simulations to calculate the gas
phase dispersion coefficient from the correlation proposed by Joshi (1982). A
comparison is shown in Figure 7.6.

0.12
Diboun (1967)
0.1 Men'shchikov (1967)
Towell (1972)
Joshi (1982)
0.08
Dg [m2/s]

Heijnen (1984)
Wachi (1990)
0.06 DBM

0.04

0.02

0
0 0.005 0.01 0.015 0.02 0.025 0.03
ug [m/s]

Figure 7.6: Comparison of the simulated gas dispersion coefficient with liter-
ature correlations.

The comparison suggests that the simulation results agree very well with
the correlation of Wachi and Nojima (1990) which is derived based on the recir-

170
7.7. Results and discussion

culation theory. The gas phase dispersion coefficients from the simulations,
however, are far below the correlation of Men’shchikov and Aerov (1967).
Men’shchikov and Aerov (1967) proposed the correlation according to the ex-
perimental data from a bubble column with the diameter of 0.3 m and did not
consider the effect of the column diameter in the correlation. The dependence
of the gas phase dispersion on the column diameter, however, is used in most
of the correlations. Moreover, the effect of the column diameter on the gas
phase dispersion coefficient has been proven by Mangartz and Pilhofer (1981).
In addition, the simulation results are close to the correlations of Diboun and
Schügerl (1967), Towell and Ackermann (1972), Joshi (1982) and Heijnen and
Van’t Riet (1984) at low superficial gas velocities (i.e. u g ≤ 0.01 m/s), while
the difference becomes larger when the superficial gas velocity increases.

7.7.3 Liquid phase dispersion


Comparison of the interpolation methods

The two methods for interpolating the Eulerian velocity to the Lagrangian
velocity, tricubic interpolation and trilinear interpolation are compared with
each other. For instance, the mean-square displacement of the tracer particles
and the autocorrelation of vertical component of the tracer particle velocities
at superficial gas velocity u g = 0.005 m/s are shown in Figure 7.7.

(a) Meas-square displacement (b) Autocorrelation

Figure 7.7: Comparison of interpolation methods (u g = 0.005 m/s).

It can be seen that there are some minor differences between the two

171
7. Gas holdup and phase mixing in bubble column reactors

methods with respect to both the mean-square displacement and the autocor-
relation, which suggests that the interpolation scheme only slightly influences
the resulting Lagrangian time series. In the following discussions, only results
obtained with the tricubic interpolation scheme are considered.

Turbulent diffusion coefficient

The turbulent diffusion coefficient of the liquid phase in the bubble column
is determined according to Equation 7.40. The Lagrangian time scale TL is
determined as the time for the autocorrelation to decrease to 1/e of its initial
value (Squires and Eaton, 1991). The results obtained from both sets of tracer
particles are listed in Table 7.3. The turbulent diffusion coefficient determined
from the tracer particles “Tracer0” is denoted as Dl0 and that determined from
the tracer particles “Tracer1” is represented as Dl1 . Moreover, for each set of
tracer particles, the turbulent diffusion coefficient is determined four times
and the average of these four samples is presented in Table 7.3 along with the
standard deviation which are denoted as σ0 and σ1 .

Table 7.3: Turbulent diffusion coefficient of the liquid phase.

u g [m/s] 0.005 0.01 0.015 0.02 0.025


Dl0 [m /s]
2
0.006 0.010 0.019 0.016 0.021
σ0 [m2 /s] 0.002 0.004 0.002 0.002 0.003
Dl1 [m2 /s] 0.008 0.011 0.020 0.018 0.023
σ1 [m2 /s] 0.003 0.004 0.003 0.003 0.001

The table shows that the turbulent diffusion coefficients obtained from the
two set of tracer particles are quite close to each other. However, the standard
deviation of the turbulent diffusion coefficient is relatively large, particularly
at low superficial gas velocity (i.e. u g = 0.005 m/s). Moreover, it can be found
that the turbulent diffusion coefficient of the liquid phase nearly increases
with the superficial gas velocity.

172
7.7. Results and discussion

Comparison with literature correlations

The turbulent diffusion coefficient of the liquid phase obtained from the sim-
ulations are compared with the correlations in the literature and shown in
Figure 7.8.

0.04
Ohki (1970) Towell (1972)
Deckwer (1974) Hikita (1974)
Zehner (1986) Heijnen (1984)
0.03
DBM (Tracer0) DBM (Tracer1)
Dl [m2/s]

0.02

0.01

0
0 0.005 0.01 0.015 0.02 0.025 0.03
ug [m/s]

Figure 7.8: Comparison of the simulated liquid phase dispersion coefficient


with literature correlations.

According to the dispersion coefficient correlations of the liquid phase


introduced earlier, the correlations of Deckwer et al. (1974), Baird and Rice
(1975), Heijnen and Van’t Riet (1984), Zehner (1986) and Kawase and Moo-
Young (1986) are very close to each other. All of them exhibit similar de-
pendence on the column diameter and the superficial gas velocity, and the
constants of the correlations are also quite similar. Hence, only two of them
are used for the comparison in the plot (i.e. Heijnen and Van’t Riet (1984)’s
correlation with the constant equal to 0.33 and Zehner (1986)’s correlation).
Moreover, Towell and Ackermann (1972)’s correlation possesses a similar
trend with the above correlations because they use a very similar form of the
expression with only slightly different powers of the column diameter and the
superficial gas velocity. Hikita and Kikukawa (1974)’s correlation reveals a
weaker dependence on the superficial gas velocity than the others. Moreover,

173
7. Gas holdup and phase mixing in bubble column reactors

the dispersion coefficients calculated from Hikita and Kikukawa (1974)’s cor-
relation are higher than those obtained from the others at low superficial gas
velocity (i.e. u g < 0.02 m/s). Ohki and Inoue (1970)’s correlation increases
with the superficial gas velocity faster than the others and is larger than the
others when u g > 0.02 m/s. It is clearly found that the turbulent diffusion
coefficients obtained from the previous section are very close to the disper-
sion coefficients calculated from the correlations in the literature within the
applied range of the superficial gas velocity.

7.8 Conclusions

The present chapter utilized the discrete bubble model (DBM) to investigate
the overall gas holdup and the mixing of the both phases in bubble columns.
One of the advantages of the Eulerian-Lagrangian approach is that the dis-
persed phase is treated as individual elements, i.e. bubbles that are tracked
individually. The total number and the residence times of bubbles in the
bubble column reactors can be determined conveniently during simulations
and thus, the overall gas holdup and the residence time distribution of the
gas phase can be obtained accordingly. Hence, the dispersion coefficient de-
scribing the extent of the gas phase mixing can be calculated. Furthermore,
by introducing liquid phase tracer particles, it is possible to study the liquid
phase mixing as well.
In the first part of this chapter, the overall gas holdups from the dis-
crete bubble model are obtained within the range from u g = 0.005 m/s to
u g = 0.025 m/s. The results are then compared with correlations in the litera-
ture. The comparison reveals that the simulation results agree well with the
correlations of Kumar et al. (1976), Heijnen and Van’t Riet (1984) and Ruzicka
et al. (2001). The results suggest that the overall gas holdup increases nearly
linearly with the superficial gas velocity with a scaling constant of 4 in the
adopted range of the superficial gas velocity.
Secondly, the residence time distributions (RTD) of the gas phase are de-
termined in the residence times of bubbles from the discrete bubble model. A
sharp peak in the residence time distribution is obtained, which means that

174
Nomenclature

most of the bubbles in the bubble column exit the column at the corresponding
time. Moreover, the long tail of the curve suggests that some bubbles remain
in the column for longer residence times. As the superficial gas velocity in-
creases, one can find that the peak of the residence time distribution shifts to
smaller residence time, which is attributed to the stronger liquid circulation.
In addition, the distribution becomes wider when the superficial gas velocity
is higher, which is reflected by the increase of the variance of the residence
time distribution.
The comparison with the gas dispersion correlations in the literature shows
that the simulation results agree very well with Wachi and Nojima (1990)’s
correlation which is derived on basis of the recirculation theory.
Finally, the liquid phase mixing is studied through the introduction of
liquid phase tracer particles. Two methods for the interpolation scheme of
the Lagrangian velocity from the Eulerian velocity of the liquid phase are
compared with each other. Only small differences have been found between
the different methods. The turbulent diffusion coefficient is calculated from
the Lagrangian velocities and compared with correlations in literature. The
comparison shows that the obtained turbulent diffusion coefficients are very
close to those from the literature within the applied range of the superficial
gas velocity.

Nomenclature

C model coefficient, [-]; concentration of a species, [kg m−3 ]


Bo Bodenstein number, [-]
d bubble diameter, [m]
D column depth, [m]; dispersion coefficient, [m2 s−1 ]
f function, [-]
(ρl − ρ g )gd2
Eo Eötvös number, Eo = , [-]
σ
E(t) residence time distribution, [s−1 ]
F force vector, [N]
g gravity acceleration, [m s−2 ]

175
7. Gas holdup and phase mixing in bubble column reactors

H column height, [m]


I unit tensor, [-]
n unit normal vector, [-]
p pressure, [N m−2 ]
Pe Peclet number, [-]
r position vector, [m]
R bubble radius, [m]
ρl |v − u|d
Re bubble Reynolds number, Re = , [-]
µl
S characteristic filtered strain rate, [s−1 ]
t time, [s]
tm mean residence time, [s]
u liquid velocity vector, [m s−1 ]
U mean bubble velocity, [m s−1 ]
v bubble velocity vector, [m s−1 ]
V volume, [m3 ]
W column width,[ m]

Greek letters

α local volume fraction, [-]


∆ subgrid length scale, [m]
δ hole diameter of perforated plate, [m]
 gas holdup, [-]
µ viscosity, [kg m−1 s−1 ]
φ column diameter, [m]
Φ volume averaged momentum transfer due to interphase forces,
[kg m−2 s−2 ]
ρ density, [kg m−3 ]
σ surface tension, [N m−1 ]; standard deviation
σ2 variance, [s2 ]
σ2θ dimensionless variance, [-]
τ stress tensor, [N m−2 ]

176
Nomenclature

Indices

b bubble
cell computational cell
D drag
eff effective
g gas phase
G gravity
i i direction
j j direction
l liquid
L lift; molecular viscosity
P pressure
S subgrid
T turbulent
VM virtual mass
W wall

177
7. Gas holdup and phase mixing in bubble column reactors

7.A Runge-Kutta-Fehlberg method

The Runge-Kutta-Fehlberg method (denoted RKF45) is a numerical technique


used to solve ordinary differential equations of the form:

dr
= v(r(t), t) (A.I)
dt

Formal integration of A.I yields:


Z t+∆t
r(t + ∆t) = r(t) + v(r(t), t) dt (A.II)
t

The Runge-Kutta-Fehlberg method has a procedure to determine if the


proper step size h is being used. At each step, two different approximations
from both O(h4 ) and O(h5 ) methods for the solution are made and compared.
If the two results are in close agreement, the approximation is accepted. If
the two results do not agree to a specified accuracy, the step size is reduced.
If the results agree to more significant digits than required, the step size is
increased. A brief description is given below.
When integrating Equation A.II with the step size h using the Runge-Kutta-
Fehlberg method, let rk be the current position of a certain particle. Each step
requires the use of the following six values:

k1 = hv (rk , tk )
 
1 1
k2 = hv rk + k1 , tk + h
4 4
 
3 9 3
k3 = hv rk + k1 + k2 , tk + h
32 32 8 (A.III)
 
1932 7200 7296 12
k4 = hv rk + k1 − k2 + k3 , tk + h
2197 2197 2197 13
 
439 3680 845
k5 = hv rk + k1 − 8k2 + k3 − k4 , tk + h
216 513 4104
 
8 3544 1859 11 1
k6 = hv rk − k1 + 2k2 − k3 + k4 − k5 , tk + h
27 2565 4104 10 2

178
7.B. Interpolation methods

Then an approximation to the solution is made using a Runge-Kutta


method of order 4:
25 1408 2197 1
rk+1 = rk + k1 + k3 + k4 − k5 (A.IV)
216 2565 4101 5
In addition, a better value for the solution is determined using a Runge-
Kutta method of order 5:
16 6656 28561 9 2
r0 k+1 = rk + k1 + k3 + k4 − k5 + k6 (A.V)
135 12825 56430 50 55
The two approximations are then compared with each other. If the error
of approximation falls within the appropriate error bound, the approximation
will be accepted. Otherwise, an optimal step size will be computed based on
the two approximations.

7.B Interpolation methods

Trilinear interpolation is a fast and simple interpolation scheme for a three


dimensional Cartesian grid. It approximates the value of a point within the
rectangular grid cell linearly, using data on the corners of the cell. If point p is
in the cell shown in Figure B.1 and has the fractional offsets (dx, dy, dz) from
the grid point p1 , then:

f (dx, dy, dz) = f (p1 )(1 − dx)(1 − dy)(1 − dz) +


f (p2 )dx(1 − dy)(1 − dz) +
f (p3 )(1 − dx)dy(1 − dz) +
f (p4 )dxdy(1 − dz) +
f (p5 )(1 − dx)(1 − dy)dz +
f (p6 )dx(1 − dy)dz +
f (p7 )(1 − dx)dydz +
f (p8 )dxdydz (B.I)

In addition, Lekien and Marsden (2005) introduced a local tricubic interpo-


lation scheme in three dimensions that is both C1 and isotropic. The algorithm

179
7. Gas holdup and phase mixing in bubble column reactors

is based on a specific 64 × 64 matrix that gives the relationship between the


derivatives at the corners of the elements and the coefficients of the tricubic
interpolant for this element. A brief description of this method is as follows.

Figure B.1: Element for interpolation in three dimensions

As shown in the Figure B.1, the function f at the position p in the regular
grid can be represented by:

X
N X
N X
N
f (dx, dy, dz) = ai jk dxi dy j dzk (B.II)
i=0 j=0 k=0

The coefficients ai jk are determined in such a way that the function f is C1 ,


that is, f is continuous and its 3 first derivatives are also continuous.
The 64 coefficients aijk are stacked in a vector α by defining:

α1+i+4j+16k = aijk ∀i, j, k ∈ 0, 1, 2, 3 (B.III)

Similarly, the constraints on f and its derivatives are stacked in a vector b

180
References

by defining:






 f (pi ) if 1≤i≤8





 ∂

 9 ≤ i ≤ 16

 f (pi−8 ) if

 ∂x



 ∂

 17 ≤ i ≤ 24

 ∂y
f (pi−16 ) if





 ∂



 f (pi−24 ) if 25 ≤ i ≤ 32

 ∂z
bi = 
 (B.IV)

 ∂2

 f (pi−32 ) if 33 ≤ i ≤ 40

 ∂x∂y





 ∂2

 if 41 ≤ i ≤ 48

 f (pi−40 )

 ∂x∂z



 ∂2

 if 49 ≤ i ≤ 56

 ∂y∂z
f (pi−48 )





 ∂3



 f (pi−56 ) if 57 ≤ i ≤ 64
∂x∂y∂z

The derivatives of f can be computed and evaluated for the 8 points pi .


This gives a linear system in the 64 unknown coefficients αi :

Bα = b (B.V)

The matrix B is 64 × 64 and has only integer entries. The determinant of B


is 1. As a result, B is invertible and the coefficients αi can be computed using
the linear relationship:
α = B−1 b (B.VI)

The matrix B−1 is the core of the tricubic interpolator and can be found in
Lekien and Marsden (2005).

References

T. R. Auton. The lift force on a spherical body in a rotational flow. J. Fluid


Mech., 197:241–257, 1987.

181
7. Gas holdup and phase mixing in bubble column reactors

M. H. I. Baird and R. G. Rice. Axial dispersion in large unbaffled columns.


The Chemical Engineering Journal, 9(2):171–174, 1975.

D. Bauer, R. Peikert, M. Sato, and M. Sick. A case study in selective visu-


alization of unsteady 3d flow. In VIS ’02: Proceedings of the conference on
Visualization ’02, pages 525–528, Washington, DC, USA, 2002. IEEE Com-
puter Society.

R. S. Brodkey. Mixing: Theory and Practice, volume 1. Academic Press Inc.,


1972.

P. V. Danckwerts. Continuous flow systems : Distribution of residence times.


Chemical Engineering Science, 2(1):1–13, February 1953.

D. Darmana. On the multiscale modelling of hydrodynamics, mass transfer and


chemical reactions in bubble columns. PhD thesis, Enschede, 2006.

W. -D. Deckwer and A. Schumpe. Improved tools for bubble column reactor
design and scale-up. Chemical Engineering Science, 48(5):889–911, 1993.

W. -D. Deckwer, R. Burckhart, and G. Zoll. Mixing and mass transfer in tall
bubble columns. Chemical Engineering Science, 29(11):2177–2188, 1974.

E. Delnoij, F. A. Lammers, J. A. M. Kuipers, and W. P. M. van Swaaij. Dynamic


simulation of dispersed gas-liquid two-phase flow using a discrete bubble
model. Chemical Engineering Science, 52(9):1429–1458, 1997.

E. Delnoij, J. A. M. Kuipers, and W. P. M. van Swaaij. A three-dimensional


CFD model for gas-liquid bubble columns. Chemical Engineering Science, 54
(13-14):2217–2226, 1999.

J. R. Fair, A. J. Lambright, and J. W. Andersen. Heat transfer and gas holdup


in a sparged contactor. IandEC Process Design and Development, 1(1):33–36,
1962.

J. J. Heijnen and K. Van’t Riet. Mass transfer, mixing and heat transfer phe-
nomena in low viscosity bubble column reactors. The Chemical Engineering
Journal, 28(2):B21–B42, 1984.

182
References

H. Hikita and H. Kikukawa. Liquid-phase mixing in bubble columns: Effect


of liquid properties. The Chemical Engineering Journal, 8(3):191–197, 1974.

Andrea J. S. Hin and Frits H. Post. Visualization of turbulent flow with


particles. In VIS ’93: Proceedings of the 4th conference on Visualization ’93,
pages 46–53, Washington, DC, USA, 1993. IEEE Computer Society.

B. P. B. Hoomans. Granular dynamics of gas-solid two-phase flows. PhD thesis,


Enschede, January 2000.

B. P. B. Hoomans, J. A. M. Kuipers, W. J. Briels, and W. P. M. van Swaaij. Dis-


crete particle simulation of bubble and slug formation in a two-dimensional
gas-fluidised bed: A hard-sphere approach. Chemical Engineering Science,
51(1):99–118, 1996.

H. A. Jakobsen. Chemical Reactor Modeling: Multiphase Reactive Flows. Springer,


1st edition, July 2008.

J. B. Joshi. Axial mixing in multiphase contactors - A unified correlation.


Transactions of the Institution of Chemical Engineers, 58(3):155–165, 1980.

J. B. Joshi. Gas phase dispersion in bubble columns. The Chemical Engineering


Journal, 24(2):213–216, 1982.

J. B. Joshi, A. B. Pandit, K. L. Kataria, R. P. Kulkarni, A. N. Sawarkar, D. Tandon,


Y. Ram, and M. M. Kumar. Petroleum residue upgradation via visbreaking:
A review. Industrial and Engineering Chemistry Research, 47(23):8960–8988,
2008.

N. Kantarci, F. Borak, and K. O. Ulgen. Bubble column reactors. Process


Biochemistry, 40(7):2263–2283, 2005.

Y. Kawase and M. Moo-Young. Liquid phase mixing in bubble columns with


newtonian and non-newtonian fluids. Chemical Engineering Science, 41(8):
1969–1977, 1986.

Y. Kawase and M. Moo-Young. Theoretical prediction of gas hold-up in


bubble columns with newtonian and non-newtonian fluids. Industrial and
Engineering Chemistry Research, 26(5):933–937, 1987.

183
7. Gas holdup and phase mixing in bubble column reactors

R. Krishna, P.M. Wilkinson, and L.L. Van Dierendonck. A model for gas
holdup in bubble columns incorporating the influence of gas density on flow
regime transitions. Chemical Engineering Science, 46(10):2491–2496, 1991.

A. Kumar, T. E. Degaleesan, G. S. Laddha, and H. E. Hoelscher. Bubble swarm


characteristics in bubble columns. Canadian Journal of Chemical Engineering,
54(6):503–508, 1976.

D. A. Lane. Scientific visualization of large-scale unsteady fluid flows. In


Scientific Visualization, Overviews, Methodologies, and Techniques, pages 125–
145, Washington, DC, USA, 1997. IEEE Computer Society.

F. Lekien and J. Marsden. Tricubic interpolation in three dimensions. Interna-


tional Journal for Numerical Methods in Engineering, 63(3):455–471, 2005.

O. Levenspiel. Chemical Reaction Engineering. John Wiley & Sons, 3rd edition,
1999.

S. K. Majumder. Analysis of dispersion coefficient of bubble motion and


velocity characteristic factor in down and upflow bubble column reactor.
Chemical Engineering Science, 63(12):3160–3170, 2008.

K. -H. Mangartz and Th. Pilhofer. Interpretation of mass transfer measure-


ments in bubble columns considering dispersion of both phases. Chemical
Engineering Science, 36(6):1069–1077, 1981.

Y. Ohki and H. Inoue. Longitudinal mixing of the liquid phase in bubble


columns. Chemical Engineering Science, 25(1):1–16, 1970.

C. P. Ribeiro Jr. and P. L. C. Lage. Gas-liquid direct-contact evaporation: A


review. Chemical Engineering and Technology, 28(10):1081–1107, 2005.

H. Rusche. Computational fluid dynamics of dispersed two-phase flows at high phase


fractions. PhD thesis, Imperial College, London, 2002.

M. C. Ruzicka, J. Zahradnı́k, J. Drahos, and N. H. Thomas. Homogeneous-


heterogeneous regime transition in bubble columns. Chemical Engineering
Science, 56(15):4609–4626, 2001.

184
References

H. Scott Fogler. Elements of Chemical Reaction Engineering. Prentice Hall PTR,


4th edition, August 2005.

Y. T. Shah, B. G. Kelkar, S. P. Godbole, and W. -D. Deckwer. Design parameters


estimations for bubble column reactors. AIChE Journal, 28(3):353–379, 1982.

A. Shaikh and M. H. Al-Dahhan. A review on flow regime transition in bubble


columns. International Journal of Chemical Reactor Engineering, 5, 2007.

K. D. Squires and J. K. Eaton. Lagrangian and eulerian statistics obtained from


direct numerical simulations of homogeneous turbulence. Physics of Fluids
A, 3(1):130–143, 1991.

A. Tomiyama, T. Matsuoka, T. Fukuda, and T. Sakaguchi. A simple numerical


method for solving an incompressible two-fluid model in a general curvi-
linear coordinate system. Advances in Multiphase Flow, pages 241–252,
Amsterdam, 1995. Society of Petroleum Engineers Inc.

A. Tomiyama, I. Kataoka, I. Zun, and T. Sakaguchi. Drag coefficients of single


bubbles under normal and micro gravity conditions. JSME international
journal. Ser. B, Fluids and thermal engineering, 41(2):472–479, 1998.

A. Tomiyama, H. Tamai, I. Zun, and S. Hosokawa. Transverse migration of


single bubbles in simple shear flows. Chemical Engineering Science, 57(11):
1849–1858, 2002.

A. W. Vreman. An eddy-viscosity subgrid-scale model for turbulent shear


flow: Algebraic theory and applications. Physics of Fluids, 16(10):3670–3681,
2004.

S. Wachi and Y. Nojima. Gas-phase dispersion in bubble columns. Chemical


Engineering Science, 45(4):901–905, 1990.

R. M. Wellek, A. K. Agrawal, and A. H. P. Skelland. Shape of liquid drops


moving in liquid media. AIChE Journal, 12(5):854–862, 1966.

K. R. Westerterp, W. P. M. van Swaaij, and A. A. C. M. Beenackers. Chemical


Reactor Design and Operation. John Wiley & Sons Ltd., 2nd edition, 1987.

185
7. Gas holdup and phase mixing in bubble column reactors

P. K. Yeung and S. B. Pope. An algorithm for tracking fluid particles in


numerical simulations of homogeneous turbulence. Journal of Computational
Physics, 79(2):373–416, 1988.

P. Zehner. Momentum, mass and heat transfer in bubble columns. part 2. axial
blending and heat transfer. International chemical engineering, 26(1):29–35,
1986.

186
Summary
Experimental and Numerical Investigation of Bubble Column
Reactors

Due to various advantages, such as simple geometry, ease of operation, low


operating and maintenance costs, excellent heat and mass transfer character-
istics, bubble column reactors are frequently used in chemical, petrochemical,
biochemical, pharmaceutical, metallurgical industries for a variety of pro-
cesses, i.e. hydrogenation, oxidation, chlorination, alkylation, chemical gas
cleaning, various bio-technological applications, etc. However, complex hy-
drodynamics and its influence on transport phenomena (i.e. heat and mass
transport) make it difficult to achieve reliable design and scale-up of bubble
column reactors. Many factors influence the performance of this type of re-
actors significantly, such as column dimensions, column internals design, gas
distributor design, operating conditions, i.e. pressure and temperature, su-
perficial gas velocity, physical and chemical properties of the involved phases.
A large variety of scientific studies on bubble column reactors utilizing both
experimental and numerical techniques has been carried out during the past
decades. In this study a bubble column with a square cross-sectional area
has been studied in detail using a combined experimental and computational
approach.
Chapter 1 introduces bubble column reactors and their variants according
to practical requirements. Both advantages and disadvantages of these types
of reactors are presented. Key parameters related to the performance of bubble
column reactors are also presented. In addition, in Chapter 1 a brief literature
review is presented on both experimental and numerical techniques utilized
in investigations on the performance of bubble column reactors during the
past decades.

187
Summary

In Chapter 2, accuracy of a four-point optical fibre probe for measuring


bubble properties is investigated in a flat bubble column. Photography is used
to validate results obtained from the four-point optical fibre probe. According
to the comparison, it is found that the liquid properties have a profound
influence on bubble velocity measured by the optical probe. Finally, it is
found that the extent of inaccuracy in the determination of bubble velocity
can be characterized with the Morton number.
The accuracy of the four-point optical fibre probe and its intrusive effect are
further studied in a square bubble column operating at higher superficial gas
velocity in Chapter 3. Besides bubble velocity, other bubble properties, such as
local void fraction, chord length and specific interfacial area, are obtained from
measurements with the four-point optical fibre probe. Furthermore, bubble
size is determined in different ways. Possible reasons for the discrepancy in
the bubble size determination are discussed. The effect of the initial liquid
height in the bubble column on the bubble properties is also investigated.
Chapter 4 studies the effect of the gas sparger properties on the hydrody-
namics in a square bubble column with an Eulerian-Lagrangian model. The
performance of the model is first evaluated by comparison with experimental
data. Subsequently, the effects of different sparged areas and the sparger loca-
tion on hydrodynamics, i.e. liquid velocity, turbulent kinetic energy and void
fraction are investigated. Furthermore, the residence time distribution of the
gas phase is extracted from the numerical simulations. These distributions are
used to characterize the gas phase mixing in the bubble column by employing
a standard axial dispersion model. The results reveal that the extent of mixing
increases when the sparged area decreases. The axial dispersion coefficient
increases as the sparged area is shifted towards the side wall.
For numerical simulation of bubbly flows, reliable closures are required to
represent the interfacial momentum transfer rate (i.e. the effective drag acting
on bubbles). Furthermore, the presence of neighboring bubbles in a bubble
swarm may result in deviation of the drag force acting on isolated bubbles.
Chapter 5 investigates the performance of several drag correlations reported
in literature for bubble swarms with the aid of a discrete bubble model. By
comparing with experimental data, it is found that Lima Neto’s drag model

188
and Wen & Yu’s model have a better performance at low superficial gas
velocity and Rusche’s model can predict the hydrodynamics of the bubbly
flows better compared to the other models at high superficial gas velocity.
In Chapter 6, breakup models developed in literature are implemented
into the Eulerian-Lagrangian model. Moreover, the critical Weber number
for bubble breakup studied by many authors in turbulent flows is also in-
corporated in the model. The performance of different breakup models and
the critical Weber number for predicting hydrodynamics and the bubble size
distribution are compared with experimental data.
Finally, the Eulerian-Lagrangian model is further extended to study the
performance of bubble column reactors, i.e. predicting overall gas holdup
and phase mixing in Chapter 7. The residence time distribution of the gas
phase and tracer particles introduced in the liquid phase are used to study the
mixing of both the gas and liquid phase. It is found that the applied model
shows very good agreement with empirical correlations reported in literature.

189
Samenvatting
Experimentele en Numerieke Studie van Bellenkolom
Reactoren

Vanwege de vele voordelen, zoals eenvoudige geometrie, eenvoudige oper-


atie, lage operatie- en onderhoudskosten, uitstekende warmte- en stofover-
dracht karakteristieken worden bellenkolom reactoren veelvuldig gebruikt
in de chemische, petrochemische, biochemische en metallurgische industrie
voor diverse processen, zoals hydrogenering, oxidatie, chlorering, alkyler-
ing, chemische gas zuivering, diverse biochemische toepassingen, etc. Echter,
de complexe hydrodynamica en de invloed hiervan op de transport verschi-
jnselen (d.w.z. warmte- en stofoverdracht) bemoeilijken een betrouwbaar
ontwerp en opschaling van bellenkolommen.
Er zijn veel factoren die grote invloed hebben op de prestaties van dit
type reactoren, zoals kolom afmetingen, het ontwerp van kolom internals,
gasverdeler ontwerp, operatiecondities, d.w.z. druk en temperatuur, gassnel-
heid, fysische en chemische eigenschappen van de verschillende fasen. De
afgelopen decennia is een grote verscheidenheid aan wetenschappelijke stud-
ies over bellenkolom reactoren uitgevoerd, waarbij gebruik werd gemaakt van
zowel experimentele als numerieke technieken. In dit werk wordt een bel-
lenkolom met een vierkante dwarsdoorsnede in detail bestudeerd met behulp
van een gecombineerde experimentele en numerieke aanpak.
In hoofdstuk 1 worden bellenkolom reactoren en hun varianten geı̈ntro-
duceerd aan de hand van praktische randvoorwaarden. Zowel voor- en
nadelen van dit soort reactoren worden gepresenteerd en belangrijke param-
eters worden gerelateerd aan het functioneren van bellenkolom reactoren.
Tenslotte bevat hoofdstuk 1 een beknopte literatuurstudie van experimentele
en numerieke technieken die de afgelopen decennia gebruikt zijn in het on-

191
Samenvatting

derzoek naar het functioneren van bellenkolom reactoren.


In hoofdstuk 2 wordt de nauwkeurigheid van een vier-punts glasvezel
probe onderzocht voor het meten van beleigenschappen in een platte bel-
lenkolom. Digitale fotografie is gebruikt om de resultaten verkregen uit
de probe te valideren. Uit deze vergelijking is geconstateerd dat de
vloeistofeigenschappen een grote invloed hebben op de door de probe geme-
ten belsnelheid. Ten slotte is geconstateerd dat de onnauwkeurigheid in de
bepaling van de belsnelheid kan worden gekarakteriseerd door het Morton
kental.
De nauwkeurigheid van de glasvezel probe en de mate van verstoring van
de vloeistof bij een hogere gassnelheid zijn verder onderzocht in een vierkante
bellenkolom in hoofdstuk 3. Naast de belsnelheid zijn andere beleigenschap-
pen, zoals de lokale gasfractie, de koordelengte en het specifieke oppervlak
verkregen uit metingen met de glasvezel probe. Bovendien is de belgrootte
bepaald op verschillende manieren. Mogelijke redenen voor de discrepantie
in de belgrootte bepaling worden besproken. De invloed van de initiële
vloeistofhoogte in de kolom op de beleigenschappen is ook onderzocht.
In hoofdstuk 4 wordt een Euler-Lagrange model gebruikt om het effect van
de eigenschappen van de gasverdeler op de hydrodynamica in een vierkante
bellenkolom te bestuderen. De prestaties van het model worden eerst geëval-
ueerd aan de hand van vergelijking met experimentele data. Vervolgens
worden de effecten van verschillende afmetingen en positionering van de
gasverdeler op de hydrodynamica, d.w.z. de vloeistofsnelheid, turbulente
kinetische energie en de gasfractie onderzocht. Bovendien zijn de verblijfti-
jdspreidingen van de gasfase uit de numerieke simulaties verkregen. Deze
worden gebruikt om de menging van de gasfase in de bellenkolom te karak-
teriseren aan de hand van een standaard axiaal dispersie model. De resultaten
tonen aan dat de mate van menging toeneemt wanneer het oppervlak van de
gasverdeler afneemt. De axiale dispersie coëfficiënt neemt toe naarmate de
gasverdeler verder naar de zijwand wordt verschoven.
Voor numerieke simulatie van bellenstromen, zijn betrouwbare sluit-
ingsrelaties nodig voor de impulsuitwisseling tussen de fasen (d.w.z. de
effectieve wrijvingskracht op bellen). Bovendien kan de aanwezigheid van

192
naburige bellen in een bellenzwerm resulteren in een afwijking op de wrijv-
ingskracht die geldt op individuele, geı̈soleerde bellen. In hoofdstuk 5 worden
de prestaties van verschillende wrijvingskrachtcorrelaties voor bellenzwer-
men uit de literatuur onderzocht met behulp van een discrete bellenmodel.
Aan de hand van een vergelijking met experimentele data is vastgesteld dat
bij lage gassnelheden de wrijvingskrachtcorrelaties van Lima Neto en Wen
& Yu de beste voorspellingen geven, terwijl bij hoge gassnelheden Rusche’s
correlatie de beste resultaten geeft.
In hoofdstuk 6 zijn literatuurmodellen voor het opbreken van bellen
geı̈mplementeerd in het Euler-Lagrange model. Bovendien zijn verschillende
literatuurmodellen voor het kritische Weber kental bij opbreken van bellen
in turbulente stromingen verdisconteerd in het model. De prestaties van de
verschillende modellen voor het opbreken van bellen en het kritische Weber
kental in termen van het voorspellen van hydrodynamica en de belgroot-
teverdeling zijn vergeleken met experimentele data.
Ten slotte is in hoofdstuk 7 het Euler-Lagrange model verder uitgebreid
om de prestaties van de bellenkolom reactoren te bestuderen, in het bijzonder
voor het voorspellen van de totale gasfractie en het mengen van beide fasen.
De verblijftijdverdeling van de gasfase en tracer deeltjes geı̈ntroduceerd in de
vloeistoffase worden gebruikt om het mengen van zowel de gas- als vloeistof-
fase te bestuderen. Het is gebleken dat de door het model voorspelde resul-
taten zeer goed overeenkomen met empirische correlaties gerapporteerd in
de literatuur.

193
总结
鼓泡塔反应器的实验与数值研究

由于具有结构简单、易操作、低运行与维护成本、高传热传质性质等优点,
鼓泡塔反应器被广泛地应用在化学、石油化学、生物化学、制药、冶金等工
业中的各种过程中,例如制氢、氧化、烷基化、化学气体清洗以及各种生物
技术应用等。然而,复杂的流体力学特征及其对传输性质(例如传热传质)
的影响,使得很难获得可靠设计和按比例放大的鼓泡塔反应器。很多因素,
极大地影响这类反应器的性能,例如反应器尺寸、反应器内部设计、气体分
布器设计、运行条件(如压力、温度)、表观气速以及各相的物理与化学性
质。在过去几十年里,人们做了大量关于鼓泡塔反应器的科学研究,运用了
实验与模拟技术。本研究利用实验与模拟技术详细地研究了一个具有方形截
面的鼓泡塔。
第一章介绍了鼓泡塔反应器及其根据不同实际需求的变体,以及此类反应
器的优缺点。并且介绍了与鼓泡塔反应器性能相关的关键参数。此外,第一
章中,文献综述简要地介绍了在过去几十年里,研究鼓泡塔反应器性能的实
验以及模拟技术。
在第二章里,研究了用于测量气泡性质的四针光纤探针的准确度。高速照
相技术所获得的气泡性质用于比较四针光纤探针的测量结果。比较发现,液
体物性显著地影响着探针所测量的气泡速度。最后,通过Morton数,描述了
利用探针测量气泡速度的不准确程度。
第三章通过一个方形鼓泡塔,进一步地研究了在较高的表观气速下,光
纤探针测量的准确度及其侵入性的影响。除了气泡速度,四针光纤探针还测
量了其他气泡性质,例如局部气含率、气泡弦长和比界面积。此外,通过不
同的方法,确定了气泡的尺寸。并且讨论了造成确定气泡尺寸差异的可能原
因。最后,研究了鼓泡塔内初始液体高度对气泡性质的影响。
第四章利用欧拉-拉格朗日模型研究了气体分布器对鼓泡塔内流动性质的影
响。首先,通过比较实验数据,验证了模型。随后,利用模型研究了不同的

195
总结

鼓泡区域对流体力学性质,如液体速度,湍动能及气含率的影响。此外,利
用模拟结果,获得了气相驻留时间分布。通过轴向分布模型,利用气相驻留
时间分布,描述了鼓泡塔内的气相混合。结果显示,混合程度随着鼓泡区域
的减小而变大。鼓泡区域越靠近侧壁,轴向分布系数越大。
泡状流的数值模拟需要可靠的封闭模型来描述界面动量传输率,例如作用
在气泡上的有效曳力。此外在大量气泡里,由于周围气泡的影响,作用在气
泡上的曳力可能会不同于作用在单个气泡上的曳力。第五章借助离散气泡模
型,研究了文献中的一些关于气泡群中的曳力关联式。通过比较实验数据发
现,在低表观气速下,Lima Neto以及Wen和Yu的模型优于其他模型。而在
高表观气速下,同其他模型相比,Rusche的模型能够较好地预测泡状流的流
动。
第六章将文献中发展的破碎模型应用到欧拉-拉格朗日模型。此外,湍流里
关于气泡破碎的临界Weber数也被应用到了此模型中。通过比较实验数据,研
究了这些不同破碎模型以及临界Weber数预测流动和气泡尺寸分布的性能。
最后,在第七章中,欧拉-拉格朗日模型被进一步扩展到用于研究鼓泡塔反
应器的性能,例如预测整体气含率与相混合。气相驻留时间分布和液相中的
示踪粒子被用于研究气相和液相的混合。结果表明,模型所预测的结果同文
献中的实验关联式吻合得很好。

196
List of publications
Publications

• W. Bai, N. G. Deen, J.A.M. Kuipers. Numerical analysis of the effect of


gas sparger on the performance of bubble column, Ind.Eng.Chem.Res.,
submitted.

• W. Bai, Y. M. Lau, N. G. Deen, J. A. M. Kuipers. Discrete bubble modeling


of bubbly flows: Swarm effects, Chem.Eng.J., submitted.

• W. Bai, Y. M. Lau, N. G. Deen, J. A. M. Kuipers. Discrete bubble modeling


of bubbly flows: Implementation of breakup models, Ind.Eng.Chem.Res.,
in preparation.

• W. Bai, N. G. Deen, J. A. M. Kuipers. Numerical investigation of gas


holdup and phase mixing in bubble column reactors, Ind.Eng.Chem.Res.,
in preparation.

Conference papers

• W. Bai, N. G. Deen, and J. A. M. Kuipers. Bubble properties of heteroge-


neous bubbly flows in a square bubble column, International Symposium
of Multiphase Flows, July 11-15, 2009, Xi’an, China.

• W. Bai, N. G. Deen, R. F. Mudde and J. A. M. Kuipers. Accuracy of four-


point optical probe for determination of bubble velocity, 6th International
Conference on Computational Fluid Dynamics in the Oil & Gas, Metallurgical
and Process Industries, Trondheim, Norway 10-12 June 2008.

197
Acknowledgement
In this thesis the results of a four-year research study are presented, which
was conducted at the research group Fundamentals of Chemical Reaction
Engineering, Faculty of Science and Technology, University of Twente, the
Netherlands. I’d like to thank the Institute of Mechanics, Processes and Con-
trol - Twente (IMPACT) for the financial support to the project. Furthermore,
I would like to express my sincere thanks to those who have contributed their
help, support and care to both my work and my life in Enschede during the
past four years.
First of all, I would like to convey my thanks to my promotor, Prof. Hans
Kuipers. Your careful attitude on research and profound knowledge of process
engineering have made a strong impression on me. Your patient guidance
during monthly meetings and every group activity organized by your family
will be part of my beautiful memories for the rest of my life.
A special thank to my co-promotor, Dr. Niels Deen, who acted as a daily
supervisor. I couldn’t forget our fruitful weekly discussions, your quick-
witted mind and the encouragement you gave. Thank you for being there to
help me always. I would also like to thank you for all the grammar checking
of the thesis, the Dutch translation of the summary and all other support you
offered. I wish both Hans and you great success at Eindhoven University of
Technology.
I would like to thank the secretary of the group, Nicole Haitjema, for
preparing all the paperwork for my appointment as a PhD student and ar-
ranging housing for me before I came to Enschede. Everything you have done
for me during the past four years is highly appreciated.
Sincere thanks to Gerrit Schorfhaar. I’d like to thank you for constructing
the experimental setup for me. I also want to thank you for the fruitful
conversations and the sharing of your travelling experience and your life.

199
Acknowledgement

I’m impressed that you can still run 5 kilometers twice a week as a senior
person. I’d like to thank the other technicians in the group, Wim Leppink,
Johan Agterhorst and Erik Analbers for their help and kindness. I would also
like to thank Robert Meijer for his help on PC installing, email settings and
everything related to electronics.
I’d like to send my gratitude to the former group members for their sup-
port and kindness: Mao Ye, Christiaan Zeilstra, Dadan Darmana, Dongsheng
Zhang, Sabita Sarkar, Wouter Dijkhuizen, Tymen Tiemersma, Sander Noor-
man, Willem Godlieb, Jan Albert Laverman and Micheline Abbas. Many
thanks to warm-hearted Mao and Chris for their enthusiastic help when I first
came here. Special thanks to Dongsheng for his helping hand and inviting
me for dinners and outings with his family. I would also like to thank all the
current members of the group: Prof. Martin van Sint Annaland, Dr. Martin
van der Hoef, Maureen van Buijtenen, Nhi Dang, Jelle de Jong, Tom Kolk-
man, Sebastian Kriebitzsch, Yuk Man Lau, Ivo Roghair, Lianghui Tan, Martin
Tuinier, Olasaju Olaofe and Carles Mesado Melia for bringing a friendly and
creative atmosphere, nice X’mas parties, a memorable sailing event, borrels
and outings. Many thanks to Ivo and Tom for helping me on solving prob-
lems with clusters and other things. Thank Yuk Man for discussing the DBM
Code with me and offering other help. I’d also like to thank Martin (小马) for
talking to me in Chinese and offering his help.
I would like to extend my sincere thanks to Fausto Gallucci and his family.
The delicious Italian food and the hospitality of his family have been embed-
ded into my memory. Many thanks to Junwu Wang for the comments on my
work and the nice food at his place.
Many thanks to Wei Zhao & Wei Zhou(F), Yang Zhang & Rui Ge, Wei
Zhou(M), Xinhui Wang & Xuehui He, Xin Wang, Pengxiang Xu & Xia Bian, Yan
Song, Rui Guo and Hongmei Zhang for those happy times of dinner gathering,
table tennis and dizhu fighting. I would like to thank my former roommates
Xuexin Duan, Songyue Chen and Lixian Xu for the warm atmosphere at home,
nice conversation and pleasant dinner. I also thank Songyue’s husband Yifan
He and Lixian’s family, Hongping Luo and their son, Beibei for unforgettable
memories. I would like to thank Huaping Xu for enjoying chats and beer with

200
me. Thank Xiaoying Shao and Shu-Han Hsu for inviting me for dinners and
the trip to Greece. I also thank all the other friends: Rongmei Li, Yixuan Li,
Weihua Zhou, Xia Shang, Peng Zhang, Chunlin Song, Jinping Han, Yu Song,
Yu Zeng and Xinyan Wen.
Special thanks to Ying Zhang and her parents for taking care of me and
treating me as part of their family. I am grateful to Jie Fan for designing the
cover of my thesis.
最后,特别感谢我的家人多年来对我的关心和支持:父母业已年迈,尚未
回报丁点养育之恩,深感内疚,而父母仍理解并支持儿子的求学之情。感谢
我的哥哥、嫂子和姐姐、姐夫,他们给我了很大的关心,并赡养父母,使得
我不论在何处学习或工作,均能够安心于学业与工作。感谢我的小外甥和小
侄女,给我带来的开心和快乐。诚挚感谢周彦婷小姐所给予的关心以及在英
文写作上的帮助,使得我能够充满信心地完成论文。

柏巍 Wei Bai
September 2010, Ames

201
About the author
Wei Bai was born on March 14th 1977 in Hanzhong, Shaanxi, China. After
completing his secondary education in 1996 at Mian county’s 1st high school
in Shaanxi, China, he continued his study at Xi’an Jiaotong University, China
and obtained his bachelor’s degree in Heating, Ventilation and Air Condition-
ing Engineering on the subject of “Heating system design of high buildings”
in 2000. In 2001, he started his master study under the supervision of Prof.
Qiuwang Wang at Xi’an Jiaotong University. He received the master degree in
Engineering Thermophysics with a thesis titled ”Three-dimensional numeri-
cal simulation of periodical flow and heat transfer in primary surface channel”
in 2004. He then joined the Division of Building Science and Technology as a
research assistant at City University of Hong Kong, China. From 2006 to 2010,
he started his PhD research and was supervised by prof.dr.ir. J.A.M. Kuipers
and dr.ir. Niels Deen in the Fundamentals of Chemical Reaction Engineering
group at University of Twente, the Netherlands. The results of this research
are presented in this dissertation. In June 2010, he joined Prof. Rodney O.
Fox’s group as a post doc at Iowa State University in the U.S.A.

203

Вам также может понравиться