You are on page 1of 208

Detailed Simulations of Liquid and

Solid-Liquid Mixing

Turbulent agitated flow and mass transfer


DETAILED SIMULATIONS OF LIQUID AND
SOLID-LIQUID MIXING

Turbulent agitated flow and mass transfer

PROEFSCHRIFT

ter verkrijging van de graad van doctor


aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof. dr ir J. T. Fokkema,
voorzitter van het College voor Promoties,
in het openbaar te verdedigen op
vrijdag 18 november 2005 om 10.30 uur

door

Hugo HARTMANN

natuurkundig ingenieur
geboren te Zuidelijke IJsselmeerpolders
Dit proefschrift is goedgekeurd door de promotor:
Prof. dr ir H. E. A. van den Akker

Samenstelling promotiecommissie:

Rector Magnificus, voorzitter


Prof. dr ir H. E. A. van den Akker, Technische Universiteit Delft, promotor
Prof. dr M. Yianneskis, Kings College London, Verenigd Koninkrijk
Prof. dr C. D. Rielly, Loughborough University, Verenigd Koninkrijk
Prof. dr ir P. Wesseling, Technische Universiteit Delft
Prof. dr D. J. E. M. Roekaerts, Technische Universiteit Delft
Dr ir J. J. Derksen, Technische Universiteit Delft
Prof. dr ir M. M. C. G. Warmoeskerken, Universiteit Twente

The project OPTIMUM was financially supported by the Commission of the European
Union under the program Promoting Competitive and Sustainable Growth, Contract
G1RD-CT-2000-00263.

Printed by:
Ponsen & Looijen B.V.
Nudepark 142
6702 DX Wageningen The Netherlands
http://www.p-l.nl
ISBN 90-6464-234-6

Keywords: stirred tank, simulations, turbulence, scalar mixing, solids suspension, mass
transfer.

2005
c by Hugo Hartmann.

All rights reserved. No part of the material protected by this copyright notice may be
reproduced or utilized in any form or by any means, electronic or mechanical, includ-
ing photocopying, recording or by any information storage and retrieval system without
written permission from the publisher.
Zoek uw kracht in de Heer,
in de kracht van zijn macht.

Efezirs 6:10 (De Nieuwe Bijbelvertaling)

Aan Mirjam en Lon


Contents

Summary xi

Samenvatting xv

1 Introduction 1
1.1 Mixing in the process industries . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Two mixing applications; OPTIMUM project . . . . . . . . . . . . . . . 3
1.2.1 Motivation, aim and project partners . . . . . . . . . . . . . . . . 3
1.2.2 Suspension polymerization . . . . . . . . . . . . . . . . . . . . . 5
1.2.3 Particle coating . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Mixing and LBM at the Kramers Laboratory . . . . . . . . . . . . . . . . 11
1.4 Motivation and aim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.5 Stirred tank geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.6 Outline of this thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2 Assessment of large eddy simulations on the flow in a stirred tank 19


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 Stirred vessel configuration . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 Experimental method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.4 Computational method . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4.1 RANS simulation . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4.2 Large eddy simulation . . . . . . . . . . . . . . . . . . . . . . . 25
2.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.5.1 Behavior of the subgrid scale models . . . . . . . . . . . . . . . 28
2.5.2 Phase-averaged flow field and turbulence levels . . . . . . . . . . 29
2.5.3 Phase-resolved flow field and turbulence levels . . . . . . . . . . 34
2.5.4 Energy dissipation . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.5.5 Turbulence anisotropy . . . . . . . . . . . . . . . . . . . . . . . 40

vii
2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3 Macro-instability uncovered in a Rushton turbine stirred tank 47


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.1.1 Flow phenomena in stirred tanks . . . . . . . . . . . . . . . . . . 47
3.1.2 MI significance to multi-phase chemical processes . . . . . . . . 48
3.1.3 RANS simulation . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.1.4 Large eddy simulation . . . . . . . . . . . . . . . . . . . . . . . 50
3.2 Flow system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.3 Simulation procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.3.1 Large Eddy Simulation . . . . . . . . . . . . . . . . . . . . . . . 51
3.3.2 Simulation aspects . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.4.1 Single flow field realizations . . . . . . . . . . . . . . . . . . . . 54
3.4.2 Time-series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.4.3 Frequency analysis . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.4.4 Effect of Reynolds number and impeller off-bottom clearance . . 60
3.4.5 Pseudo turbulence determination procedure . . . . . . . . . . . . 61
3.4.6 Kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.4.7 Grid size effects . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4 Development and validation of a scalar mixing solver based on finite


volume discretization 75
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2 Finite volume discretization of the convection diffusion equation . . . . . 76
4.3 Time discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.4 The treatment of convection . . . . . . . . . . . . . . . . . . . . . . . . 78
4.4.1 Monotone schemes . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.4.2 TVD discretization scheme . . . . . . . . . . . . . . . . . . . . . 80
4.4.3 TVD multidimensional application . . . . . . . . . . . . . . . . 82
4.5 The fully discretized convection diffusion equation . . . . . . . . . . . . 83
4.6 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.6.1 Neumann boundary condition implementation in 1-D . . . . . . . 86
4.6.2 Extension to multiple dimensions . . . . . . . . . . . . . . . . . 87

viii
4.7 Validations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.7.1 Laminar cavity flow in a square box . . . . . . . . . . . . . . . . 88
4.7.2 Laminar cavity flow in a box with inclined side walls . . . . . . . 92
4.7.3 Mixing behavior in a cylindrical tank with a side-entry mixer . . . 95
4.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5 A parameter study of the mixing time in a turbulent stirred tank by


means of LES 105
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.1.1 Experimental research on turbulent scalar mixing . . . . . . . . . 105
5.1.2 Potential of CFD on scalar mixing . . . . . . . . . . . . . . . . . 107
5.2 Flow system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.3 Simulation procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.3.1 Large eddy simulation . . . . . . . . . . . . . . . . . . . . . . . 109
5.3.2 Scalar transport . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.3.3 Immersed boundary technique . . . . . . . . . . . . . . . . . . . 112
5.3.4 Simulation aspects . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.4.1 Snapshots of the scalar concentration field . . . . . . . . . . . . . 115
5.4.2 Time series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.4.3 Mixing time and coefficient of mixing . . . . . . . . . . . . . . . 121
5.4.4 Mixing time correlations . . . . . . . . . . . . . . . . . . . . . . 125
5.4.5 The Ruszkowski (1994) mixing time correlation . . . . . . . . . 127
5.4.6 Performance immersed boundary technique, mass conservation . 131
5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

6 Numerical simulation of a dissolution process in a stirred tank reactor139


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
6.2 Flow system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.3 Simulation procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
6.3.1 Flow solver . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
6.3.2 Particle transport solver . . . . . . . . . . . . . . . . . . . . . . 144
6.3.3 Mass transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
6.3.4 Scalar mixing solver . . . . . . . . . . . . . . . . . . . . . . . . 146
6.3.5 Simulation aspects . . . . . . . . . . . . . . . . . . . . . . . . . 147

ix
6.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.4.1 Snapshots of the particle distributions and concentration fields . . 148
6.4.2 Stages in the dissolution process . . . . . . . . . . . . . . . . . . 154
6.4.3 Evolution of particle size distribution in time . . . . . . . . . . . 156
6.4.4 Dissolution time . . . . . . . . . . . . . . . . . . . . . . . . . . 158
6.4.5 Time series of concentration . . . . . . . . . . . . . . . . . . . . 159
6.4.6 Mass conservation finite volume scheme . . . . . . . . . . . . . . 160
6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

7 Conclusions and perspectives 167


7.1 General discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
7.2 Turbulent flow phenomena . . . . . . . . . . . . . . . . . . . . . . . . . 168
7.3 Scalar mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
7.4 Solid-liquid mixing including mass transfer . . . . . . . . . . . . . . . . 170
7.5 Perspectives and recommendations . . . . . . . . . . . . . . . . . . . . . 171

Bibliography 175

Dankwoord 185

List of publications 189

About the author 190

x
Summary

Mixing operations involving turbulent flow are widely applied in the process industries for
bringing two or three phases into intimate contact to achieve inter-phase heat and mass
transfer, chemical reactions, etc. Although mixing is only part of the whole process, it
often has a significant impact on the product quality, process reliability and economy.
An inadequate understanding of mixing costs a few billion dollars per year in the USA
alone. Furthermore, the lack of insight in the hydrodynamic phenomena involved and
their coupling with chemistry, and heat and mass transfer must be overcome to achieve
competitiveness improvement. In the last decades, experiments have contributed signifi-
cantly to the understanding of the flow phenomena and processes, and to better design and
operation methods. However, parameters dominating the smallest scales are still hardly
accessible for experimental techniques. A promising alternative gaining more attention in
the recent years, is to resolve the relevant turbulence quantities by simulation.
This work was part of the European project OPTIMUM in which experimental and nu-
merical tools have been developed and employed to increase the understanding of a sus-
pension polymerization and a particle coating process. Through addressing this topic at a
European level, know-how on the specific chemical reactions and products of the partic-
ular processes, fundamental information on hydrodynamic phenomena, and expertise on
experimental and numerical techniques were combined.
The main aim of the present work arose from the OPTIMUM project and was formulated
as a contribution to reliable and accurate numerical predictions of complex, multi-phase
processes. For this purpose, three issues were treated successively: (i) turbulent flow
phenomena, (ii) scalar mixing, and (iii) solid-liquid mixing including mass transfer. The
flow geometry used throughout this work is a baffled tank equipped with a Rushton turbine
operated under turbulent conditions.
A detailed and accurate representation of the flow field is the first step toward a successful
simulation of a complex, multi-phase process. For this purpose, a large eddy simulation
(LES) was used. The flow solver based on the lattice-Boltzmann discretization scheme
was critically assessed in order to get confidence in the method. Two subgrid-scale mod-

xi
els were compared. The focus was on the accuracy of the LES flow field predictions (in
terms of velocity, turbulent kinetic energy, energy dissipation, and turbulence anisotropy),
and the recovery of coherent velocity fluctuations induced by large-scale precessing vor-
tices (i.e. a so-called macro-instability). The important result from this assessment was
that LES predicted the turbulence levels to a high level of accuracy, which contrasts with
the underprediction of 50% due to a simulation based on the RANS (Reynolds-Averaged
Navier-Stokes) approach. This result is of high importance as e.g. the mixing patterns,
the motion of particles, and inter-phase mass transfer are highly affected by the velocity
fluctuations.
In order to describe the mixing performance of the turbulent flow generated by the impeller
or inter-phase mass transfer between the continuous and disperse phases, information of
a scalar concentration in the continuous phase is required. Therefore, a solver for the
description of scalar transport was developed. An Eulerian approach based on the finite
volume method has been followed. Subsequently, attention was paid to the implemen-
tation of complexly shaped boundaries which are in general off-grid. The zero-gradient
constraint at these walls was imposed by a newly developed immersed boundary tech-
nique. The scalar mixing solver was assessed in various laminar and turbulent flow cases.
Subsequently, LES including scalar transport have been carried out on a blending process.
The focus was on the mixing time as a function of impeller size and injection position of
the passive tracer. An important observation was that in the literature there is no standard-
ization of mixing time experiments. In literature, measured or simulated mixing times are
often not carefully compared with data and correlations. Our simulations, however, make
it possible to relate all data and correlations properly. The predicted mixing times agreed
within 30% with experimental results and values obtained from two correlations presented
in the literature. According to the simulations, the mixing time scales with the tank over
impeller diameter ratio to the power 2.5. The mixing time was not significantly influenced
by the position of the feed point.
Finally, a particle transport solver was coupled to the flow and scalar mixing solvers to
carry out a simulation of a solid-liquid suspension including mass transfer. The particle
transport solver is based on a Lagrangian approach in which spherical, solid particles are
tracked in the Eulerian flow field through solving the dynamic equations of linear and rota-
tional motion of the particles. Particle-particle and particle-wall collisions were included.
The force exerted by the fluid on the particle was fed back to the fluid (two-way coupling).
A solubility process was chosen to serve as the benchmark case. The simulation was re-
stricted to a lab-scale vessel with a volume equal to 10 liter. A set of 7 million spherical
particles was released in the top part of the vessel. At the moment of release, the local vol-
ume fraction amounted to 10%, which corresponds to a vessel averaged volume fraction of

xii
1%. The particle properties were such that they resemble those of calcium-chloride beads.
The focus was on the dispersion of the particles throughout the vessel, the solubility rate
in terms of solids and scalar concentration distributions, particle size distributions, and the
solubility time. For the particular process considered, various stages were identified, and
the solubility time was found to be at most one order of magnitude higher than the time
needed to disperse the solids throughout the vessel. Initially, the particles organized in
streaky patterns. Due to decreasing inertia of the particles, they were behaving more like
fluid tracers. Till the point the particles were dispersed throughout the vessel, the spatial
solids and scalar concentration distributions were very inhomogeneous. Subsequently, the
distributions were becoming more homogeneous.

xiii
Samenvatting

In de procesindustrie wordt uitgebreid gebruik gemaakt van menging om intiem contact


tussen twee of drie fasen te bewerkstelligen ten behoeve van stofoverdracht, chemische re-
acties en warmtetransport. Hoewel menging maar een onderdeel is van het gehele proces,
heeft menging een belangrijke invloed op de kwaliteit van het produkt, de betrouwbaarheid
van het proces en de winstgevendheid. Onvoldoende inzicht in menging kost alleen al in
de Verenigde Staten een aantal miljard dollar per jaar. Daarnaast kan vergroting van het
inzicht in de interactie tussen hydrodynamica en chemie, massa- en warmtetransport bij-
dragen aan de verbetering van de concurrentiepositie. In de laatste eeuw hebben experi-
menten significant bijgedragen aan het begrip van de hydrodynamica en van de processen,
en aan betere ontwerp- en handelingsmethoden. De belangrijke parameters op de kleine
schalen zijn echter nog steeds moeilijk te verkrijgen met behulp van experimentele tech-
nieken. Een veelbelovend alternatief dat steeds meer aandacht krijgt in de laatste jaren is
het verkrijgen van de relevante turbulente grootheden door middel van simulaties.
Dit werk maakte deel uit van een Europees project getiteld OPTIMUM waarin experi-
mentele en numerieke technieken ontwikkeld en aangewend werden om de kennis te ver-
beteren van een suspensiepolymerisatieproces en van een deeltjescoatingproces. Door dit
project uit te voeren op Europees niveau konden knowhow van de specifieke chemische
reacties en produkten, fundamentele informatie over de hydrodynamica, en expertise betr-
effende experimentele en numerieke methoden gecombineerd worden.
De belangrijkste doelstelling van dit werk is voortgekomen uit het OPTIMUM project
en was geformuleerd als een bijdrage tot betrouwbare en nauwkeurige numerieke voor-
spellingen van complexe, meerfaseprocessen. Daartoe zijn drie kwesties achtereenvol-
gens behandeld: (i) turbulente stromingsfenomenen, (ii) scalaire menging, en (iii) het sus-
penderen van vaste deeltjes in een vloeistof inclusief stofoverdracht. In dit werk is gebruik
gemaakt van een tank met keerschotten voorzien van een Rushton turbine welke opereert
onder turbulente condities.
Een gedetailleerde en nauwkeurige representatie van het stromingsveld is de eerste stap
naar een succesvolle simulatie van complexe, meerfase processen. Hiervoor werd een

xv
large-eddy simulatie (LES) gebruikt. De stromingscode, gebaseerd op een rooster Boltz-
mann discretizatieschema, was kritisch beoordeeld om vertrouwen te winnen in de ge-
bruikte methode. Twee subgrid-schaal modellen zijn vergeleken. De focus lag op de
nauwkeurigheid van de LES voorspellingen van het stromingsveld (in termen van de snel-
heid, de turbulente kinetische energie, de energie dissipatie en de turbulente anisotropie),
en het oplossen van coherente snelheidsfluctuaties veroorzaakt door grote precesserende
vortexen (een zogenaamde macro-instabiliteit). Het belangrijke resultaat van deze studie
is dat een LES de turbulente kinetische energie nauwkeurig voorspelt, in tegenstelling
tot de 50% onderschatting in een simulatie gebaseerd op de RANS (Reynolds-Averaged
Navier-Stokes) benadering. Dit resultaat is van groot belang, omdat bijvoorbeeld de meng-
patronen, de beweging van deeltjes, en de stofoverdracht in grote mate benvloed worden
door de snelheidsfluctuaties.
Om bijvoorbeeld het menggedrag van de turbulente stroming gegenereerd door een roerder,
of de stofoverdracht tussen de continue en disperse fases te beschrijven, is informatie nodig
over de scalaire concentratie in de continue fase. Hiervoor is een code ontwikkeld die
scalair transport beschrijft. Een Euleriaanse benadering is gebruikt die gebaseerd is op
de eindige volume methode. Vervolgens is er aandacht besteed aan de implementatie van
complexe wanden die in het algemeen niet gelijk lopen aan het grid. De nul-gradint re-
strictie op deze wanden is opgelegd door middel van een nieuw ontwikkelde techniek. De
code voor scalair transport is getest voor drie eenvoudige gevallen van laminaire en turbu-
lente stromingen. Daarna zijn pas de large-eddy simulaties met scalair transport uitgevoerd
voor een mengproces. De focus lag daarbij op de mengtijd als functie van roerdergrootte
en injectiepositie van de passieve tracer. Een belangrijke observatie daarbij was dat er geen
overeenkomst bestaat in de literatuur over standaard mengtijdexperimenten. Gemeten of
gesimuleerde mengtijden worden vaak onzorgvuldig vergeleken met data en correlaties in
de literatuur. Onze simulaties zijn geschikt om data en correlaties zorgvuldig te vergeli-
jken. De voorspelde mengtijden kwamen binnen 30% overeen met de experimentele re-
sultaten and waarden verkregen uit twee correlaties in the literatuur. Volgens de simulaties
schaalt de mengtijd met de ratio van de tank- en roerderdiameter tot de macht 2.5. De
mengtijd hangt niet significant af van de positie van het injectiepunt.
Tenslotte is een deeltjestransportcode gekoppeld aan de stromings- en scalair transport-
codes om een simulatie uit te voeren van een geroerde vast-vloeibare suspensie met sto-
foverdracht. De deeltjestransportcode is gebaseerd op een Lagrangiaanse benadering waarin
ronde vaste deeltjes gevolgd worden in een Euleriaanse benadering van het stromingsveld
door de dynamische vergelijkingen van beweging en rotatie van de deeltjes op te lossen.
Deeltje-deeltje en deeltje-wand botsingen zijn opgenomen in de code. De kracht uitgeoe-
fend door de vloeistof op de deeltjes is weer teruggekoppeld naar de vloeistof (two-way

xvi
koppeling). Als toepassing is gekozen voor een oplosproces in een vat op laboratorium
schaal met een volume van 10 liter. Een verzameling van 7 miljoen deeltjes is losgelaten
in het bovenste gedeelte van het vat. Op het moment van loslaten was de lokale vol-
umefractie gelijk aan 10%, wat overeenkomt met een vat-gemiddelde volumefractie van
1%. De eigenschappen van de deeltjes komen overeen met die van calciumchloride. De
simulatie betreft de dispersie van de deeltjes in het vat, de oplossnelheid in termen van
tijdsafhankelijke distributies van de deeltjes en de scalaire concentratie, de deeltjesgroot-
tedistributie in de tijd en de oplostijd. Voor het proces dat bestudeerd is, zijn verschillende
regimes gedentificeerd, en de oplostijd was hoogstens een orde van grootte hoger dan de
tijd benodigd voor het volledig dispergeren van de deeltjes door het hele vat. Aanvankelijk
groeperen de deeltjes zich in streepvormige patronen. Door het afnemen van de deelt-
jesenertie gingen de deeltjes zich meer gedragen als vloeistoftracers. Tot aan het tijdstip
waarop de deeltjes volledig gedispergeerd waren in het vat waren de ruimtelijke deeltjes en
scalaire concentratiedistributies erg inhomogeen. Vervolgens werden de distributies meer
homogeen.

xvii
1 Introduction

1.1 Mixing in the process industries


The concept of mixing is of high importance in many (chemical) applications. For fluid
systems, mixing operations are necessary to mix either miscible fluids (e.g. blending of
petroleum products) without chemical reaction or mass transfer, or immiscible fluids (e.g.
emulsification) with inter-phase mass transfer. Mixing operations are also widely applied
in multiphase contacting processes where they are necessary for bringing two (gas and
liquid; liquid and solid; gas and solid) or three phases (gas, liquid and solid) into intimate
contact to achieve inter-phase heat and mass transfer, chemical reactions, etc.
Mixing tasks are mainly carried out in stirred tanks or static mixers, where the mixing
is brought about by revolving impellers (of various shapes and sizes) or internals, respec-
tively. Also bubble columns are frequently applied in view of gas/liquid mixing. Gas-solid
mixing is often achieved in a fluidized bed, where a gas passes through a bed of solid par-
ticles at a high speed causing a certain degree of fluidization of the bed. As a result,
inter-phase mass transfer or segregation of light and heavy particles is achieved.
Mixing is a central and essential feature of many processes in the food, pharmaceuti-
cal, paper, plastics, ceramics and rubber industries. Industrial applications of mixing are
ranging from very simple cases to very complex processes such as polymerization, crys-
tallization, and precipitation. Although in such processes mixing is just part of the whole
process, it often has an important impact on product quality, process reliability, and pro-
cess economy. As a result, the financial investment in both the capital and running costs
of mixing processes, when viewed on international scale, is considerable. It has been esti-
mated that the cost to process industries due to an inadequate understanding of mixing is
a few billion dollars per year in the USA alone (Tatterson, 1994). Also, rules of thumb for
scaling up contacting processes cannot be applied with confidence. This requires expen-
sive pilot tests and/or results in either process failures or too conservative process designs.
The major challenges for process industry to improve their competitiveness are short-
2 Chapter 1. Introduction

ening of process development time, enhancement of product quality, waste reduction and
usage of existing plants up to full capacity. These challenges are triggered by the desire of
industrial companies to make products at the lowest possible cost, and by environmental
constraints imposed by national and European laws. For these processes the lack of in-
sight into the hydrodynamic phenomena involved and their coupling with chemistry and
heat and mass transfer must be overcome to achieve competitiveness improvement.
All this has resulted in the need for more detailed insights in the turbulent flow phenom-
ena and processes and for better design and operation methods. The role of experiments
in order to meet this need has shifted over the years. The characterization of the flow in
a stirred tank by means of global parameters, such as the power drawn by the impeller
(Rushton et al., 1950), the pumping capacity of the impeller, and circulation times within
the tank (Holmes et al., 1964), made way for local, mostly optical, flow measurements.
Detailed measurement techniques, such as laser Doppler anemometry (LDA; e.g. Schfer
et al., 1997), particle image velocimetry (PIV; e.g. Sharp & Adrian, 2001) and laser in-
duced fluorescence (LIF; e.g. Distelhoff et al., 1997) have great significance in resolving
the large-scale as well as small-scale flow structures. Furthermore, parameters that dom-
inate the smallest scales (e.g. energy dissipation rates (Micheletti et al., 2004), spectral
information at the micro-scale and shear rates) are becoming more accessible for experi-
mental techniques.
A promising route to assess the small scales, which is becoming feasible with the avail-
ability of large computational resources, is to resolve the relevant turbulence quantities by
simulation. Computational fluid dynamics (CFD) has proven to be a versatile tool for
studying a wide range of hydrodynamic phenomena. CFD codes based on finite volume
discretization have been introduced in the eighties (FLUENT in 1983 and CFX in 1987).
Next to commercial packages, numerous in-house codes exist that are based on existing
discretization schemes or novel schemes. One scheme that is gaining more attention in
the academic as well as in the industrial world is based on the Lattice-Boltzmann Method
(LBM), which is a solver for the Navier-Stokes equation. There is a commercial code
(Power FLOW, EXA Corporation) available, that can be used for flow problems related to
e.g. the automotive and aerospace industries. The Navier-Stokes solver used in this work
is an LBM-based academic research tool.
Today, CFD is incapable of solving all flow scales in industrial applications, within
a reasonable amount of time and memory space. A direct numerical simulation (DNS)
of the turbulent flow at industrially relevant Reynolds numbers (Re 104 ) would require
an enormous amount of grid cells and time steps to capture all relevant time and length
scales in the flow. In a large eddy simulation (LES), the smallest scales are modeled with
a subgrid-scale model, thereby reducing the amount of computational resources. While a
1.2. Two mixing applications; OPTIMUM project 3

LES is still computationally expensive, attention to this method is increasing in the last
years. Nowadays, most CFD codes solve the Reynolds-Averaged Navier-Stokes (RANS)
equations in combination with a closure model for the Reynolds stresses, which compared
to a LES or DNS is computationally less expensive. However, turbulence models in a
RANS approach rely on empirical parameters obtained in standard flow problems (e.g.
turbulent channel flow), and therefore do not perform that well in more complicated flow
problems. Furthermore, there is no clear distinction between the part of fluctuations that
is explicitly resolved and the part that is taken into account by a turbulence model.
The available experimental and numerical techniques of today are of great help to
broaden our understanding of the flow phenomena, in order to improve process scale-up,
process design, product quality, etc. The drive toward larger degrees of sustainability in
the process industries has urged for lower amounts of solvents and for higher yields and
higher selectivities in chemical reactors. These goals are often achieved via cooperation of
industrial companies and universities in European projects. The next section focuses on the
EU project OPTIMUM in which experimental and numerical tools have been employed
to increase the understanding of a suspension polymerization process and a particle coating
process.

1.2 Two mixing applications; OPTIMUM project


This research was part of a European project called OPTIMUM (OPTimization of Indus-
trial MUlti-phase Mixing), funded by the European Community under the Competitive
and Sustainable Growth Program. The project start date was December 1, 2000, and
the closure date was November 30, 2003. The total budget amounted to approximately 4
million euros, of which 10.5% was on the account of TU Delft.
In the subsections below, motivation and aim of the project are described, together with
the two chemical processes under consideration. This description is somewhat concise,
because of confidentiality imposed by the industrial partners. A significant research effort
has been made in this project, and my relevant contributions will be briefly addressed in
the next subsections.

1.2.1 Motivation, aim and project partners


Chemical manufacturing currently constitutes one of the major European industries on
which many communities depend economically. A trend to increase chemical production
capacities in non-European countries with lower costs (less strict protection of environ-
4 Chapter 1. Introduction

ment, labor costs) is witnessed. Therefore, sustained innovation of products and processes
is absolutely necessary to keep Europe a chemical manufacturing region of significance.
This is the main reason why the OPTIMUM project was initiated.
Recent progress in CFD research and development brought considerable progress in
numerical fluid dynamics and reaction engineering, to the effect that complex multi-phase
mixing and reaction processes come within reach of predictive computations. Therefore,
major parts of conventional process design methods can be replaced by predictive compu-
tations, which in turn result in optimized processes and in decreasing development time
and costs. The reliability of the computations need to be verified and validated against de-
tailed experimental data. The purpose of the project is the development of computer tools
especially for multi-phase mixing. This ambitious goal requires know-how on the specific
chemical reactions and products of the particular processes under investigation, fundamen-
tal information on hydrodynamic phenomena (e.g. turbulence modeling, micro-mixing,
break-up and coalescence models) and expertise on experimental measuring techniques,
and a software company that integrates the information into a CFD code. Therefore, this
project was addressed at European level.
Two chemical processes, a polymerization process forming a solid granular product
(suspension polymerization), and a process for making coated particles of various func-
tional properties (particle coating), have been selected for the OPTIMUM project for a
number of reasons that are summarized below.

Suitability: Both processes are highly complex, multi-phase and mixing sensitive.

Generalization: Each process constitutes a separate class of problems accessible to


CFD for multi-phase flow, giving the project the broadness needed for application
to other processes.

Socioeconomics: The processes are mature and already employed in large scale
European production facilities. Scaling-up difficulties and cost effectiveness hamper
new investment in large European production facilities. Consequently, secure and
growing employment and waste product reduction are important issues in the project.

Potential: New products developed, particularly nano-materials, can be produced


by particle coating. The computer tools developed within the project will greatly
facilitate research and contribute to the European technical progress.

The seven partners with complementary expertise from three EU member states form
the consortium:
1.2. Two mixing applications; OPTIMUM project 5

Merck KgaA, Germany; Particle coating process.

Dow Benelux NV, The Netherlands; Suspension polymerization process.

INVENT Umwelt- und Verfahrenstechnik GmbH & Co KG, Germany; Mixing tech-
nology.

AEA Technology, United Kingdom (Today merged with Ansys Inc., USA); CFX
code supplier.

TU Delft, The Netherlands; Modeling expertise, experience large eddy simulations.

LSTM Erlangen, Germany; Experimental and numerical competencies on stirred


tank flows.

Kings College London, United Kingdom; Expertise in design and execution of ex-
periments in a large range of complex engineering processes.

A detailed study on turbulence models employed in the CFD codes is an important


issue, since the simulation accuracy of multi-phase flows will heavily rely on correct
turbulence parameter prediction. The flow agitated in a stirred tank is inherently time-
dependent, due to the motion of the impeller. The existing engineering turbulence models
implicitly assume steady-state behavior, and are notorious for their underprediction of the
turbulence kinetic energy. Therefore, these RANS models (e.g. k ) need to be evalu-
ated through experimental data and with respect to numerical results based on large eddy
simulations. The large eddy simulations performed within this context are described in
chapters 2 and 3.
In the next two subsections, a brief description is given of the two processes considered
in the OPTIMUM project.

1.2.2 Suspension polymerization


The Dow company is one of the major engineering plastics manufacturers. The suspen-
sion polymerization process of styrene is employed in the production of one of the plastics
products (e.g. insulation foam, consumer electronics, rigid packaging, food service dis-
posables). Today, the major market opportunity growth is in Asia, and consequently the
competition of this market segment from Asian manufacturers is increasing rapidly. In or-
der to withstand this competition, one needs to utilize existing facilities in Europe better to
increase the profitability and capacity of existing installations. As a result, the OPTIMUM
6 Chapter 1. Introduction

project aims at a more competitive process, thereby enabling opportunities to expand pro-
duction capacity in Europe.
Suspension polymerization of styrene is a batch process in which monomer(s), rel-
atively insoluble in water, is (are) dispersed as liquid droplets with steric stabilizer or
the so-called Pickering stabilizer and under vigorous stirring (which is maintained during
polymerization) to produce polymer particles as a dispersed solid phase. The suspension
of monomer in water has volume fractions up to about 50%. The turbulence generated by
the revolving impeller induces break-up and coalescence of droplets, and consequently the
formation of a droplet size distribution. The rate-limiting break-up and coalescence pro-
cesses take place at the small scales (comparable to the droplet size), and are controlled by
the rate at which turbulent kinetic energy is dissipated. Suspension stabilizers are used in
order to prevent uncontrolled coalescence in the final stage of the process and to control
the particle size distribution.
The process takes place in a jacketed stirred vessel. At the start of the process an
initiator is added. The initiator molecules decompose into free radicals, which start the
polymerization process. During polymerization the monomer is slowly converted into
polymer. As a result, the viscosity in the disperse phase increases. The process can glob-
ally be divided into three stages. The first stage is called the low-viscosity stage with
relatively small monomer/polymer droplets. With increasing conversion in the dispersed
phase, the viscosity increases sharply (caused by the so-called Trommsdorff- or gel-effect),
the properties of the inter-facial layer and the density change slightly, causing a change in
equilibrium of breakage and coalescence toward coalescence. The particle size distribu-
tion shifts to a broader distribution with larger particle size. This stage is called the sticky
phase. If the suspension remains stable (due to suspension stabilizers), the particle size
distribution reaches its final state, which is called the identity point.
A schematic picture of the vessel geometry, accompanied by an averaged velocity field
in the baffle plane (obtained by an LES computation), and a schematic course of the Sauter
mean diameter in time (including the three stages mentioned) is shown in Figure 1.1. The
average velocity field shows that, because of the low mounting position of the impeller,
one large and one small recirculation zone is created above and below the impeller, re-
spectively. The impeller pumps in a radial direction.
The particle size distribution in the suspension was modeled based on a population
balance approach (according to Ramkrishna (1985)). Initially, the evolution (in space and
time) of the particle size distribution was based on two parameters, the degree of polymer-
ization and the particle size. The characteristic response times for heat transfer phenom-
ena in the tank have been analyzed and presented by Hartmann (2001) in a confidential
report. Main conclusion from that report was that the process time scales are an order of
1.2. Two mixing applications; OPTIMUM project 7

(b)
Sauter diameter

low viscosity
stage identity point

sticky
stage final stage

time

(a) (c)

Figure 1.1: (a) Schematic representation of the vessel geometry with styrene droplets sus-
pended in water with stabilizer. The vessel has a heated jacket to regulate the
temperature of the contents. (b) Averaged flow field (LES) in the vertical baffle
plane, the well-mixed region shown is depicted in (a) with a dashed box. (c)
Schematic time trace of the Sauter mean diameter, including the three stages.

magnitude larger than the particle response times. Consequently, effects of non-uniform
temperatures due to heating up of the process could be ignored. Since the degree of poly-
merization (i.e. conversion) is directly related to the temperature, it could be assumed
uniform throughout the reactor vessel. As a result, the population balance equations de-
scribe the evolution (in space and time) of a single parameter being the particle diameter.
The population balances are conservation equations which account for transport of
8 Chapter 1. Introduction

droplets into and out of the reactor volume, and generation and disappearance of droplets
through break-up and coalescence processes. The population balance model is in the CFD
code CFX often referred to as the MUSIG (MUltiple SIze Group) model. Coalescence
and break-up correlations are required to provide the rates at which the droplets change
size during the course of the simulation. The popular break-up model of Luo & Svendsen
(1996) and coalescence model of Prince & Blanch (1990) are valid for gas-liquid systems
(bubbly flows). Therefore, they are not not applicable for liquid-liquid dispersions such as
the suspension polymerization process.
The break-up and coalescence correlations highly depend on the rheology of the dis-
persed phase. Two models have been investigated within this project (Hartmann, 2002b).
The first model is introduced by Vivaldo-Lima et al. (1998), and assumes the rheological
properties of the dispersed phase (styrene/polystyrene mixture) to be similar to a power-
law fluid. The other model has been introduced by Alvarez et al. (1994) and Maggioris
et al. (2000), and assumes the fluid of the dispersed phase to behave as a Maxwell fluid
with visco-elastic properties. From a confidential report of Hartmann (2002b), the follow-
ing observations based on these two models are summarized.

The power-law model is poorly documented, and Vivaldo-Lima et al. (1998) use
a power-law index equal to one. As a result, they model the disperse phase as a
Newtonian fluid.

The power-law model has 10 adjustable constants.

Visco-elastic effects are discarded in the power-law model.

The Maxwell model is poorly documented by Alvarez et al. (1994) and Maggioris
et al. (2000), the papers containing numerous typographical errors in the equations.
Most of these errors and other inconsistencies were solved in the report (Hartmann,
2002b), partly through a personal communication with Kotoulas, a PhD successor
of Maggioris.

The Maxwell model needs 5 adjustable constants.

The Maxwell model takes into account visco-elastic effects in the dispersed phase.

Main conclusion from the report is that the Maxwell model was found to be most
appropriate for the suspension polymerization process. This model has been implemented
in the CFX code, and has been selected for the ultimate particle size simulations with the
MUSIG model.
1.2. Two mixing applications; OPTIMUM project 9

NaOH sprinkling

TiOCl 2 feed
mica solution
(a) (b)

Figure 1.2: (a). Schematic representation of the vessel geometry in the vertical baffle
plane. (b) Averaged flow field (LES) in a vertical plane which is perpendicu-
lar to the baffle plane. The well-mixed region shown is depicted in (a) with a
dashed box.

1.2.3 Particle coating

The particle coating process is executed by Merck, which is the only European company
in this field. Similar to the Dow company, Merck is challenged by Asian competitors.
In order to participate in the market growth, Merck is forced to meet ever more stringent
quality requirements from its main customers from the plastics, paint and cosmetics in-
dustries. In the particle coating process, natural mineral mica particles are covered with a
thin layer of metal oxides (e.g. titanium dioxide). Through an interplay of transparency,
10 Chapter 1. Introduction

(a) (b)

Figure 1.3: (a) Illustration of the mica particles. (b) Illustration of a coated mica particle.

refractive index, coating and multiple reflections, a variety of color effects is obtained.
The overall coating process can be described as a titration method. An aqueous, strong
acid T iOCl2 solution is continuously dosed to a mica suspension. The mass fraction of the
mica particles is around 6%. The mica partices have the shape of plates (i.e. particles with
a high width/height over thickness aspect ratio. The precipitation of the titanium oxide-
hydroxide occurs at a certain pH value. The pH value is kept constant by continuous
addition of N aOH according to the following reaction scheme:

T iOCl2 + 2N aOH + mica T iO(OH)2,mica + 2N aCl (1.1)

A schematic representation of the mixing vessel geometry accompanied by an averaged


flow field obtained by a LES computation is shown in Figure 1.2. The impeller pumps
axially, thereby creating a large and a small recirculation zone.
The species react most likely in a homogeneous reaction, and form nano particles.
Subsequently, these nano particles agglomerate on the mica particles to form the coating.
This process, and consequently the product quality, is strongly dependent on parameters
such as the location of the feed point, and the pH. Also heterogeneous reactions (i.e. re-
actions at the mica surface) take place. The situation of heterogeneous reactions is akin
to growth in crystallization, which is mainly determined by the rate at which supersatura-
tion can be transported to the crystals. The turbulence generated by the revolving impeller
plays an important role in the formation of the crystals. Similar to break-up and coales-
cence phenomena mentioned with a view to the suspension polymerization process, the
transportation rate of supersaturation to the crystals is rate-limiting and controlled by the
energy dissipation rate. Figure 1.3 illustrates the mica particles at the start of the process
and the coated layer on a mica particle at the end of the process.
In the final computations of the particle coating process, the particles are tracked in the
1.3. Mixing and LBM at the Kramers Laboratory 11

turbulent flow field. Since the mica particles have a high width/height of thickness aspect
ratio, standard correlations for the drag force are not valid. A theoretical investigation
has been carried out by Hartmann (2002a) to provide an estimation of the drag force
experienced by the mica particles. The main conclusions will be briefly described.
The non-spherical mica particles immersed in the fluid experience a drag force that
acts against their direction of motion. The orientation of the particles is highly dependent
on the particle Reynolds number and shape, and so is the drag force. Based on the report
of Hartmann (2002a), the mica particles do not have preferred orientations throughout the
tank (i.e. in both the turbulent and quiescent regions). Therefore, a drag force correlation
has been proposed, based on a mean resistance of translation. The drag coefficient pro-
posed depends on the aspect ratio being the thickness to height/width (the shape) and is
inversely proportional to the particle Reynolds number.

1.3 Mixing and LBM at the Kramers Laboratory


Subsequent to the processes described within the OPTIMUM project, this section deals
with earlier successful research in the Kramers Laboratory on (multi-phase) complex pro-
cess applications with the help of the previously mentioned lattice-Boltzmann Navier-
Stokes solver (see section 1.1). Since the introduction in 1996 of the lattice-Boltzmann
method (LBM) in our Laboratory, a significant research effort has been undertaken to
study the flow phenomena in a large range of real-life applications.
One of the lattice-Boltzmann discretization schemes used in the Kramers Laboratory
(in particular in the work presented in this thesis) originates from the scheme presented by
Somers (1993). The first LES study in a stirred tank geometry with LBM was reported by
Derksen & Van den Akker (1999). The results revealed a range of flow field characteristics
(e.g. the so-called trailing vortex structure, kinetic energy, dissipation data). A newly
developed immersed boundary technique, that imposes the no-slip velocity constraint at
the off-grid walls, is successfully applied.
Various subgrid-scale models have been explored. Derksen (2001) assessed the per-
formance of the structure function model (Metais & Lesieur, 1992) and the widely used
Smagorinsky (1963) model in the prediction of a turbulent flow agitated by a pitched blade
impeller. Furthermore, the Voke (1996) model was assessed by Hartmann et al. (2004b),
which is a subject of chapter 2.
Turbulently agitated solid-liquid suspensions are encountered in various industrial pro-
cesses (e.g. catalytic slurry reactors, crystallization). Derksen (2003) reports on solid-
liquid simulations of the turbulent flow in a stirred tank geometry according to an Eu-
12 Chapter 1. Introduction

lerian/Lagrangian approach. Particle-particle collisions proved to play a critical role in


order to have the particles in suspension in accordance with the just-suspended criterion
(Zwietering, 1958).
Another industrial application is a crystallization process. Agglomeration is most often
an unwanted phenomenon, as it tends to broaden the crystal size distribution. Hollander
et al. (2001) have investigated the agglomeration behavior in a stirred tank and observed a
huge discrepancy between the agglomeration kernel obtained by Mumtaz et al. (1997) in
simple shear flow and stirred tank flow. This discrepancy was attributed to hydrodynamic
effects.
A similar industrial application was studied by Ten Cate et al. (2001). They studied
crystal-crystal collisions in a draft tube baffled industrial crystallizer by means of LES.
The information on the local turbulence in the crystallizer was then used to determine
local collision frequencies and intensities through a detailed DNS of an isotropic turbulent
field in a box with thousands of particles.
Gas-solid cyclone separators are simple devices (no moving parts) to separate solid
particles from gas streams. These devices are used in various applications (e.g. in the
oil industry). Derksen & Van den Akker (2000) studied the turbulent flow in a reverse-
flow cyclone by means of LES. They recovered the average flow field with a high level
of accuracy. Vortex core precession was observed in the simulations, its frequency being
close to the experimentally found value.
The lattice-Boltzmann discretization scheme, known as the LBGK (Lattice Bhatnagar-
Gross-Krook) method (Qian et al., 1992), is definitely the most well-known and widely
used approach for LBM. Kandhai et al. (2003) used this scheme for investigating the
average drag that acts on particles in a fluidized bed reactor under steady-state conditions.
The applicability of existing drag models has been checked by simulations through varying
Reynolds number, number of spherical particles and particle volume fraction.
Van Wageningen et al. (2004) used LBGK to investigate the flow in a Kenics static
mixer for a range of Reynolds numbers. They assessed their results with results based on
simulations with FLUENT. An important observation was that, for Re> 300, the LBM
simulations were faster and computationally less demanding than the FLUENT simula-
tion.

1.4 Motivation and aim


When we look back to the processes mentioned in the previous sections, it is clear that a
detailed knowledge of the flow phenomena involved is only a first step in gaining more
1.4. Motivation and aim 13

understanding of complex multi-phase processes. The previous section illustrated the


amount of research already undertaken in our laboratory to improve the knowledge of
(multi-phase) process applications. In this section we again consider particle coating (see
section 1.2.3), in order to make clear the issues this thesis aims to focus on. The ulti-
mate goal is to contribute to reliable and accurate predictions of processes such as particle
coating.
Inter-phase mass transfer from the continuous phase to the mica particles (solid phase)
plays an important role in the heterogeneous reactions. If this process is to be solved
via simulation, a scalar concentration distribution in the continuous phase is required for
determining the mass transfer rate. The transport of a scalar is governed by the convection-
diffusion equation, and solution of this equation can either be obtained in an Eulerian
(e.g. finite volume, finite difference methods) or a Lagrangian framework (e.g. Monte
Carlo modeling). The work of Van Vliet (2003) has shown that the Lagrangian approach
is feasible but expensive. Therefore, an Eulerian approach based on the finite volume
formulation is attempted in this work for scalar mixing calculations. Furthermore, in the
absence of second-order (or higher-order) chemical reactions in the liquid phase we do not
need to take concentration fluctuations into account. A simple blending process is used as
a reference case in chapter 5.
Scalar mixing calculations, eventually in combination with particle transport, are not
possible without a detailed knowledge of the flow phenomena occurring in the continuous
phase. For instance, particles will not feel the average flow but the turbulent eddies with
size comparable to the particle size. Furthermore, mass transport is not likely to be con-
trolled by the large-scale flow. As a result, the growth rate of the layer thickness on the
mica particles is expected to be in-homogeneously distributed throughout the tank, which
induces the formation of a particle (mica) size distribution in the tank. These arguments
motivate a detailed resolution of the flow field (i.e. the first step mentioned in the first
paragraph). In this thesis, this is accomplished by performing large eddy simulations.
Large eddy simulations need to be critically assessed against experimental and numeri-
cal data based on the RANS approach. This aspect is the subject of chapter 2. Furthermore,
in a LES fluctuations are resolved down to the size of the numerical grid. One may wonder
if all these fluctuations are of a purely turbulent nature. Experimental data have confirmed
that part of the fluctuations are induced by so-called macro-instabilities that severely affect
the turbulence levels (Nikiforaki et al., 2002). Large eddy simulations have been executed
to determine whether it is possible to determine these instabilities in a numerical simula-
tion. Chapter 3 deals with the recovery of the macro-instability by means of large eddy
simulations.
The implementation of complex boundaries (e.g. cylindrical tanks, curved impeller
14 Chapter 1. Introduction

blades) is a difficult task. Nowadays, commercial codes make use of unstructured meshes
and body fitted grids. However, in structured codes (such as LBM or CFX-4) the complex
boundaries are in general off-grid. Within LBM on a Cartesian mesh, Derksen & Van den
Akker (1999) have successfully developed an immersed boundary technique to impose the
no-slip velocity constraint at the walls that are off-grid.
With a view to the finite volume scalar mixing solver developed in chapter 4 and used
in chapters 5 and 6, a novel immersed boundary technique needs to be developed. The
stair-case approximation is often applied with a fine grid resolution, that is an inaccurate
representation of the wall and which is computationally expensive. Several other tech-
niques have been developed and tested in the literature for a wide range of applications.
In this thesis, a novel technique is developed (see chapter 4) in order to describe a zero-
gradient boundary condition for the scalar concentration at off-grid boundaries.
In summary, the focus in this thesis is on

Assessment of the lattice-Boltzmann Navier-Stokes solver (LES)

Development of a scalar mixing solver in the finite volume framework in combina-


tion with an immersed boundary technique

Assessment of coupled flow (LES) and scalar mixing solvers through simulation of
a blending process

Study on mass transfer and solids suspension through simulation of a dissolution


process with coupled flow (LES), scalar mixing and particle transport solvers.

1.5 Stirred tank geometry


Fluid dynamics research on stirred tanks has adopted a few de facto standard impeller
and tank geometries, for which quite a few experimental as well as numerical data are
available. The tank geometry is mostly a flat-bottomed cylindrical tank, with a diameter
equal to the liquid height. Four baffles positioned along the perimeter of the tank prevent
a solid body rotation of the fluid, and consequently enhance the mixing performance of
the tank. In Figure 1.4, two tanks equipped with a standard impeller are shown. The
first impeller consists of four, 45o -pitched blades mounted on a shaft (Figure 1.4a). The
pitched blade impeller pumps axially downward or upward, depending on the direction of
rotation. The second impeller consists of six flat, vertically oriented blades mounted on a
disk (the so-called Rushton turbine, see Figure 1.4b). This impeller pumps the fluid in the
1.6. Outline of this thesis 15

(a) (b)

Figure 1.4: Standard stirred tank configurations. (a) Cylindrical tank equipped with four
baffles and a pitched blade turbine. The four blades are under a 45o angle. (b)
Cylindrical tank equipped with four baffles and a disk (Rushton) turbine with
six, flat blades.

radial direction. The Rushton turbine has been used for all the simulations presented in
this thesis.

1.6 Outline of this thesis


In view of the motivation of this work described in section 1.4, the thesis is structured in
the following order.

Turbulent flow phenomena

The first two chapters of this thesis deal with single-phase flow simulations in a Rush-
ton turbine stirred tank geometry. These large eddy simulations are performed with the
lattice-Boltzmann Navier-Stokes solver based on the scheme published by Somers (1993).
Both chapters have been published in scientific journals.
16 Chapter 1. Introduction

In chapter 2, a large eddy simulation is assessed against a transient RANS simulation


based on the commercial software package CFX-5, and LDA experiments performed by
Schfer et al. (1997). As the Reynolds number was relatively low (Re=7300), the flow
is fully resolved in some parts of the tank. Therefore, next to the standard Smagorinsky
model, an alternative subgrid-scale model introduced by Voke (1996) has been tested. This
latter model blends smoothly between a DNS in the bulk flow and a LES in the impeller
region and discharge flow.
Chapter 3 deals with the so-called macro-instability that is present in stirred tank
flows. A large-scale vortical structure precessing about the tank centerline has been iden-
tified experimentally in the literature. The frequency of this precession lies approximately
at 0.02 times the impeller speed. Large eddy simulations have been performed with a fo-
cus on the flow structures, and the spectral characteristics of the velocity components. The
flow has been simulated at various Reynolds numbers and grid resolutions.

Scalar mixing

Chapter 4 treats the development of a finite volume code in order to solve the mixing
of a passive scalar in a (turbulent) flow field. The code is based on an explicit formu-
lation of the convection-diffusion equation. Numerical diffusion is minimized through a
second-order TVD scheme for the discretization of the convection term. A first test is per-
formed on laminar flow inside a cavity. In order to model complex (off-grid) boundaries,
an immersed boundary technique is developed. This technique is tested in an inclined
cavity geometry (laminar flow) and a cylindrical tank equipped with a side-entry mixer
(turbulent flow).
Subsequently, detailed scalar mixing simulations are performed in a Rushton turbine
stirred tank geometry. Chapter 5 describes the simulation procedure and the results in
terms of snapshots of the concentration field, simulated time series of the scalar concentra-
tion at various monitoring points, and the mixing time. The developed boundary technique
is critically assessed (mass conservation).

Solid-liquid mixing including mass transport

A particle transport solver has been developed by Derksen (2003). The code solves the
transport equation for the particle motion, and takes into account particle-wall and particle-
particle collisions. Simulations in a Rushton turbine stirred tank geometry have shown that
the exclusion effect (i.e. the particles are not allowed to physically overlap) brought about
by the particle-particle collisions plays a crucial role to keep the particles suspended. A
combination of the lattice-Boltzmann Navier-Stokes solver, the scalar mixing solver, and
1.6. Outline of this thesis 17

the particle transport solver allows for detailed simulations of solid-liquid mixing with
mass exchange between the liquid and solid phases (as occurring in e.g. crystallization or
dissolution processes).
Chapter 6 describes a simulation of a dissolution process. Suspended particles in
an unsaturated liquid dissolve under the action of a concentration gradient. Results are
presented through snapshots of the particle distribution, scalar concentration field, and the
dissolution time (i.e. the time needed for all particles to fully dissolve in the continuous
liquid).

Conclusions and outlook

Finally, chapter 7 reflects on the work presented in the previous chapters. Main con-
clusions are summarized, and an outlook for future work is presented.

Note from the author

Chapters 2 and 3 have been published in scientific journals, and chapters 5 and 6 have
been submitted as journal papers. Therefore, some parts of the text in this thesis may
appear more than once.
2 Assessment of large eddy
simulations on the flow in a stirred
tank

Large eddy simulations and RANS calculations were performed on the flow
in a baffled stirred tank, driven by a Rushton turbine at Re = 7300. The LES
methodology provides detailed flow information as velocity fluctuations are
resolved down to the scale of the numerical grid. The Smagorinsky and Voke
subgrid-scale models used in the LES were embedded in a numerical lattice-
Boltzmann scheme for discretizing the Navier-Stokes equations, and an adap-
tive force-field technique was used for modeling the geometry. The uniform,
cubic computational grid had a size of 2403 grid nodes. The RANS calcu-
lations were performed using the computational fluid dynamics code CFX
5.5.1. A transient sliding mesh procedure was applied in combination with
the Shear-Stress-Transport (SST) turbulence closure model. The mesh used
for the RANS calculation consisted of 241464 nodes and 228096 elements
(hexahedrons). Phase-averaged and phase-resolved flow field data, as well as
turbulence characteristics, based on the LES and RANS results, are compared
both mutually and with a single set of experimental data.
Key words: stirred tank, mixing, turbulence, fluid dynamics, simulations,
LDA
This chapter has been published in Chem. Eng. Sci. 59(12), 2419-2432 (2004).

2.1 Introduction
The flow structures in a turbulently stirred tank are highly three-dimensional and complex
with vortical structures and high turbulence levels in the vicinity of the impeller. Under the
20 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

action of a revolving impeller, the fluid is circulated through the tank. Mixing is enhanced
by baffles along the tank wall preventing a solid-body rotation of the fluid.
As stirred tanks have always played an important role in industry, the understanding
of the flow phenomena encountered in such tanks have been subject of research studies.
These often document power number - Reynolds number relationships, and circulation
times, characterizing the global flow field. Also integral investigations for selected stir-
rer/vessel combinations were used to derive scaling rules. In modern chemical engineer-
ing, however, there is an increasing demand for local flow information. For instance, yield
prediction of chemically reacting fluids for a specific reactor is difficult, since most indus-
trial flows are highly turbulent, inhomogeneous, time dependent, and cover a wide range
of spatial and temporal scales (Van Vliet et al., 2001). In other applications, mixing at
the small scales is important since rate-limiting phenomena take place at these scales. For
example, droplet breakup is controlled by the dissipation rate of turbulent kinetic energy
at the scale of the drop size, and the sizes of the turbulent eddies at this scale (Tsouris &
Tavlarides, 1994), (Luo & Svendsen, 1996).
Many investigations to date have been concerned with the flow in tanks stirred by a
Rushton impeller. This type of impeller is much-studied, and a significant amount of
experimental and numerical (RANS) data are available in the literature that can be applied
for verification. The Rushton impeller induces a strong radial discharge stream. In the
wakes behind the impeller blades, three dimensional vortices are formed. The variation
of the radial velocity in the impeller discharge flow due to the passage of the impeller
blades was found to cause an increase of the turbulence levels (Van t Riet & Smith, 1975).
This non-random part of the turbulent fluctuations is sometimes presented as pseudo-
turbulence.
Detailed flow measurement techniques, such as Laser Doppler Anemometry (LDA)
and Particle Image Velocimetry (PIV) (Sharp & Adrian, 2001) are suitable for the inves-
tigation of highly vortical flows, and are able to resolve large-scale as well as small-scale
flow structures. LDA investigations have identified the formation of the trailing vortex
pair behind each impeller blade and provided information about the flow periodicity in
the impeller vicinity and its associated increase of the turbulence levels (e.g. Yianneskis
et al., 1987; Schfer et al., 1998). Laser Doppler Anemometry has proven to be more ac-
curate in the measurement of flow fields in stirred tank reactors than other techniques (e.g.
Pitot-tubes, hot-wire anemometers), since it provides flow information even in unsteady
and highly turbulent flow regions as well as in the return flow areas of the tank and it oper-
ates without fluid contact. Furthermore, parameters that dominate the smallest scales (e.g.
energy dissipation rates (Micheletti et al., 2004), spectral information at the micro-scale
and shear rates) are becoming more accessible for experimental techniques.
2.1. Introduction 21

Since computational resources increase year-by-year, detailed computational informa-


tion about the flows at the scales relevant to mixing, chemical reactions and physical pro-
cesses (e.g. bubble break-up, particle collisions) can be obtained at a fraction of the cost
of the corresponding experiment. Consequently, computational modeling of the flow has
become an alternative route of describing flows in stirred-tanks. One of the options is a Di-
rect Numerical Simulation (DNS), which resolves all turbulent length and time scales by
directly integrating the Navier-Stokes equations using a very fine grid. A direct simulation
of stirred tank flow at industrially relevant Reynolds numbers (Re 104 ) is not feasible, as
the resolution of all length and time scales in the flow would require enormous amounts of
grid cells and time steps. Even at the lower limit of the turbulent regime (Re= 7300) con-
sidered here, a true DNS would require very fine grids in the critical flow regions (such as
the impeller swept volume) in order to fully resolve the flow. Bartels et al. (2001) have at-
tempted a DNS for the same model system operating at Re=7300 using a clicking/sliding
mesh technique for the coupling of the rotating and fixed frame of reference. They as-
sumed two-fold periodicity of the solution domain, to keep the computational cost at an
acceptable level, and the finite thickness of the blades and disk was not taken into account.
We have used the technique of Large Eddy Simulation (LES). In LES, the Navier-
Stokes equations are low-pass filtered. This way, the range of the resolved scales is reduced
by the numerical grid, and the effect of the unresolved subgrid scales on the resolved
large scales is taken into account with a subgrid-scale model. It has been proven to be a
powerful tool to study stirred-tank flows (Eggels, 1996), as it accounts for its unsteady and
periodic behavior and it can effectively be employed to explicitly resolve the phenomena
directly related to the unsteady boundaries (the impeller blades). The LES methodology
embedded in the lattice-Boltzmann solver has been successfully applied to single phase
flow phenomena (macro-instabilities in a stirred tank by Hartmann et al. (2004a)) as well
as multi phase systems (solids suspension in a stirred tank by Derksen (2003)). With LES,
the computational effort is high but smaller than that of a DNS.
Most of the Computational Fluid Dynamics (CFD) simulations solve the Reynolds-
Averaged Navier-Stokes (RANS) equations in combination with a closure model for the
Reynolds stresses. Flow periodicity and unsteadiness are taken into account with the slid-
ing mesh procedure, in which the geometry of the individual blades is modeled using a grid
rotating with the impeller, whereas the bulk flow is calculated with a stationary frame. In
contrast to LES, in the RANS approach it is difficult to distinguish the part of fluctuations,
that is explicitly resolved and the part represented by the Reynolds stresses. In contrast to
the DNS work of Bartels et al. (2001), the full solution domain has been simulated in the
present LES and RANS simulations, and the blades and disk were of finite thickness.
Nowadays, the RANS approach is commonly applied for the calculation of the com-
22 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

plex flows in large-scale industrial reactors. As computational resources are expected to


increase further in future, detailed local flow information of industrial, chemical reactors
becomes within reach. As a result, the LES methodology is believed to become an engi-
neering tool. Therefore, extensive comparisons with experimental data and RANS predic-
tions are necessary, which was the purpose of the work presented in this paper. The flow
geometry considered in this research was a Rushton turbine stirred vessel. In the LES,
the standard Smagorinsky subgrid-scale model and the Voke subgrid-scale model were
adopted for turbulence modeling. For the RANS calculation, the Shear-Stress-Transport
(SST) model was used. First, the behavior of the LES subgrid scale models is checked.
Subsequently, comparisons of the velocity and turbulent kinetic energy fields are made;
phase-averaged and phase-resolved (near impeller). Attention is also paid to the computed
energy dissipation field and the (an)isotropy of the flow field. These latter predictions are
hardly accessible experimentally, but they are of interest for many mixing applications.

2.2 Stirred vessel configuration


The stirred vessel reactor used in this research was a standard configuration cylindrical
vessel of diameter T = 150 mm, with four equi-spaced baffles of width 0.1T with a small
clearance of 0.017T and a liquid column height of H = T , as shown in Figure 2.1. A lid
was positioned at height H, to prevent entrainment of air bubbles.
The working fluid was silicon oil of density 1039 kg/m3 and dynamic viscosity 15.9
mPas. The impeller rotational speed was 2672 rpm, resulting in a flow Reynolds number
(Re=N D 2 /, is the fluid density, is the fluid dynamic viscosity and N is the impeller
rotational speed in rev/s) of 7300 and a tip speed of 7 m/s.

2.3 Experimental method


As a standard LDA system only supplies flow information at a single point, data acqui-
sition was automated to enable a rapid determination of the entire flow field. A multi-
functional stirrer test rig has been developed at the Institute of Fluid Mechanics (LSTM),
and the entire measuring circuit was refractive index matched. The rotational speed of
the impeller was measured by means of an optical shaft encoder, which was coupled to
the impeller shaft. The detailed experimental results reported in this paper are taken from
(Schfer, 2001). More detailed description of the experimental setup can be found in the
paper of Schfer et al. (1997), and the experimental procedure is described by Schfer
et al. (1998).
2.4. Computational method 23

T/10 0.017 T

D = T/3
0.75D

0.04D 0.2D
H=T
0.04D
0.16D
T/3 0.32D
0.25D

Figure 2.1: Cross-section of the tank (left). Plan view and cross-section of the impeller
(right). At the top level there is a lid. The impeller is a Rushton turbine.

The vertical measurement plane was located mid-way between two baffles. The com-
plete phase-averaged flow field was measured with a high resolution of 4 mm in axial and
radial directions. Phase-resolved measurements in a range of 0o 60o behind the blade
were made in the vicinity of the impeller stream with a resolution in axial and radial direc-
tions of 1 mm and 2 mm inside and outside the impeller swept volume, respectively. The
measurements were performed for the axial, radial and tangential mean velocity compo-
nents and their corresponding RMS levels.

2.4 Computational method


2.4.1 RANS simulation
The three dimensional simulations were performed using the computational fluid dynam-
ics code CFX version 5.5.1, a finite-volume based computational fluid dynamics analysis
program which solves the Navier-Stokes equations. Turbulence was modeled using the
SST (Shear-Stress-Transport) model (Menter, 1994), which is a combination of the k-
model near the wall and the k- model away from the wall. In this way, both models are
used in areas where they perform best. The reason for making use of this model was that
the mesh was well resolved, resulting in y + values that tended to be lower than 11. By
using the SST model, CFX-5 uses an automatic wall function, which blends smoothly be-
tween a low Reynolds formulation and a standard log wall function. This gives a more
accurate representation of the friction at the wall in the areas where y + is smaller than 11.
For the mixing vessel simulations performed, the computational mesh is made up of
24 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

(a) (b)

Figure 2.2: Vertical (a) and horizontal (b) view of the mesh in the RANS calculations.
The interface between the rotating and sationary meshes was located at r/T =
0.25, and it extents from the bottom of the tank to the free surface.

two parts: an inner rotating cylindrical volume enclosing the impeller, and an outer, sta-
tionary, annular volume containing the rest of the vessel. The sliding mesh procedure was
applied, which is in fact a transient method. At each time step the inner grid is rotated
with a small incremental angle, and the flow field is recalculated taking into account the
additional velocity due to the motion of the grid. The location of the interface between the
two volumes was such that the region of flow periodicity was contained within the sliding
mesh (i.e. at r/T = 0.25 and extending all the way from the bottom of the tank to the free
surface). The mesh used for the RANS simulation consists of 241464 nodes and 228096
elements (hexahedrons). An illustration of the computational mesh is given in Figure 2.2.
In the sliding mesh procedure, the transient rotor-stator frame change model was used,
which predicts the true transient interaction of the flow between a stator and rotor passage.
This model forms part of the General Grid Interfaces (GGI), which refers to a class of grid
connection technology (CFX-5, 2002). As a consequence of the sliding mesh technique,
the computer resources are relatively large, in terms of simulation time, disk space and
quantitative post processing of the data.
Upon simulation, a blended advection scheme for the momentum and continuity equa-
tions was used to minimize effects of false diffusion. This scheme is based on upwinding
along a streamline with the addition of Numerical Advection Correction (NAC). With a
2.4. Computational method 25

10

n e,V/ n (-)

b=0
b = 2/9
0.1
0.1 1 10
n e / n (-)

Figure 2.3: The Voke subgrid scale model, Eq.(2.4). While at large e / numbers the Voke
eddy viscosity (solid line) approaches the Smagorinsky eddy viscosity (dashed
line), at small numbers it is significantly reduced.

blending factor of 1, the scheme is fully second-order but not bounded. Therefore, the
high resolution scheme has been used for the turbulence model equations. This scheme is
also a blend between a first-order and second-order scheme, but here the blending factor
is calculated based on the solution. The high resolution scheme is always bounded, and
therefore does not cause overshoots or undershoots in the solution. Hence, this scheme is
suitable for variables that are always positive (like k, and ) or bounded (e.g. volume
fractions).
First a run was made with a relatively large time step of 1.25 ms (corresponding to
20 degrees of impeller rotation), in order to allow the solution to reach a quasi steady
state. The run has then been carried on with a smaller time step of 250 s (i.e. 4 degrees of
impeller rotation). Per time step 10 iterations were performed. Typically within every time
step, the RMS residuals at the end of the time step are three orders of magnitude smaller
than at the beginning of the time step, which indicates a reasonable level of convergence.
The simulation was run on Xeon dual processors (Pentium IV) machines (Dell) with 2 Gb
of memory, 2 GHz clock frequency and a Linux operating system. The final run was done
on 6 processors in parallel.

2.4.2 Large eddy simulation


In a Large Eddy Simulation (LES), the small scales in the flow are assumed to be universal
and isotropic (i.e. independent of the specific flow geometry). The effect the small scales
26 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

have on the larger scales is modeled with a subgrid-scale model. In this research, the
well-known Smagorinsky model (Smagorinsky, 1963) was used. This is an eddy-viscosity
model with a subgrid-scale eddy viscosity, e , which is related to the local, resolved de-
formation rate:
e = 2mix S 2 (2.1)
where mix is the mixing length of the subgrid-scale motion and S 2 is the resolved defor-
mation rate, defined as:  2
2 1 ui uj
S = + (2.2)
2 xj xi
with ui the resolved ith velocity component. Note the summation convention over the
repeated indices i, j. In the standard Smagorinsky model, the ratio between the mixing
length and the lattice spacing, , is constant, called the Smagorinsky constant (c s ). How-
ever, near the walls, the subgrid-scale stresses should vanish, which is not automatically
guaranteed in the standard Smagorinsky model. In general, this is accomplished by a re-
duction of the length scale mix toward the wall. In this research, the Van Driests wall
damping function (Van Driest, 1956) is used at the tank walls, which determines the mix-
ing length as:  
y+
mix = cs 1 e A+ (2.3)

where y + denotes the distance from the wall in viscous wall units (y + = yu /, u is
the so-called friction velocity) and A+ is a constant taken equal to 26. A value of 0.10
was adopted for cs , which is a typical value used in LES computations of shear-driven
turbulent flows (Piomelli et al., 1988).

At relatively low Reynolds numbers (Re = 2000 10000) and a fine grid, as is the
case in this research, the flow in quiescent regions in the tank is completely resolved. Still,
the Smagorinsky model predicts a nonzero eddy viscosity in these regions (Eq. (2.1)).
According to Voke (1996), the Smagorinsky model is valid for high e / numbers, that
is for high Reynolds numbers. Based on a form of the dissipation spectrum proposed by
Pao (1965), a modified Smagorinsky model was introduced by Voke (1996), which is also
applicable for relatively low e / numbers:
 e

e,V = e 1 e (2.4)

with = 2/9, e is the Smagorinsky eddy viscosity and e,V is the Voke eddy viscos-
ity. Figure 2.3 illustrates the deviation of the Voke eddy viscosity with respect to the
2.4. Computational method 27

Smagorinsky eddy viscosity. For large e / numbers the Voke eddy viscosity approaches
the Smagorinsky eddy viscosity and for small e / numbers (i.e. more or less resolved
flow) the Voke eddy viscosity is greatly reduced. E.g. for e / = 10 and e / = 1 the
eddy viscosity is reduced to 2.2% and 22%, respectively.
With a view to the flow system studied at a Reynolds number at the lower limit of the
turbulent regime, we compared the performance of the Voke subgrid scale model (because
of its adequacy at low Reynolds numbers) with that of the commonly used Smagorinsky
model. Hence, more advanced models like the mixed-scale models and dynamic models
(see e.g. Lesieur & Metais, 1996) have not been attempted in this work. Derksen (2001)
has assessed the structure function model with the same numerical scheme as has been
used here, but this model did not lead to significant changes in the results.
For the large eddy simulations, a uniform, cubic computational grid of 2403 lattice
cells was defined. The diameter of the tank corresponds to 240 lattice spacings, resulting
in a spatial resolution of T /240 = 0.625 mm. As a result, the mean Kolmogorov length
scale (based on the specific power input, and a power number equal to 5) equals 90 m
or 0.145. A Lattice-Boltzmann numerical solver has been used for the finite difference
solution of the filtered momentum equations (Derksen & Van den Akker, 1999). This
scheme is based on a microscopic model of many particles that can shuffle and collide
on the numerical grid according to completely local collision rules. It can be shown that
in the incompressible limit the macroscopic behavior of the particles obeys the Navier-
Stokes equations (Chen & Doolen, 1998). In the current work, N = 1/2900, that is, the
impeller makes a full revolution in 2900 time steps (one time step equals (2900N ) 1 =
7.74 s). The diameter of the impeller equals 80 lattice spacings. As a result, the tip
speed, vtip , of the impeller was 0.09 (in lattice units, i.e., lattice spacings per time step),
which is sufficiently low for meeting the incompressibility limit in the Lattice-Boltzmann
discretization scheme.
At the bottom of the tank a no-slip boundary condition was imposed, whereas at the top
a free-slip boundary condition was set to mimic the free surface. Inside the computational
domain, the cylindrical tank wall, the baffles, the impeller, and the impeller shaft were
defined by sets of points. A forcing algorithm (Derksen & Van den Akker, 1999) takes
care of the no-slip boundary conditions at these points. The algorithm calculates forces
acting on the flow in such a way that the flow field has prescribed velocities at points
within the domain. As a geometry point does not need to coincide with the lattice nodes,
second-order interpolation of the flow velocities at the surrounding lattice sites is used to
obtain the flow velocity at that point. The spacing between the geometry points has to be
set such (i.e. smaller than a lattice spacing), that the fluid cannot penetrate at the surface
of e.g. an impeller blade. This technique is very flexible, as there is no need for building a
28 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

(a) LES: Smagorinsky model (b) LES: Voke model

Figure 2.4: Phase-averaged plot of the eddy viscosity.

new computational mesh when adapting the geometry.


On every grid node 21 (18 directions for the LB particles and 3 force components)
single-precision, real values need to be stored. The memory requirements of the simulation
are proportional to the grid size, resulting in an executable that occupies 240 3 21 4
1.16 GByte of memory. The simulations were performed on an in house PC cluster with
Intel 500 MHz processors using a MPI message passing tool for communication. The
simulations were run on 4 or 6 processors in parallel. Since the entire tank was simulated,
statistical information of the flow was stored in four vertical planes mid-way between two
baffles. Hence, four statistically independent realizations of the flow contributed to the
averaging procedure. The results were averaged over 15 impeller revolutions.

2.5 Results
2.5.1 Behavior of the subgrid scale models
Because the Reynolds number investigated in the present work is at the lower limit of the
turbulent regime (Re=7300), the performance of the two subgrid scale models applied in
2.5. Results 29

the large eddy simulations should be checked. As a fine grid resolution was used, the
results of the large eddy simulations performed might nearly resemble the results of a
Direct Numerical Simulation (DNS).
Figure 2.4 shows phase-averaged plots of the eddy viscosity distribution in a vertical
plane mid-way between two baffles for both subgrid scale models applied. The results of
the eddy viscosity distribution for both subgrid scale models clearly show that the eddy
viscosity in the bulk flow is less than 10% of the kinematic viscosity. From that it can
be concluded that the Kolmogorov scale in the bulk flow becomes of the order of the
grid spacing, and as a consequence the results would resemble those of a direct numerical
simulation. However, in the impeller region and the discharge flow the eddy viscosity
becomes of the order of the kinematic viscosity. Moreover, instantaneous realizations
showed eddy viscosities of 2-4 times the kinematic viscosity. As a result, the present
grid resolution would have been too coarse to capture all the turbulent scales. A direct
numerical simulation with the present grid resolution may therefore not accurately capture
the flow phenomena occurring in the impeller region and the discharge flow where most
of the mixing takes place. The authors note that for further evaluations of the subgrid
scale models, the present results should also be compared to those of a direct numerical
simulation with the same grid resolution. Further evaluations of the Smagorinsky model
at a higher Reynolds number (i.e. Re=29000) with the same numerical scheme as is used
in the present work are presented in the paper by Derksen & Van den Akker (1999).
A direct comparison between Figure 2.4a and Figure 2.4b shows a reduction of the
Voke eddy viscosity with respect to the Smagorinsky eddy viscosity. The relative reduction
of the eddy viscosity in the bulk flow is more pronounced than in the impeller outflow
region, as expected (see Figure 2.3).

2.5.2 Phase-averaged flow field and turbulence levels


In comparing the LES results with experimental and RANS based data, the focus is first
on the global, phase-averaged flow field. Figure 2.5 shows the experimental, RANS and
LES based results of the global, phase-averaged flow field and its associated turbulent
kinetic energy levels, mid-way between two baffles. The dominant flow feature are the
large two circulation loops; one below the impeller with downward flow near the tank wall
and upward flow near the axis and one above the impeller with upward flow near the tank
wall and downward flow near the axis. The upper circulation loop does not extend to the
top of the tank, and consequently the fluid volume in the upper part of the tank (about
14% 18% of the total tank volume) is badly mixed.
The LES for both subgrid-scale models (see Figures 2.5c,d), as well as the RANS
30 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

(a) Experiment, phase-averaged turbu- (b) RANS, not all vectors are plotted
lent kinetic energy was only measured for clarity
in the impeller region

(c) Smagorinsky model, not all vectors (d) Voke model, not all vectors are
are plotted for clarity plotted for clarity

Figure 2.5: Phase-averaged plot of turbulent kinetic energy (based on random velocity
fluctuations) and velocity vector field in a mid-way baffle plane. The spac-
ing between the vectors in all subfigures is approximately the same.
2.5. Results 31

(a) Experiment (b) RANS

(c) LES: Smagorinsky model (d) LES: Voke model

Figure 2.6: Phase-averaged plot of turbulent kinetic energy (based on random velocity
fluctuations in the vicinity of the impeller.

simulation show good correspondence with respect to the form and center of the lower
circulation loop. With respect to the upper circulation loop, a deviation of the separation
point is observed between the LES and experimental results. The experimental result
shows an average upward flow near the tank wall till z/T = 0.8, whereas the LES results
with the Smagorinsky and Voke model show an upward flow till z/T = 0.73 and z/T =
0.7, respectively. The separation location of the upper circulation loop (z/T = 0.8) has
been well captured by the RANS simulation, although the latter shows some spurious flow
at the top of the tank. It is believed that this spurious flow is associated with the fact that
the run has not fully reached its pseudo-stationary state, since it is more pronounced in the
early stages of the simulation (not shown). The circulation in the lower circulation loop
is stronger compared to the experimental result. We further note that the k and DNS
32 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

0.4
0.38 r/T=0.183 r/T=0.25 r/T=0.317
0.36
z/T (-)

0.34
0.32
0.3 LDA LDA LDA
RANS RANS RANS
0.28 LES(S) LES(S) LES(S)
LES(V) LES(V) LES(V)
0.26
0 0.25 0.5 0.75 0 0.25 0.5 0.75 0 0.25 0.5 0.75
vr /v tip (-) vr /v tip (-) vr /v tip (-)

(a) Radial velocity component at radial locations r/T = 0.183 (left), r/T = 0.25 (middle) and r/T =
0.317 (right).

0.4
0.38 r/T=0.183 r/T=0.25 r/T=0.317
0.36
z/T (-)

0.34
0.32
0.3 LDA LDA LDA
RANS RANS RANS
0.28 LES(S) LES(S) LES(S)
LES(V) LES(V) LES(V)
0.26
0 0.25 0.5 0.75 0 0.25 0.5 0.75 0 0.25 0.5 0.75
vt /v tip (-) vt /v tip (-) vt /v tip (-)

(b) Tangential velocity component at radial locations r/T = 0.183 (left), r/T = 0.25 (middle) and
r/T = 0.317 (right).

0.4
0.38 r/T=0.183 r/T=0.25 r/T=0.317
0.36
z/T (-)

0.34
0.32
0.3 LDA LDA LDA
RANS RANS RANS
0.28 LES(S) LES(S) LES(S)
LES(V) LES(V) LES(V)
0.26
0 0.025 0.05 0.075 0 0.025 0.05 0.075 0 0.025 0.05 0.075
k/v 2tip (-) k/v 2tip (-) k/v 2tip (-)

(c) Kinetic energy of the random velocity fluctuations at radial locations r/T = 0.183 (left),r/T = 0.25
(middle) and r/T = 0.317 (right).

Figure 2.7: Phase-averaged axial profiles of the radial (a) and tangential (b) velocity com-
ponents and the kinetic energy of the random velocity fluctuations (c) at three
radial locations.
2.5. Results 33

results of Bartels et al. (2001) also predict the wall separation point at a too low position
(z/T = 0.7).
Velocity fluctuations in a turbulently stirred tank are partly periodic (directly related
to the blade passage frequency) and partly random (turbulence). As a result, the kinetic
energy can be divided in a random part and a coherent part. The total kinetic energy, k tot ,
in the velocity fluctuations is:

1 2 
ktot = kcoh + kran = ui u i 2 (2.5)
2
where kcoh and kran are the coherent and random contributions to the total turbulent ki-
netic energy, respectively. Note the summation convection over the repeated index i. The
(time) averages are over all velocity samples, irrespective of the angular position of the
impeller.
The random part of the kinetic energy can be determined if phase-resolved average
data are available:
1 
kran = < ui2 > < ui 2 > (2.6)
2
with ui the ith resolved velocity component sample at the angular position . Conse-
quently, the time averages are over the velocity samples at the angular position . The
notation < > denotes averaging over all angular positions.
The kinetic energy in a large eddy simulation is partly resolved and partly unresolved.
The kinetic energy residing at the subgrid levels can be estimated based on isotropic, local-
equilibrium mixing-length reasoning (Mason & Callen, 1986). A comparison between the
grid scale and subgrid scale kinetic energy (instantaneous realization) in a mid-way baffle
plane showed that about 0.1% of the kinetic energy is at the subgrid scale level near the
impeller and less in the rest of the tank (not shown), and the subgrid scale contribution is
therefore negligible. In the following results, the resolved kinetic energy is presented.
The turbulent kinetic energy distributions plotted in Figures 2.5 and 2.6 are based on
the random velocity fluctuations. In case of the measurements and LES (Figures 2.5a,c,d
and 2.6a,c,d), the random part of the turbulent kinetic energy was obtained by averaging
the phase-resolved experimental/numerical results, Eq (2.6). In case of the transient RANS
simulation (Figures 2.5b and 2.6b), the coherent fluctuations are resolved in the transient
calculations, but they are not included in the turbulent kinetic energy solved for by the
RANS model. As a result, the RANS turbulent kinetic energy is regarded here to be based
on random velocity fluctuations.
The results show low turbulence levels in the bulk flow. The result of the Voke subgrid
scale model shows a slight decrease in turbulent kinetic energy in the bulk flow, with re-
34 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

spect to the result based on the Smagorinsky model (compare Figure 2.5c and Figure 2.5d).
This observation contrasts to what is expected; a reduction of the eddy viscosity in the bulk
would lead to increased levels of turbulent kinetic energy.
High turbulent kinetic energy levels are encountered in the impeller outflow region,
see Figure 2.6. The structure of the kinetic energy distribution of the LES calculations
agrees better with the experimental results than the agreement between the RANS and
experimental results.
Figure 2.7 shows axial profiles of the radial and tangential velocity components and
the kinetic energy based on random velocity fluctuations at three radial locations. The
agreement of the radial velocity component between simulations and experiments is good,
although the upward directed radial impeller outflow is more pronounced in the LES with
the Voke subgrid scale model (see Figure 2.7a). All simulations overestimate the tangen-
tial velocity component at the impeller tip (i.e. r/T = 0.183; z/T = 0.32 0.35). While
the agreement of the tangential velocity component for the LES is good at r/T = 0.25 and
r/T = 0.317, the RANS predictions differ from the experimental results. At r/T = 0.183
the turbulent kinetic energy is underpredicted by all simulations, but the agreement of the
LES predictions with experimental data is better than that for RANS. Also at r/T = 0.25
and r/T = 0.317 the RANS predictions significantly underpredict the turbulent kinetic
energy, whereas LES shows an overestimation between z/T = 0.32 and z/T = 0.37.
However, the spreading rate (i.e. the width of the profiles) of turbulent kinetic energy pre-
dicted by LES is in accordance with the experimental data. Except for the radial velocity
component, the differences between the two subgrid scale models applied in the LES are
small.
Mesh refinement and/or time-step refinement are not expected to improve substantially
the predictions of the turbulent kinetic energy in the transient RANS simulation (Ng et al.,
1998). It is believed that the discrepancies stem from the turbulence model employed. The
SST eddy viscosity model used here assumes locally isotropic turbulence. However, if the
local flow field is anisotropic (i.e. u0 6= v 0 6= w0 ) the application of such a turbulence
model may lead to an inaccurate prediction of the local turbulent kinetic energy.

2.5.3 Phase-resolved flow field and turbulence levels


In a phase-resolved experiment or simulation the position of the impeller blade, with re-
spect to the plane where experimental and numerical results were extracted, is recorded.
This allows for a reconstruction of the mean flow field and its fluctuations as function of
the impeller angle.
Phase-resolved, mean flow fields in the vicinity of the impeller are shown in Figures 2.8
2.5. Results 35

(a) Experiment (b) RANS

(c) LES: Smagorinsky model, not all (d) LES: Voke model, not all vectors
vectors are plotted for clarity are plotted for clarity

Figure 2.8: Phase-resolved plot of turbulent kinetic energy (based on random velocity fluc-
tuations) and velocity vector field at 15o behind the impeller blade.

and 2.10, at angles 15o and 45o , respectively. An important flow phenomenon is the trail-
ing vortex system, that develops in the wake of the turbine blade (Figure 2.8a), and is then
advected by the impeller stream into the bulk of the tank. The trailing vortices are shown
in more detail in Figure 2.9. They die out, and the radial impeller outflow is directed
slightly upward (Figure 2.10a). The LES prediction with the Smagorinsky subgrid scale
model (Figures 2.8c and 2.10c) and the RANS prediction (Figures 2.8b and 2.10b) show a
good representation of the trailing vortex system, although the upward directed radial out-
flow was not predicted by RANS. The Voke subgrid scale model (Figures 2.8d and 2.10d)
predicts the lower trailing vortex as being much stronger than the upper vortex at 15 0 ,
and a remnant of that lower vortex at 450 , which does not correspond to the experimental
36 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

(a) (b) (c) (d)

Figure 2.9: Phase-resolved flow field at 15o behind the impeller blade in the domain
0.07 r/T 0.2; 0.29 z/T 0.38. The experimentally measured flow
field is depicted in (a), and the flow field simulated by RANS, LES (Smagorin-
sky) and LES (Voke) in (b-d), respectively.

findings.

It can be seen that the vortex structure is accompanied with high turbulent kinetic
energy levels. At 15o behind the impeller blade at r/T values in between about 0.24
and 0.28 a region is observed with high turbulent kinetic energy levels. This region of
high turbulent kinetic energy is caused by the vortex system related to the previous blade
passage. With the LES model, this region is slightly overpredicted. The reduction of the
Smagorinsky eddy viscosity in that region by the Voke subgrid scale model is relatively
small (see Figure 2.4), and consequently differences of the kinetic energy between both
subgrid scale models are marginal. At 45o , the structure and levels of turbulent kinetic
energy are well resolved. For the RANS calculation, the turbulent kinetic energy is mostly
underestimated in areas where the trailing vortices are observed (i.e. near the tip of the
blade). This underestimation may again be associated to the turbulence model used here.

Figure 2.11 shows profiles of the turbulent kinetic energy at r/T = 0.183, r/T = 0.25
(15 degrees behind impeller blade) and r/T = 0.217 (45 degrees behind the impeller
blade), which cover high regions of turbulent kinetic energy caused by the trailing vortices.
In general, this figure shows a better agreement between LES and experimental results,
compared to that between RANS and experiments. The RANS results significantly and
systematically underpredict the turbulent kinetic energy. The agreement between the LES
and experimental results of the width of the profiles is good, whereas Figure 2.11a shows
an underestimation of the turbulence levels between z/T = 0.33 and z/T = 0.36 and
Figure 2.11c shows an overestimation between z/T = 0.32 and z/T = 0.36.
2.5. Results 37

(a) Experiment (b) RANS

(c) LES: Smagorinsky model, not all (d) LES: Voke model, not all vectors
vectors are plotted for clarity are plotted for clarity

Figure 2.10: Phase-resolved plot of turbulent kinetic energy (based on random velocity
fluctuations) and velocity vector field at 45o behind the impeller blade.

2.5.4 Energy dissipation


The dissipation rate of turbulent kinetic energy () is important in many mixing applica-
tions, since it controls the flow at the micro-scale and acts as a controlling parameter in,
e.g, break-up and coalescence processes. Unfortunately, it is hardly possible to directly
measure the dissipation rate. Indirect ways of measuring via turbulence intensities and
length scales require assumptions about the nature of turbulence (e.g. isotropy or equilib-
rium) which are not appropriate in stirred-tank flow. Since the predictions of the turbulent
kinetic energy correspond fairly well to the experimental results, it seems fair to assume
that the energy dissipation rate distribution is predicted with acceptable accuracy by the
simulations. By assuming local equilibrium between production and dissipation at and
38 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

0.4
0.38 15; r/T=0.183 15; r/T=0.25 45; r/T=0.317
0.36
z/T (-)

0.34
0.32
0.3 LDA LDA LDA
RANS RANS RANS
0.28 LES(S) LES(S) LES(S)
LES(V) LES(V) LES(V)
0.26
0 0.025 0.05 0.075 0.1 0.125 0 0.025 0.05 0.075 0.1 0.125 0 0.025 0.05 0.075 0.1 0.125
k/v 2tip(-) k/v 2tip(-) k/v 2tip (-)

Figure 2.11: Phase-resolved profiles of turbulent kinetic energy (based on random velocity
fluctuations) at 15o degrees behind the impeller blade, r/T = 0.183 (left)
and r/T = 0.25 (middle), and at 45o degrees behind the impeller blade at
r/T = 0.217 (right).

below subgrid-scale level, the energy dissipation rate can be coupled to the deformation
rate:
= ( + e )S 2 (2.7)
Figure 2.12 shows that the dissipation rate distribution (i.e. the sum of the resolved scale
and subgrid scale contributions) in the tank is very inhomogeneous, stressed by the log-
scale used for the gray-scale coding. The energy dissipation rate in the impeller region is
several orders of magnitude higher than in the bulk region.

The transient RANS result of the energy dissipation rate reveals higher energy dissipa-
tion rates compared to the LES results. Furthermore, the RANS prediction of the spreading
rate of energy dissipation in the discharge flow is smaller than that predicted by LES. Be-
yond that, the structure of the distributions are similar. A region at the top of the tank and
the two lobes in the impeller outflow region show increased energy dissipation rates, which
are thought to be due to the fact that the run had not yet fully reached its pseudo-stationary
state.

In comparing the results of the Voke model to those of the Smagorinsky model (see
Figures 2.12b,c), the energy dissipation rate is reduced in the bulk flow, whereas the energy
dissipation rate in the impeller region is more or less unaffected. This effect is expected,
since the eddy viscosity in the Voke model is more suppressed in the bulk region (more
or less resolved flow) of the tank, leading to a lower energy dissipation rate in that region,
see Eq. 2.7.
2.5. Results 39

(a) RANS (b) Smagorinsky model

(c) Voke model

Figure 2.12: Phase-averaged plot of dissipation rate of turbulent kinetic energy in a mid-
way baffle plane.
40 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

(a) Smagorinsky model (b) Voke model

Figure 2.13: Phase-averaged plot of | A | in a mid-way baffle plane.

2.5.5 Turbulence anisotropy


The k- model is still widely used for simulations in stirred tank configurations, also in
the present work. This model is an eddy-viscosity model and it locally assumes isotropic
turbulent transport. In rotating and/or highly three-dimensional flows, the k- model is
known to be inappropriate (Wilcox, 1993). LES at least predicts the resolved part of the
Reynolds stresses and thus may provide an assessment of the isotropy assumptions in the k-
model. A treatment in terms of (an)isotropy is beneficial for a meaningful interpretation
of the Reynolds stresses (Derksen et al., 1999). The Reynolds stress data will be presented
in terms of the anisotropy tensor aij and its invariants. The anisotropy tensor, defined as:

ui uj 2
aij = ij (2.8)
k 3
has a first invariant equal to zero by definition. The second and third invariant respectively
are A2 = aij aji and A3 = aij ajk aki . The range of physically allowed values of A2 and
A3 is bounded in the (A3 , A2 ) plane by the so-called Lumley triangle (Lumley, 1978). In
order to characterize anisotropy with a single parameter, the distance from the isotropic
2.6. Conclusions 41

Figure 2.14: The locations within the Lumley triangle of the phase-averaged points in a
plane mid-way between two baffles. Not all points are plotted for clarity. The
borders represent different types of turbulent flows: 3-D isotropic turbulence,
2-D axisymmetric turbulence, 2-D turbulence and 1-D turbulence (Lumley,
1978).

p
state | A |= A22 + A23 was defined.
Phase-averaged results of | A | in a plane mid-way between two baffles are presented
in Figure 2.13. The strongest deviations from isotropy, i.e. the highest levels of | A |,
occurred in the discharge flow, the inflow regions (i.e. below and above the impeller),
the boundary layers and at the separation points. The turbulence anisotropy at the separa-
tion points is advected into the bulk flow. In the recirculation loops, turbulence is nearly
isotropic. Figure 2.14 shows the distribution of the invariants in the (A3 , A2 ) plane ob-
tained with the Smagorinsky model. The points are mostly clustered in the lower part (i.e.
almost isotropic turbulence) of the so-called Lumley triangle. However, Figures 2.13 and
2.14 both show that the assumption of mean isotropic turbulence in stirred vessel config-
urations is questionable. This may explain the underestimation of the turbulent kinetic
energy levels in the transient RANS results observed in sections 2.5.2 and 2.5.3.

2.6 Conclusions
In this paper, results of Large-Eddy Simulations (LES) and RANS simulations were as-
sessed by means of detailed LDA experiments globally throughout the tank and locally
near the impeller. Two subgrid scale models were investigated, the standard Smagorinsky
model and a modified Smagorinsky model (i.e. the Voke subgrid scale model). With a
view to chemical mixing processes, the LES methodology has clear, distinct advantages
42 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

over a RANS type of approach, as velocity fluctuations, and consequently the Reynolds
stresses and turbulent kinetic energy, are resolved down to the scale of the numerical grid.
As a result, there is a clear (spectral) distinction between the resolved and unresolved
scales that is indefinite in a RANS approach.
As the experiments were conducted at the lower limit of the turbulent regime, the
behavior of the subgrid scale models applied was checked. In the bulk flow, all scales were
more or less resolved, whereas in the impeller region and discharge flow the Smagorinsky
and Voke eddy viscosity became of the order of the kinematic viscosity. As a result, a
direct numerical simulation with the LES grid resolution will not accurately capture the
smallest scales in the impeller (outflow) region.
The simulated phase-averaged flow fields were in good agreement with the experimen-
tal result. The RANS flow field showed deviations in the tangential velocity component in
the discharge flow, whereas the wall separation point of the upper circulation loop in the
LES flow field differed from the experimental result for both subgrid scale models. All
simulations overpredicted the tangential velocity component at the center of the impeller
tip. The disagreement between the experimental and LES results of the wall separation
point is thought to be due to the wall damping function used in this work. Therefore, other
wall damping functions will be explored in future investigations.
The development of the trailing vortex system was well represented by the RANS
simulation and LES with the Smagorinsky subgrid scale model. The LES with the Voke
subgrid scale model showed the lower vortex as being much stronger than the upper vor-
tex, which did not correspond to the experimental findings. The upward directed radial
impeller outflow was well represented with LES, and was not found in the RANS simula-
tion.
In general, the structure and levels of the turbulent kinetic energy in the impeller dis-
charge flow were better represented by LES than by the RANS simulation. The levels
predicted by the RANS simulation systematically and significantly underpredicted the ex-
perimentally obtained levels. The levels of turbulent kinetic energy obtained by LES were
overestimated between z/T = 0.32 and z/T = 0.37, except for the averaged levels at
r/T = 0.183 that showed an underestimation between z/T = 0.26 and z/T = 0.4. In
order to improve the LES kinetic energy predictions in the impeller outflow region, either
the Smagorinsky model should also be evaluated at higher Reynolds numbers (e.g. Derk-
sen & Van den Akker (1999)), or more advanced subgrid scale models (e.g. mixed scale
model, dynamic Smagorinsky model, see Lesieur & Metais (1996)) might be attempted.
The dissipation of turbulent kinetic energy is very inhomogeneously distributed through-
out the tank, with high levels in the impeller outflow region and low levels in the bulk of
the tank. The spreading rate of energy dissipation predicted by the RANS simulation was
2.6. Conclusions 43

smaller than that by LES; this was also observed for the spreading rate of the turbulent
kinetic energy (i.e. by the widths of the profiles). The levels obtained by the RANS simu-
lation were higher than those obtained by LES.
As velocity fluctuations are resolved by the numerical grid, LES enables an estima-
tion of the turbulence (an)isotropy throughout the tank. Nearly isotropic turbulence was
observed in the circulation loops. However, in the impeller stream, the boundary layers,
and at the separation points turbulence was found more anisotropic, probably caused by
the local, high shear rates. Consequently, the choice for making use of an isotropic eddy
viscosity turbulence model should be treated with care.
For the Reynolds number being in the lower limit of the turbulent regime, the Voke
subgrid scale model was explored that blends smoothly between a DNS in the bulk flow
and a LES in the impeller region and discharge flow. This has been illustrated by the Voke
eddy viscosity distribution, that showed a more pronounced relative reduction of the Voke
eddy viscosity in the bulk of the tank compared to that in the impeller region and discharge
flow. Despite the potential of the Voke subgrid scale model for application at relatively low
Reynolds number flows, this work did not show significant improvements in the flow field
results.
In conclusion, a transient RANS simulation is able to provide an accurate represen-
tation of the flow field, but fails in the prediction of the turbulent kinetic energy in the
impeller region and discharge flow where most of the mixing takes place. In order to have
a fair comparison with phase-averaged and phase-resolved experimental data, the transient
RANS simulation has been executed in a full flow domain. Similar to a LES, such a sim-
ulation is time consuming as well. The large eddy simulations performed have shown that
LES provides an accurate picture of the flow field, and reasonable data of the turbulent
kinetic energy directly obtained from the resolved velocity fluctuations. As LES resolves
the velocity fluctuations locally, information of the turbulence (an)isotropy in e.g. a stirred
tank geometry is now available.
44 Chapter 2. Assessment of large eddy simulations on the flow in a stirred tank

Nomenclature

Roman Description Unit


aij anisotropy tensor -
A+ constant in Eq. (2.3) -
A2 second invariant of anisotropy tensor -
A3 third invariant of anisotropy tensor -
cs Smagorinsky constant -
D impeller diameter m
H height of the tank m
k turbulent kinetic energy m2 .s2
kcoh turbulent kinetic energy based on coherent velocity m2 .s2
fluctuations
kran turbulent kinetic energy based on random velocity m2 .s2
fluctuations
N impeller speed s1
r radial coordinate m
S resolved deformation rate s1
T tank diameter m
ui velocity component i m.s1
ui velocity component i linked to an angular position m.s1
u friction velocity m.s1
y distance from wall m
y+ distance from wall in viscous wall units -
vr radial velocity component m.s1
vt tangential velocity component m.s1
vtip impeller tip speed m.s1
z axial coordinate m

Greek Description Unit


constant in Eq. (2.4) -
lattice spacing m
energy dissipation rate m2 .s3
angle rad
mix mixing length m
dynamic viscosity kg.m1 .s1
2.6. Conclusions 45

kinematic viscosity m2 .s1


e Smagorinsky eddy viscosity m2 .s1
e,V Voke eddy viscosity m2 .s1
|A| distance from isotropic state -
3 Macro-instability uncovered in a
Rushton turbine stirred tank

Low-frequency mean flow variations, identified experimentally in various stirred


tank geometries, are studied by means of large eddy simulations in a Rushton
turbine stirred tank. The focus is on flow structures and the spectral character-
istics of the velocity components. The lattice-Boltzmann Navier-Stokes solver
and the Smagorinsky subgrid-scale model are adopted for solving the stirred
tank flow, and boundary conditions are imposed by means of an adaptive
force-field technique. Simulations performed at Reynolds numbers 20, 000
and 30, 000 on grid sizes of 1203 , 1803 and 2403 grid nodes confirm the ex-
perimentally found flow variations at various monitoring points in the bulk
flow. The period of the flow variations found in the bulk of the tank corre-
sponds to approximately 250 blade passage periods. A simulation performed
at a Reynolds number of 12, 500 showed pronounced flow variations with a
time scale of about 65 blade passage periods, which is consistent with ex-
perimental observations. Transient flow field results in the bulk flow of the
tank uncover a whirlpool type of precessing vortex of which the impact on the
kinetic energy of the velocity fluctuations is further analyzed.
Key words: stirred tank, mixing, turbulence, macro-instability, LES
This chapter has been published in AIChE J. 51(10), 2419-2432 (2004).

3.1 Introduction
3.1.1 Flow phenomena in stirred tanks
Stirred tanks have always played an important role in industry, because of their mixing
abilities brought about by the agitation of the impeller together with baffles along the tank
48 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

wall preventing a solid body rotation of the fluid. As a result, a large number of studies
focused on the flow phenomena encountered in such tanks. Rushton et al. (1950) and
Holmes et al. (1964) were the among the first to report on power consumption and circula-
tion times for various impeller and tank configurations, which characterize the global flow
field. More recently, various experimental and numerical studies reported on local flow
information, with a view to optimizing mixing processes (e.g. Yianneskis et al., 1987;
Schfer et al., 1998; Derksen et al., 1999; Derksen & Van den Akker, 1999). For in-
stance, the flow structures are identified to be highly three-dimensional and complex with
trailing vortices and high turbulence levels (and dissipation rates) in the vicinity of the
impeller. This knowledge of the local flow is of importance as rate-limiting phenomena
such as droplet break-up (Tsouris & Tavlarides, 1994; Luo & Svendsen, 1996) or chemical
reactions (Van Vliet et al., 2001) are influenced by the small-scale mixing.
One of the complications in stirred tank flows is the presence of Macro-Instabilities
(MI, i.e. low-frequency mean flow variations) that affect the flow patterns and conse-
quently mixing performance. Different types of instabilities have been identified experi-
mentally, and were reviewed in the paper of Nikiforaki et al. (2002). One of these types
was identified as a whirlpool type of vortex moving around the tank centerline (Yianneskis
et al., 1987). A large number of studies has been concerned with this instability and the
frequency of its precessing motion at different operating conditions, tank and impeller
types, and impeller clearances (e.g. Nikiforaki et al., 2002; Haam et al., 1992; Hasal et al.,
2000; Myers et al., 1997). However, the frequency was found difficult to reproduce. Niki-
foraki et al. (2002) have experimentally pointed out that the whirlpool type of vortex can
contribute significantly to the kinetic energy contained in the velocity fluctuations. This
effect has similarities to the coherent contribution of the flow periodicity near the impeller
blades to the kinetic energy (see e.g. Yianneskis et al., 1987; Derksen & Van den Akker,
1999).

3.1.2 MI significance to multi-phase chemical processes


In chemical processing, the velocity field is used to accomplish objectives such as heat
transfer, mass transfer and chemical reaction. Various investigations provide evidence of
the significance of the MI to multi-phase chemical processes. Haam et al. (1992) mea-
sured the local heat transfer flux and temperature on the tank wall. Time variations in the
local heat transfer coefficient were observed, and the authors hypothesized that these time
variations were due to the precession of an axial vortical structure. Feed stream intermit-
tency was studied by Houcine et al. (1999) by means of laser induced fluorescence (LIF).
The feed stream jet was injected at the baffle in the upper part of the tank. Three stages of
3.1. Introduction 49

the feed stream were observed, of which one of them was characterized as effective inter-
mittency with a period of 1-2s. This phenomenon can be linked to the existence of an MI.
Derksen (2003) provided another example of the presence of precessing vortices in a solids
distribution system. The lower vortex may contribute to resuspension of the solids lying
on the bottom of the tank, and consequently mass transfer is enhanced. But, the voidage
in the vortex itself may be regarded as a lost volume. Finally, the upper precessing vortex
may cause surface aeration which often is an unwanted process.

3.1.3 RANS simulation

The presence of low-frequency coherent fluctuations associated with macro-instabilities


in stirred tanks have important implications for computational flow models. Commercial
software packages generally solve the Reynolds-Averaged Navier-Stokes (RANS) equa-
tions in combination with a closure model for the Reynolds stresses. The stationary and
rotating components are handled by means of interface models, such as the multiple frames
of reference (MFR) technique (CFX-5, 2002). Precessing vortices moving around the tank
centerline contribute to the flow unsteadiness, and therefore exclude models that allow
for a steady-state approximation and domain reduction through geometrical symmetries
(e.g. frozen-rotor and circumferential-average models). The transient sliding mesh inter-
face, on the contrary, is able to simulate the unsteady fluid motion as the flow variation
both in time and space is taken into account. However, it is unclear what the effect is
of the location of the interface between the rotating and stationary meshes on the simu-
lated unsteadiness of the mean flow. Subsequently, the turbulence modeling is a point for
discussion. The k- model is often used to account for the turbulent fluctuations. This
eddy-viscosity model, which assumes isotropic turbulent transport, is known to be inap-
propriate in rotating and/or highly three-dimensional flows (Wilcox, 1993). Assessment
of the isotropy assumptions by LDA experiments (Derksen et al., 1999) and large eddy
simulations (Hartmann et al., 2004b) has indicated turbulence anisotropy in the impeller
stream. Furthermore, in a transient RANS simulation it is not a priori clear which part
of the fluctuations is temporally resolved, and which part is taken care of by the turbu-
lence model. This especially applies to flows with no clear spectral separation between the
low-frequency, coherent fluctuations on one side, and turbulent fluctuations on the other.
Experiments in turbulently stirred tanks (e.g. Nikiforaki et al., 2002), as well as our LES
show no clear separation between MI-related fluctuations, and random fluctuations.
50 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

T/10 D = T/3
0.75D 0.16D
r

H=T
T/2 0.04D 0.04D 0.2D
z
0.24D
T 0.25D

Figure 3.1: Cross-section of the tank (left). Plan view and cross-section of the impeller
(right). At the top level there is a lid. The impeller, with diameter D, is a
Rushton turbine.

3.1.4 Large eddy simulation


We studied the flow macro-instability numerically by means of Large Eddy Simulation
(LES). This methodology is embedded within the lattice-Boltzmann scheme for discretiz-
ing the Navier-Stokes equations, and the use of an interface model is avoided by modeling
the geometry with an adaptive force field technique (Derksen & Van den Akker, 1999). In
LES a clear distinction is made between the resolved and unresolved scales, as the range
of resolved scales is limited by the numerical grid. The effect of the unresolved subgrid
scales on the resolved large scales is taken into account with a subgrid-scale model. The
LES methodology has proven to be a powerful tool to study and visualize stirred tank
flows (Eggels, 1996), because it accounts for the unsteady and periodic behavior of these
flows and can effectively be employed to explicitly resolve phenomena directly related to
the unsteady boundaries. In terms of velocity fields and kinetic energy contour plots in
the vicinity of the impeller and in the bulk of the tank, assessment of LES proves it to be
an accurate method for simulating stirred tank flow (Hartmann et al., 2004b). Roussinova
et al. (2003) have shown that the circulation pattern instability (i.e. another type of macro-
instability at a different time scale) can effectively be resolved by means of LES. In this
work, the LES methodology will be adopted to identify the whirlpool type of precessing
vortex and its spectral characteristics (e.g. characteristic frequency and fluctuation levels).

3.2 Flow system


The stirred vessel reactor used in this research was a standard configuration cylindrical
vessel of diameter T , with four equi-spaced baffles of width 0.1T placed at the perimeter
3.3. Simulation procedure 51

of the tank, and with liquid height H = T . A lid was positioned at height H. A standard
Rushton turbine disk impeller with six blades was mounted at mid-height of the tank. Flow
visualization results of Yianneskis (2004) indicated that the macro-instabilities were more
pronouned and more clearly defined with the impeller located at mid-height of the tank. A
schematic representation of the geometry is given in Figure 3.1. Provided that geometric
similarity is maintained, the flow system can be fully characterized by the flow Reynolds
number, defined as:

N D2
Re = (3.1)

with N the impeller speed, D the impeller diameter and the kinematic viscosity of
the working fluid.

3.3 Simulation procedure


3.3.1 Large Eddy Simulation
Many industrially relevant flows are associated with high Reynolds numbers. A Direct Nu-
merical Simulation (DNS) would require enormous amounts of grid cells and time steps,
in order to capture all length and time scales present in the flow. In a Large Eddy Simu-
lation (LES), the small scales in the flow are assumed to be universal and isotropic, and
the effect the small scales have on the larger scales is modeled with a subgrid-scale model.
In this research, the well-known Smagorinsky model (Smagorinsky, 1963) was used. This
is an eddy-viscosity model with a subgrid-scale eddy viscosity, e , which is related to the
local, resolved deformation rate:

e = 2mix S 2 (3.2)

where mix is the mixing length of the subgrid-scale motion and S 2 is the resolved
deformation rate, defined as:
 2
2 1 ui uj
S = + (3.3)
2 xj xi
with ui the resolved ith velocity component. Note the summation convention over the
repeated indices i, j. The ratio between the mixing length and the lattice spacing, , is
constant, called the Smagorinsky constant (cs ). However, near the walls, the subgrid-scale
stresses should vanish, which is not automatically guaranteed in Eq. (3.2). In general, this
52 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

Table 3.1: Numerical setup. The grid spacing and time step t are given in the first
two columns. Subsequently, the Reynolds number, the number of grid nodes
and the number of subdomains are given. The number of simulated impeller
revolutions times the (wall-clock) time to simulate one impeller revolution, ,
gives the duration of the simulation.
Case t Re Res. Subd. # Imp. revs.
I T /120 (1200N )1 20, 000 1203 4 570 0.69 h
II T /180 (1600N )1 12, 500 1803 6 570 2.03 h
III T /180 (1600N )1 20, 000 1803 6 570 2.03 h
IV T /180 (1600N )1 30, 000 1803 6 570 2.03 h
V T /240 (2500N )1 30, 000 2403 8 150 5.73 h

is accomplished by a reduction of the length scale mix toward the wall. The Van Driest
wall damping function (Van Driest, 1956) is applied at the tank walls, which determines
the mixing length as:
 
y+
mix = cs 1 e A+ (3.4)

where y + is the distance to the wall in viscous wall units and A+ is a constant taken
equal to 26. A value of 0.10 was adopted for cs , which is a typical value used in LES
computations of shear-driven turbulent flows (Piomelli et al., 1988).
A lattice-Boltzmann method (Chen & Doolen, 1998) was used for solving the filtered
momentum equations. The specific scheme we used was developed by Somers (1993).
The entire tank was simulated on a uniform, cubic computational grid. Inside the com-
putational domain, the no-slip boundary conditions at the cylindrical tank wall, the baffles,
the impeller, and the impeller shaft were imposed by an adaptive force-field technique
(Derksen & Van den Akker, 1999).

3.3.2 Simulation aspects


The MI was studied in detail at Re=30, 000. Additional simulations have been performed
at Re=20, 000 and Re=12, 500. The size of the computational grid was varied in order to
study the influence of the grid-size on the solution. The settings for the simulations are
given in Table 3.1.
The computer code runs on a parallel computer platform by means of domain de-
composition: the computational domain was horizontally (i.e. perpendicular to the tank
centerline) split in a number of equally-sized subdomains.
3.3. Simulation procedure 53

0.88

0.76

0.64
0.37 0.63

z/T (-)
0.58
0.50
0.5 0.77
0.43
0.36
2r/T (-)
0.24

0.12

Figure 3.2: Locations of the horizontal and vertical lines on which the three velocity com-
ponents are stored. The vertical plane is located mid-way between two baffles.

On every grid node, 21 (18 directions for the LB-particles and 3 force components)
single-precision, real values need to be stored. The memory requirements of the simulation
are proportional to the grid size, resulting in an executable that occupies in case of the 180 3
grid 1803 21 4 0.5 GByte of memory. The simulations were performed on an in
house PC cluster with Athlon 1800+ MHz processors using an MPI message passing tool
for communication within the parallel code.
Within the lattice-Boltzmann framework the time step and lattice spacing are non-
dimensional numbers, which are equal to one. Hence, dimensions are expressed in lattice
spacings and velocities in lattice units (i.e. lattice spacings per time step). In Case I for
example, the tank diameter corresponds to 120 lattice spacings and one impeller rotation
corresponds to 1200 time steps.
During the simulations, the axial, radial and tangential velocity components were
stored on six horizontal lines and four vertical lines. Three horizontal lines were located
below the impeller and three above the impeller. The four vertical lines were located in the
impeller outflow stream at different radial positions. The locations of the lines are shown
in Figure 3.2. In case of the 1803 -grid, the radial resolution at the horizontal lines was
0.022T , with a total number of 23 monitoring points at each line and the first monitoring
54 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

point at 2r/T = 0. The axial resolution at the vertical lines is 0.006T , with a total number
of 25 monitoring points at each line and the first monitoring point at z/T = 0.433. This
results in a total number of 238 monitoring points. The sampling frequency, f s , equals one
in the lattice-Boltzmann framework, as the data are stored at each time step.
The simulations were started from either a zero-velocity field or from a flow field of a
previous simulation. After roughly 20 30 impeller revolutions a quasi steady state was
reached. This was checked by monitoring the total kinetic energy as function of time. A
flow field was transferred from one grid to the other by means of linear interpolation.

3.4 Results
Most of the results presented in this section are based on case IV; Re=30, 000. This case
serves as a reference case.

3.4.1 Single flow field realizations


The flow field in the vicinity of the tank centerline is shown Figures 3.3 and 3.4 at axial
positions z/T = 0.12 and z/T = 0.88, respectively. The contours represent the ratio of
the axial vorticity component and the impeller angular speed; a positive value corresponds
to fluid rotation in the (clockwise) direction of the impeller rotation. In Figure 3.3a, the
main vortical structure is located in the upper-right corner of the graph. When viewing the
graphs in sequence, with a time interval of 11 impeller revolutions between each graph, a
precessional motion of the vortex in a clock-wise direction (i.e. in the same direction as
the impeller rotates) around the tank centerline is observed. Note the long period of the
precessional motion, being roughly 40 50 impeller revolutions.
Compared to Figure 3.3, Figure 3.4a shows a weaker vortical structure in the upper-left
corner of the graph. The black circle represents the impeller shaft. Again, the sequence
of the graphs shows the vortex precessing around the impeller shaft. A direct comparison
between Figures 3.3 and 3.4 indicates a phase difference between the upper and lower
moving vortices.

3.4.2 Time-series
The transient behavior of the flow strongly depends on the position in the tank. To illus-
trate the broad range of frequencies encountered in stirred tank flows, the simulated time
series of the radial velocity component is depicted in Figure 3.5 at two typical positions,
respectively. At the impeller tip, Figure 3.5a, most of the fluctuations are periodic as a
3.4. Results 55

0.2 0.2
0.5 vtip 0.5 vtip
z /0 z /0
1 1
0. 1 0.1
0.6 0.6

2y/T (-)
2y/T (-)

0.2 0.2
0 0
0.2 0.2

0.6 0.6
0.1 0.1
1 1
PSfrag replacements PSfrag replacements
0.2 0.2
0.2 0.1 0 0.1 0.2 0.2 0.1 0 0.1 0.2
2x/T (-) 2x/T (-)

(a) Snapshot at the level z/T = 0.12. (b) As in (a) after 11 impeller revolu-
tions.

0.2 0.2
0.5 vtip 0.5 vtip
z /0 z /0
1 1
0.1 0.1
0.6 0.6
2y/T (-)
2y/T (-)

0.2 0.2
0 0
0.2 0.2

0.6 0.6
0.1 0.1
1 1
PSfrag replacements PSfrag replacements
0.2 0.2
0.2 0.1 0 0.1 0.2 0.2 0.1 0 0.1 0.2
2x/T (-) 2x/T (-)

(c) As in (a) after 22 impeller revolu- (d) As in (a) after 33 impeller revolu-
tions. tions.

Figure 3.3: Impressions of the flow field (Case IV) about the tank centerline at four differ-
ent points in time. The gray scales represent the axial vorticity component, z ,
made dimensionless with the impeller angular speed, 0 = 2N .
56 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

0.2 0.2
0.5 vtip 0.5 vtip
z /0 z /0
1 1
0.1 0.1
0.6 0.6

2y/T (-)
2y/T (-)

0.2 0.2
0 0
0.2 0.2

0.6 0.6
0.1 0.1
1 1
PSfrag replacements PSfrag replacements
0.2 0.2
0.2 0.1 0 0.1 0.2 0.2 0.1 0 0.1 0.2
2x/T (-) 2x/T (-)

(a) Snapshot at the level z/T = 0.88. (b) As in (a) after 11 impeller revolu-
tions.

0.2 0.2
0.5 vtip 0.5 vtip
z /0 z /0
1 1
0.1 0.1
0.6 0.6
2y/T (-)
2y/T (-)

0.2 0.2
0 0
0.2 0.2

0.6 0.6
0.1 0.1
1 1
PSfrag replacements PSfrag replacements
0.2 0.2
0.2 0.1 0 0.1 0.2 0.2 0.1 0 0.1 0.2
2x/T (-) 2x/T (-)

(c) As in (a) after 22 impeller revolu- (d) As in (a) after 33 impeller revolu-
tions. tions.

Figure 3.4: Impressions of the flow field (Case IV) about the tank centerline at four differ-
ent points in time. The gray scales represent the dimensionless axial vorticity
component. The black circle represents the impeller shaft.
3.4. Results 57

2
1.5
Vrad /vtip (-)

1
0.5
0
-0.5
0 100 200 300 400 500 0 1 2 3 4 5
Nt (-) Nt (-) PDF (a.u.)
(a) (b) (c)
0.5

0.25
Vrad /vtip (-)

-0.25

-0.5
0 100 200 300 400 500 0 1 2 3 4 5
Nt (-) Nt (-) PDF (-)
(d) (e) (f)

Figure 3.5: Traces of the simulated (i.e. Case IV) radial velocity component at positions
z/T = 0.5; 2r/T = 0.367 (a and b) and z/T = 0.12; 2r/T = 0 (d and
e). Please note the different record lengths of 500 (a and d) and 5 (b and e)
impeller revolutions, respectively. The PDFs of the time series shown in (a)
and (d) are given in (c) and (f), respectively.

result of the regular blade passage. This can be clearly seen in the enlarged time series
(Figure 3.5b); the tangential and axial velocity components show a corresponding behav-
ior (not shown). No clear low-frequency oscillations are observed.
In the bulk of the tank, Figure 3.5d, a clear low-frequency, coherent behavior of the
radial velocity component is observed. It shows a clear, cyclic variation of the mean flow
superimposed on the turbulent fluctuations. The period of the oscillations correspond with
the approximated period extracted from the flow field snapshots, around 50 impeller rev-
olutions. This periodic behavior was found less pronounced in the axial velocity com-
ponent, but the periodicity in the axial velocity component has been observed at various
other positions in the tank (not shown). The enlarged time series (Figure 3.5e) shows the
random behavior of the fluctuations, which are superimposed on the low-frequency flow
oscillation.
58 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

10000
10000
1000
100 100

P SD ()
P SD ()

1 10

0.01 1
0.1
PSfrag replacements 0.0001 PSfrag replacements
0.01
0.001 0.01 0.1 1 10 0 0.05 0.1 0.15 0.2
f /N () f /N ()

(a) z/T = 0.5; 2r/T = 0.367 (b) z/T = 0.5; 2r/T = 0.367

10000
10000
1000
100 100
P SD ()
P SD ()

1 10

0.01 1
0.1
PSfrag replacements 0.0001 PSfrag replacements
0.01
0.001 0.01 0.1 1 10 0 0.05 0.1 0.15 0.2
f /N () f /N ()

(c) z/T = 0.12; 2r/T = 0 (d) z/T = 0.12; 2r/T = 0

Figure 3.6: Power Spectral Density functions (Case IV) of the radial velocity component
in the impeller outflow region (a and b) and the bulk region (c and d). (a and
c) are log-log plots. (b and d) are frequency enlarged lin-log plots.

3.4.3 Frequency analysis


The determination of the characteristic macro-instability frequency is essential in order to
accurately assess the time scale of the flow motion. Therefore, the power density spec-
tra of the velocity components were analyzed at all monitoring points. The time series
recorded, with a length of 570 impeller revolutions, were first detrended and multiplied
with a Hanning window (Oppenheim et al., 1983). Subsequently, the power density spec-
tra were obtained by means of the FFT technique. The frequency resolution of the power
3.4. Results 59

50 25

40 20

30 15

% points
% points

20 10

10 5
PSfrag replacements PSfrag replacements
0 0
0.001 0.01 0.1 1 10 0.001 0.01 0.1 1 10
f /N () f /N ()

(a) Case IV; Re=30, 000, 1803 -grid (b) Case II; Re=12, 500, 1803 -grid

Figure 3.7: Histogram of maximum frequency peaks found in the power density spectra of
the velocity components.

density spectra equals 5701 N 0.00175N .


As an illustration, the power density spectra of the presented time series of the radial
velocity component in Figures 3.5a and 3.5d are shown in Figure 3.6. Note that only a se-
lection of frequencies in the power density spectra are presented, the maximum detectable
frequency is 0.5fs = 0.5 (the Nyquist frequency), which equals f = 800N . The spatial
resolution of the simulations, however, cannot keep up with this high frequency. A more
meaningful estimate of the highest frequency that contains flow information is based on
(twice) the grid-spacing and typical values of the RMS velocity. If we take 0.4v tip for the
latter, we get f = 32N .
In the impeller outflow region, the turbulent fluctuations are mostly periodic due to
the passage of the six impeller blades (see Figure 3.5a). This is confirmed in the power
density spectrum of the radial velocity component, Figure 3.6a; a clear and distinct peak is
observed at f = 6N . Furthermore, the power density of this frequency peak is about three
to four orders of magnitude larger than the power density of the lower frequencies. No
other distinct peaks are observed in the spectra. These observations indicate that a large
amount of the kinetic energy (i.e. the integral of the area under peak at f = 6N ) is due to
coherent velocity fluctuations caused by the blade passages only.
The time series of the radial velocity component at the monitoring point located in
the bulk region (z/T = 0.12, 2r/T = 0) clearly shows a cyclic variation of the mean
flow superimposed on the turbulent fluctuations (Figure 3.5d). This cyclic variation is
60 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

confirmed in the power density spectrum of the radial velocity component. A clear peak
is observed at frequency f = 0.0228N , which corresponds to a period of about 44 im-
peller revolutions. The power density of this peak is more than 10 times larger than that
of the other frequencies. More important, the power density of the lower frequencies,
f < 0.1N is significantly higher with respect to that obtained at the monitoring point
in the impeller outflow region. Referring to the discussion about the turbulence model-
ing within the RANS framework in the introduction section, this power density spectrum
clearly illustrates the absence of a spectral gap between the MI-induced fluctuations, and
the random fluctuations (consistent with Nikiforaki et al. (2002)).
The power density spectra were analyzed at all monitoring points and the results of the
principal frequencies (i.e. the frequency with the maximum power density) are presented
in Figure 3.7a. These results indicate that around 50% of all monitoring points give the
principal frequency at f = 0.0228N . Also a significant amount of monitoring points
give the principal frequency at f = 6N , which is the blade passage frequency. The
characteristic frequency of f = 0.0228N was found in the simulated time series of the
axial component as well, albeit at less monitoring points (30%, not shown). This may
indicate that the axis of the vortex is essentially oriented in the axial direction.
A similar spectral analysis of the velocity components was performed on the basis of
the simulation at Re = 20, 000 (Case III). The results of the spectral analysis indicate,
in addition to the distinct blade passage frequency, a clear MI-frequency peak at f =
0.0255N .
Nikiforaki et al. (2002) observed in a similar Rushton turbine stirred tank with clear-
ance T /2, a characteristic frequency of f = 0.013N 0.018N in the Reynolds number
range of Re= 16, 000 48, 000. The frequencies we found are encouraging in the sense
that LES proves to be capable of reproducing the precessing vortex phenomenon. The ob-
served deviation from the experimentally found frequencies (which is not that significant
in view of the experimental and numerical accuracy) challenges further improvement of
the subgrid-scale model and/or the numerical settings.

3.4.4 Effect of Reynolds number and impeller off-bottom


clearance
Galetti et al. (2004) confirmed the established linear dependence of the macro-instability
frequency by Nikiforaki et al. (2002) of f = 0.02N , however they found that macro-
instabilities exhibited a different behavior when considering the laminar, transitional and
turbulent Re regions. This lead to different values of f /N . Three regions were defined;
Region 1 (400 < Re < 6, 300), with a single peak at f = 0.106N , Region 2 (6, 300 <
3.4. Results 61

Re < 13, 600), with two peaks at f = 0.106N and f = 0.015N and Region 3 (13, 600 <
Re < 54, 400) with a single peak at f = 0.015N . The latter turbulent frequency peak was
in agreement with the findings of Nikiforaki et al. (2002) and to our work. To enhance the
confidence in the LES method, a case was defined with Re= 12, 500 (i.e. Case II). Similar
to the histogram of Case IV shown in Figure 3.7a, a histogram based on the results of Case
II is presented in Figure 3.7b. The turbulent frequency peak is found not that pronounced
and clearly defined compared to Case IV at Re= 30, 000 (f 0.012N ). However, another
clear peak emerged at f = 0.092N . This frequency peak is in good agreement with the
experimentally obtained laminar frequency peak, f = 0.106N , by Galetti et al. (2004).
Galetti et al. (2004) investigated experimentally the effect of the impeller off-bottom
clearance on the macro-instability frequency. They observed no significant variation in the
macro-instability frequency, and therefore the clearance has no pronounced effect on the
macro-instability, both for the laminar and turbulent regions. However, flow visualization
results (Yianneskis, 2004) indicated that the instabilities were more pronounced and more
clearly defined with the Rushton turbine located at C/T = 0.5. This has been the major
reason for us that we set the impeller off-bottom clearance fixed at C/T = 0.5.

3.4.5 Pseudo turbulence determination procedure


Velocity fluctuations in a turbulently stirred tank are partly periodic and partly random.
The time series presented indicate that, in addition to the random turbulent fluctuations,
coherent oscillations with a low frequency are present in stirred tank flow. While in the
impeller region such oscillations are directly related to the blade passage frequency, in the
bulk flow a precessing vortex motion induces these oscillations. The previous section has
shown that the time scales related to both phenomena differ by more than two orders of
magnitude.
As a result, the kinetic energy can be divided in a coherent part and a random part. The
coherent contribution to the kinetic energy is sometimes presented as pseudo-turbulence
(Van t Riet & Smith, 1975). The kinetic energy in the velocity fluctuations, k, is defined
as follows:

1 2 
k= ui ui 2 = kran + kbpf + kM I (3.5)
2
where kran is the random contribution to the kinetic energy, and kbpf and kM I are the
coherent contributions to the kinetic energy related to the blade passage frequency and the
precessing vortex, respectively. Note the summation convention over the repeated index i.
The (time) averages are over all velocity samples, irrespective of the angular position of
62 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

0.01
bulk
0.008 impeller

kf /vtip
2
() 0.006

0.004

0.002
PSfrag replacements
0
0 0.01 0.02 0.03 0.04 0.05
fco /N ()

Figure 3.8: The kinetic energy, kf of the low-pass filtered time series of the radial veloc-
ity component as function of the low-pass filter cut-off frequency (fco ). The
labels bulk and impeller correspond with positions 2r/T = 0; z/T = 0.12
(Figure 3.5d) and 2r/T = 0.33; z/T = 0.5 (Figure 3.5a), respectively.

the impeller.
The coherent part of the kinetic energy related to the blade passage frequency, can
easily be removed if angle-resolved average data are available:

1 
k kbpf = kran + kM I = < ui2 > < ui 2 > (3.6)
2
with ui the ith resolved velocity component sample at the angular position . Con-
sequently, the time averages are over the velocity samples at the angular position . The
notation < > denotes averaging over all angular positions. Each velocity sample v j at
time tj = jt (or angle j = j = j 2N t) is ascribed to an angular position with
respect to an impeller blade as follows:


= mod(2jN, ) (3.7)
3
With an angular velocity of (1600t)1 exactly 800 velocity samples cover the angular
range (0 ). With Equation 3.7, the angular positions within this range are reflected in
the range (0 3 ), with an angular resolution of 23 N t. A velocity time series of 570
impeller revolutions results in a number of 1140 velocity samples at each angular position,
which is sufficient for meaningful flow statistics.
The above mentioned angle averaging practice is straightforward for the extraction of
the kinetic energy content related to the blade passage frequency, as this frequency and its
phase are fixed and imposed to the flow. The precessing vortex on the other hand is a flow
3.4. Results 63

0.5 0.5
vrad,f /vtip ()

vrad,f /vtip ()
0.25 0.25

0 0

-0.25 PSfrag replacements -0.25


PSfrag replacements N t ()
-0.5 -0.5
0 100 200 300 400 500
N t () P DF (a.u.)

(a) Time series of the radial velocity compo- (b) PDF of the time series shown in (a).
nent after the low-pass filtering procedure.

0.5 0.5
vrad,f /vtip ()

vrad,f /vtip ()

0.25 0.25

0 0

-0.25 PSfrag replacements -0.25


PSfrag replacements N t ()
-0.5 -0.5
0 100 200 300 400 500
N t () P DF (a.u.)

(c) Time series of the radial velocity compo- (d) PDF of the time series shown in (c).
nent after removal of the low-frequency con-
tent.

Figure 3.9: Results of the low-pass filtering procedure on the time series of the radial ve-
locity component shown in Figure 3.5d. vrad,f is the filtered radial velocity
component and vrad,f is the time series of the radial velocity component af-
ter removal of the low-frequency content.

property and although the frequency of its precessional motion is well defined, its motion is
not. The vortex certainly precesses in a clock-wise direction as is indicated in Figures 3.3
and 3.4, but due to the turbulent behavior of the flow, the vortex position fluctuates around
its mean motion. Therefore, the described angle averaging practice cannot be applied for
determining the kinetic energy in the velocity fluctuations related to the precessing vortex
64 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

0.16
bulk
0.12 impeller

k/vtip ()
0.08
2

0.04
PSfrag replacements
0
0 50 100 150 200 250
N t ()

Figure 3.10: The development of the kinetic energy as function of time series record length
(Case IV). The monitoring points at positions z/T = 0.12;2r/T = 0 and
z/T = 0.5;2r/T = 0.367 are labeled bulk and impeller, respectively.

phenomenon.
The kinetic energy content related to the MI can be estimated by means of a filtering
procedure. The difficulty is which frequency range and what amount of power contained
within this range should be accounted to the MI since (as we discussed above) the MI is
not spectrally separated from the rest of the fluctuations. Roussinova et al. (2000) applied
a moving average filter to their measured LDA signals with different time window sizes,
related to a number of blade passages. We used a low-pass filter, and (as a first approxi-
mation) attributed the power content below the cut-off frequency to the coherent velocity
fluctuations related to the precessing vortex.
The influence of the filter bandwidth is shown in Figure 3.8. At the position in the bulk
flow, the kinetic energy content of the filtered signal, kf , sharply increases at a cut-off
frequency close to fM I . This sharp increase is not found in the filtered signal obtained in
the impeller outflow region. Moreover, the filtered kinetic energy content in the bulk of
the tank is found to be much larger than in the impeller outflow region. Henceforth, the
kinetic energy content related to the MI is attributed to the power content in the frequency
range: 0 f 0.0228N . As an illustration of the filtering process, the time series of
the radial velocity component shown in Figure 3.5d is filtered and shown in Figure 3.9a,
with its pdf in 3.9b. Figure 3.9c gives the time series without the low-frequency content
and its pdf is shown in Figure 3.9d. The pdf shows that the velocity fluctuations in the
filtered time series are more or less Gaussian distributed. Therefore, it can be concluded
that the coherent low-frequency content is almost completely removed from the original
time series by the filtering procedure.
3.4. Results 65

k k
0.56 kbpf 0.56 kbpf
z/T ()

z/T ()
0.52 0.52

PSfrag replacements 0.48 PSfrag replacements 0.48

0.44 0.44

0 0.03 0.06 0.09 0.12 0 0.03 0.06 0.09 0.12


2 2
k/vtip () k/vtip ()

(a) 2r/T = 0.367 (b) 2r/T = 0.5

k k
0.56 kbpf 0.56 kbpf
z/T ()

z/T ()

0.52 0.52

PSfrag replacements 0.48 PSfrag replacements 0.48

0.44 0.44

0 0.03 0.06 0.09 0.12 0 0.03 0.06 0.09 0.12


2 2
k/vtip () k/vtip ()

(c) 2r/T = 0.633 (d) 2r/T = 0.767

Figure 3.11: Axial profiles of the kinetic energy at several radial positions (Case IV). A
distinction between kinetic energy and kinetic energy related to random ve-
locity fluctuations is made.

3.4.6 Kinetic energy


The kinetic energy in a large eddy simulation is partly resolved and partly unresolved. The
kinetic energy residing at the subgrid scales can be estimated based on isotropic, local-
equilibrium mixing-length reasoning (Mason & Callen, 1986). A comparison between the
grid scale and subgrid scale kinetic energy (instant realization) in a mid-way baffle plane
resulted in the conclusion that the subgrid scale kinetic energy was at least one order of
magnitude less than the grid scale kinetic energy (Derksen, 2003). Therefore, the subgrid
scale contribution to the total kinetic energy was neglected, and in the following results
66 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

0.02 0.02
k k
kM I kM I
0.015 0.015
()

()
k/vtip

k/vtip
0.01 0.01
2

2
PSfrag replacements PSfrag replacements
0.005 0.005

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
2r/T () 2r/T ()

(a) z/T = 0.12 (b) z/T = 0.36

0.02 0.02
k k
kM I kM I
0.015 0.015
()

()
k/vtip

k/vtip

0.01 0.01
2

PSfrag replacements PSfrag replacements


0.005 0.005

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
2r/T () 2r/T ()

(c) z/T = 0.64 (d) z/T = 0.76

Figure 3.12: Radial profiles of the kinetic energy at several axial positions (Case IV). A
distinction between kinetic energy and MI kinetic energy is made.

the total kinetic energy is the resolved total kinetic energy.


The precessing vortex motion has important implications for the determination of the
kinetic energy contained in the velocity fluctuations. For an accurate prediction of the
kinetic energy, a couple of precessing vortex periods needs to be captured, see Figure 3.10.
The line labeled bulk is obtained from the time series presented in Figure 3.5d. More
than 100 impeller revolutions (i.e. roughly two precessing vortex periods) are needed for
a more or less constant value of the kinetic energy. When the influence of the precessing
vortex decreases (at increasing 2r/T ) fewer impeller revolutions are needed to arrive at a
3.4. Results 67

constant value of the kinetic energy (e.g. for 2r/T = 0.88 around 100 impeller revolutions
are needed). In the impeller outflow region the influence of the precessing vortex on the
value of the kinetic energy is negligible, which is shown by the line labeled impeller in
Figure 3.10. Here, around 10 impeller revolutions are sufficient for an accurate prediction
of the kinetic energy.
Four axial profiles of the total kinetic energy are shown in Figure 3.11. These results
are based on record lengths of the velocity components corresponding to 570 impeller
revolutions. Close to the impeller tip at 2r/T = 0.367, a large amount of the kinetic en-
ergy is found to be related to the coherent blade passage frequency. This blade passage
contribution to the kinetic energy decreases rapidly as the distance from the impeller tip
becomes larger; from 2r/T = 0.633 radially outward this contribution becomes negligi-
ble. The contribution of the precessing vortex to the kinetic energy was found negligible
in the impeller outflow region (less than 1% of the total kinetic energy).
Figure 3.12 shows four typical radial profiles of the total kinetic energy over record
lengths of 570 impeller revolutions. This number of revolutions is sufficient for a constant
value of the kinetic energy. Next to the radial profiles of the kinetic energy, the radial pro-
files of the kinetic energy of the macro-instability related velocity fluctuations are shown.
The coherent blade passage contribution to the kinetic energy was found negligible (less
than 0.1% of the total kinetic energy) at all monitoring points located on the horizon-
tal lines. The radial profiles of the kinetic energy are approximately flat in the region
0.2 < 2r/T < 0.8. Increased k-levels are observed along the tank wall as a result of the
impeller outflow. Also around the tank centerline, increased k levels are found. This effect
is found more pronounced in the domain below the impeller. About the tank centerline
there is a significant contribution to the kinetic energy due to the passage of the precessing
vortex, being 44% of the total kinetic energy at z/T = 0.12, 26% at z/T = 0.36, 14% at
z/T = 0.64 and 21% at z/T = 0.76. These percentages illustrate in the first place that
the MI is more pronounced in absence of the impeller shaft. In the second place, the con-
tribution of the MI to the kinetic energy is most significant in the top and bottom parts of
the tank at 2r/T < 0.2. At radial positions larger than about 2r/T = 0.2, the MI effect on
the turbulence levels is less than 10%. These latter observations imply that the precessing
vortex moves in the area 2r/T < 0.2, which was already suggested by Figures 3.3 and
3.4.

3.4.7 Grid size effects


The results presented were based on simulations performed on the 1803 -grid. It is interest-
ing to study the effect the grid size may have on the (spatial) resolution of the precessing
68 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

25 25

20 20

15 15
% points

% points
10 10

5 5
PSfrag replacements PSfrag replacements
0 0
0.001 0.01 0.1 1 10 0.001 0.01 0.1 1 10
f /N () f /N ()

(a) Case I; Re=20, 000, 1203 -grid (b) Case II; Re=20, 000, 1803 -grid

Figure 3.13: Histograms of maximum frequency peaks found in the power density spectra
of the velocity components.

0.5 0.5
vrad /vtip ()

vrad /vtip ()

0.25 0.25

0 0

-0.25 -0.25
PSfrag replacements PSfrag replacements
-0.5 -0.5
0 50 100 150 0 50 100 150
N t () N t ()

(a) z/T = 0.12; 2r/T = 0.1 (b) z/T = 0.88; 2r/T = 0.1

Figure 3.14: Traces of the radial velocity component at two typical positions in the bulk
of the tank. The two horizontal line segments around N t = 40 are due to
processing errors during the simulation. This simulation was performed on
the fine 2403 -grid at Re=30,000 (Case IV).

vortex motion. Therefore, the stirred tank flow at Re= 20, 000 has been simulated on two
grid sizes; a coarser 1203 -grid and the standard 1803 -grid. The frequencies containing the
maximum power density were captured and put into a histogram, see Figure 3.13. The
3.5. Conclusions 69

histograms show that the blade passage frequency is accurately found at both grid sizes. A
broad range of low frequencies (i.e. f < N ) is found at the 1203 -grid, but a clear dominant
peak due to the MI phenomenon is not observed. Contrary to this, the histogram for the
1803 -grid does show a clear dominant frequency peak at f = 0.0255N .
Next, a second flow simulation at Re=30, 000 has been performed on a finer grid (i.e.
2403 grid nodes, Case V). Because of the higher computational demand, 150 rather than
570 impeller revolutions were simulated. This number of revolutions is not enough to
determine accurately the characteristic frequency and the consequent MI impact on the
fluctuation levels. Nevertheless, the time series of the velocity components indicated the
existence of the precessing vortex motion. As an illustration, the time series of the radial
velocity component is shown for two typical positions; one in the bottom part of the tank
and one in the top part of the tank (in Figure 3.14). They clearly show consistent behavior
compared to the results presented for Case IV in terms of the principal frequency, phase
difference between upper and lower vortices, and the strength of the mean flow variation.
The time series confirm the precessing vortex motion with a characteristic frequency of
f = 0.02N 0.025N (based on two periods). As grid refining did not lead to a better
agreement between the experimental and numerical result of the characteristic frequency,
the conclusion for the time being is that improvement of the subgrid-scale model may lead
to a better correspondence. In view of the simulations performed, a grid of 180 3 nodes is
necessary and sufficient to resolve the precessing vortex motion at the Reynolds numbers
investigated.
From Figures 3.3 and 3.4 the size of the precessing vortex can be estimated. With a
view to the discussion above, it appears that on the 1803 -grid about 10 20 grid cells are
required to accurately resolve the precessing vortex. This leads to a diameter of the vortex
of about T /18 T /9.

3.5 Conclusions
In this paper, we have investigated the precessing vortex phenomenon (i.e. a macro-
instability) numerically in a Rushton turbine stirred tank by means of large eddy simu-
lations. Such an investigation is worthwhile for the determination of the integral time
scale of the flow motion, and the significance of the precessing vortex with respect to the
fluctuation levels and the flow patterns.
The presence of precessing vortices has important implications for numerical simula-
tions of stirred tank flows. The vortex motion is a transient phenomenon which cannot
be captured in a quasi steady-state simulation, and reduction of the computational domain
70 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

because of symmetry properties of the geometry is not possible. We have performed large
eddy simulations instead of RANS simulations, because within the LES methodology the
scales of interest are directly resolved, and there is a clear distinction between the resolved
and unresolved scales. Furthermore, we avoided the interface models commonly applied
for the calculation of stirred tank flows in commercial codes, since it is unclear how signif-
icant the (spatial) resolution of the precessing vortex motion is affected by the location of
the interface between the rotating and stationary meshes, and by the interface model itself.
The sequences of snapshots of the flow field in a horizontal plane below and above
the impeller showed a vortical structure moving around the tank centerline in the same
direction as the impeller. The vortex below the impeller was found more pronounced in
strength and size compared to the vortex above the impeller, and both vortices move with
a mutual phase difference.
The presented time series of the velocity components reveal, besides the random tur-
bulent fluctuations, two types of coherent fluctuations with time scales separated by more
than two orders of magnitude. The fluctuations in the impeller outflow region are dom-
inated by the passage of the impeller blades, whereas fluctuation levels close to the tank
centerline are dominated by the low frequency motion of a vortical structure.
A frequency analysis provided the characteristic frequencies of f = 0.0255N and f =
0.0228N at flow Reynolds numbers of 20, 000 and 30, 000, respectively. The determined
values of the characteristic MI frequency are in good agreement with reported (turbulent)
frequencies found experimentally in a similar flow geometry. At a Reynolds number of
12, 500 a second frequency peak at f = 0.092N was observed, which is consistent with
the experimentally observed frequency, the so-called laminar frequency, in the literature.
Meaningful flow statistics can only be extracted if the flow is calculated in a time
span covering several integral time scales. In the impeller outflow region with 2r/T <
0.63, the flow is dominated by the passage of the impeller blades. A time span covering
several blade passage periods is sufficient here. In the bulk flow region with 2r/T < 0.2,
the flow is dominated by the low-frequency precessing vortex, and at least 100 impeller
revolutions need to be captured for an accurate prediction of the kinetic energy contained in
the velocity fluctuations. By means of a low-pass filtering procedure, it has been observed
that in the top and bottom parts of the tank at 2r/T < 0.2 a significant amount (up to 44%)
of the kinetic energy is related to the precessing vortex.
The stirred tank flow has been simulated using three different lattice node densities.
The main conclusion is that a grid size of 1803 lattice nodes is necessary and sufficient for
resolving the vortex motion at the Reynolds numbers investigated.
LES has proven its significance in the prediction of the precessing vortices present in
a stirred tank flow. The detailed results gave insight in the structure and characteristic
3.5. Conclusions 71

frequency of the vortices, and drew attention to the precessing vortex as a point for im-
provement in the existing modeling techniques. Furthermore, the results form a promising
perspective for the application of the LES methodology in future studies considering the
significance of the precessing vortex in multi phase mixing processes.
72 Chapter 3. Macro-instability uncovered in a Rushton turbine stirred tank

Nomenclature

Roman Description Unit


A+ constant in the Van Driest (1956) wall damping func- -
tion
cs Smagorinsky constant -
D impeller diameter m
f frequency s1
fco low-pass filter cut-off frequency s1
H height of the tank m
k kinetic energy in the velocity fluctuations m2 .s2
kbpf kinetic energy in the velocity fluctuations related to m2 .s2
blade passage frequency
kf kinetic energy in the velocity fluctuations after low- m2 .s2
pass filtering
kM I kinetic energy in the velocity fluctuations related to m2 .s2
the macro-instability
kran kinetic energy in the velocity fluctuations related to m2 .s2
turbulence
N impeller speed s1
r radial coordinate m
S resolved deformation rate s1
t time s
tj discrete time s
T tank diameter m
ui velocity component i m.s1
ui velocity component i linked to an angular position m.s1
u0i fluctuating velocity component i m.s1
vj velocity sample m.s1
vrad radial velocity component m.s1
vrad,f radial velocity component after low-pass filtering m.s1
vrad,f radial velocity component after removal of low- m.s1
frequency content
vtip impeller tip speed m.s1
x, y Cartesian coordinates m
y+ distance from wall in viscous wall units -
3.5. Conclusions 73

z axial coordinate m

Greek Description Unit


lattice spacing m
angle rad
j discrete angle rad
mix mixing length m
kinematic viscosity m2 .s1
e Smagorinsky eddy viscosity m2 .s1
wall-clock time needed for calculating a single im- s
peller revolution
z axial vorticity component rad.s1
0 impeller angular speed of revolution rad.s1
4 Development and validation of a
scalar mixing solver based on
finite volume discretization

4.1 Introduction
In the preceding chapters the focus was on the performance of a numerical tool to im-
prove the understanding of the flow phenomena occurring in stirred tanks. The flow solver
combined with the LES methodology has proven its potential in accurately recovering
the global and local flow features. Furthermore, the turbulence levels were more accu-
rately represented by a LES, compared to a RANS based solver. By means of LES it was
demonstrated that the turbulent fluctuations are partly random and partly coherent. The
latter fluctuations are induced by the passage of impeller blades, and slowly precessing,
large-scale vortices (a so-called macro-instability) about the tank centerline.
The influence of turbulence on the mixing performance and the description of two-
phase processes including mass transfer (e.g. the influence of turbulent velocity fluctua-
tions on the motion of particles and on the local mass transfer rate) is extremely important.
The LES flow solver has demonstrated its capability in accurately recovering the turbulent
scales down to the size of the grid. However, characterizing the mixing performance of a
stirred tank (e.g. in terms of a mixing time) or describing two-phase processes including
inter-phase mass transfer (e.g. particle coating, dissolution processes) requires additional
information on scalar transport in the continuous phase.
In this chapter, the development of a scalar transport solver is described. The solver
is based on the finite volume method, which is an Eulerian approach. Particular attention
is paid to the finite volume schemes for discretizing the various terms in the convection-
diffusion equation. Furthermore, an immersed boundary technique is described for impos-
ing the zero-gradient constraint at off-grid walls. The performance of the solver and the
76 Chapter 4. Development and validation of a scalar mixing solver

newly developed immersed boundary technique are assessed on the basis of two laminar
flow cases and one turbulent flow case.

4.2 Finite volume discretization of the convection


diffusion equation
Consider the convection-diffusion equation for a general scalar
 
ui
+ = + S (4.1)
t xi xi xi
where is the scalar diffusion coefficient and S is the source term. Note that total
mass conservation is part of this class as can be seen by taking =1, =0 and S =0. Finite
volume integration of the above model equation over an arbitrary control volume , and
using the Gauss-theorem leads to


Z Z Z Z
d + ni ui dS = ni dS + S d (4.2)
t S S xi

where S is the surface of the control volume and ni is the ith component of the
normal vector of S. From now on we suppose an uniform, cubic grid (i.e. identical lattice-
spacing in all directions, ) and ignore the source term (S =0). The above equation can
be written as follows

  i+ 21 ,j,k
()i,j,k
V = A FC +
t x1 i 12 ,j,k
  i,j+ 21 ,k

A FC + (4.3)
x2 i,j 21 ,k
  i,j,k+ 21

A FC
x3 i,j,k 1 2

i+ 1 ,j,k
where e.g. (.)|i 21 ,j,k = (.)i+ 21 ,j,k (.)i 12 ,j,k , x1 , x2 , x3 are the spatial coordinates in
2
the directions i, j, k, respectively, V is the volume of cell i, A is the area of the relevant
cell face and FC is a convective prefactor, e.g.
4.3. Time discretization 77

FC;i+ 21 ,j,k = (u1 A)i+ 21 ,j,k (4.4)


where u1 is the velocity normal to the cell face (or the x1 component of the velocity
vector), obtained by linear interpolation of the surrounding nodal values. The diffusive
fluxes are handled by second-order accurate central approximations, e.g.
 
i+1,j,k i,j,k
= (4.5)
x i+ 1 ,j,k
2

The treatment of the convective fluxes and the time discretization are the subjects of
the next two sections.

4.3 Time discretization


The finite volume integration of equation 4.3 over a control volume must be augmented
with a further integration over a finite time step t. Considering a first-order accurate
Euler forward time integration, the 1-D version of equation 4.3 reads

t+t   i+ 12

Z
V ()n+1 ()ni =

i A FC dt (4.6)
t x i 1
2

where n refers to the scalar at time t and n+1 refers to the scalar at time t + t. To
evaluate the right-hand side of the above equation we need to make an assumption about
the variation of with time. We could use the scalar at time t or at time t + t to calculate
the integral or, alternatively, a combination of the scalar values at time t and t + t. We
may generalize the approach by means of a weighting parameter between 0 and 1 and
write the integral of scalar with respect to time as
Z t+t  
dt = ()n+1 + (1 )()n t (4.7)
t

The exact form of the eventual discretized equation depends on the value of . When
is zero, only scalar values at the old time level t are used at the right-hand side of equation
4.6 to evaluate at the new time; resulting in an explicit scheme. When 0 < 1
scalar values at the new time level are used on both sides of the equation; the resulting
schemes are called implicit. The extreme case of = 1 is termed fully implicit and the
case corresponding to = 0.5 is called the Crank-Nicolson scheme. The latter scheme is
second-order accurate in time.
78 Chapter 4. Development and validation of a scalar mixing solver

The convection-diffusion equation is solved on the LES grid, and the time step is the
same as the LES time step. The LES time step was found to be sufficiently small in
order to satisfy the Courant stability criterion. Therefore, an explicit scheme for the scalar
calculations will be considered in the following discussions. The explicit scheme has a
first-order Taylor series truncation error accuracy with respect to time. This scheme can
be made second-order accurate in time by means of the Adams-Bashford time integration
(i.e. linear interpolating polynomial):

t
()n+1
i = ()ni + (3Qni Qin1 ) (4.8)
2
where Qi is
  i+ 12
1
Qi = V A FC (4.9)
x i 12

4.4 The treatment of convection


In this section special attention is paid to various existing convection schemes. Issues of
importance are stability aspects and the way to minimize numerical diffusion.

4.4.1 Monotone schemes


Consider the one-dimensional scalar convection-diffusion equation in absence of a source
term
 

+ u =0 (4.10)
t x x
Any linear general 3-point discretization can be written in the following way

ai n+1
i = bi n+1 n+1 n
i+1 + ci i1 + i (4.11)

where ai = bi +ci +1 for consistency. This scheme is called positive or monotone when
its coefficients bi , ci are positive. Positivity implies that whenever the initial conditions
are positive, the solution will remain positive during time stepping. It also means that
no spurious wiggles will occur. A non-monotone scheme may allow wiggles to occur in
regions where steeps gradients exist. When the grid density is sufficiently fine to accurately
resolve these gradients, the size of these wiggles may not be very disturbing in most cases.
4.4. The treatment of convection 79

Mathematical models that solve turbulent flow, multi-phase flow or scalar transport
contain, however, a number of variables that are inherently positive, e.g. the turbulent
kinetic energy and dissipation, phase fractions, concentration, etc. Oscillations in the so-
lution of these equations cannot be tolerated, for they could result in negative values of
these variables, which in turn may result in divergence of the algorithm. Therefore, some
form of monotone discretization is essential for the solution of these physically positive
variables. Small oscillations may slow down or prohibit convergence in nonlinear systems
and the usage of monotone schemes can be of help to overcome this problem.
Many convection schemes have been proposed in the past. The first-order upwind
scheme (Patankar, 1980)
(
i , ui+ 12 0
i+ 21 = (4.12)
i+1 , ui+ 21 < 0

for the evaluation of the face value has long been very popular. This scheme is un-
conditionally positive and the implicit discretization leads to a linear system with favor-
able solution properties (diagonally dominant). Furthermore, it takes into account flow
direction when determining the value at a cell face, which is very important in strongly
convecting flows. On the other hand, the scheme is first-order accurate only, which is in-
sufficient for most practical applications. On top of this, it is notorious for its numerical
diffusion (or false diffusion), especially when the flow is skewed with respect to the mesh.
Consequently, the upwind scheme is inapplicable for convection dominated (high Schmidt
number) flows. In spite of the deficiencies, it is still widely applied in present-day CFD.
The central difference scheme, also widely used, is based on a linear interpolation of
the surrounding grid nodes

1
i+ 21 = (i+1 + i ) (4.13)
2
Although this scheme has second-order accuracy, it fails in identifying flow direction.
Furthermore, this scheme is positive under the condition that the cell Peclet number de-
fined as

ui+ 12
Pei+ 12 = (4.14)
i+ 21

does not exceed the value of 2.


Combinations of the above schemes have been proposed, such as the hybrid and pow-
erlaw variants (Patankar, 1980). These schemes behave as the central difference scheme at
80 Chapter 4. Development and validation of a scalar mixing solver

low Peclet numbers and switch to first-order upwinding for higher values. These variants
are unconditionally positive, but share the property of numerical diffusion when the flow
is convection dominated.
As mentioned above, second-order accuracy is desirable, especially when the problem
is convection dominated. Moreover, Godunov (1959) has shown that linear monotone
schemes can have an order accuracy of at most 1 (order barrier). This barrier does not
apply to nonlinear discretizations. Higher-order upwind schemes have been developed
in the past, such as the second-order upwind scheme and the QUICK scheme (Leonard,
1979). These schemes are, however, only conditionally positive. More recently, there have
been strong developments on nonlinear discretizations in the area of compressible flows.
These techniques are gaining attention in sciences outside the aeronautics related fields.

4.4.2 TVD discretization scheme


For the last two decades there has been a sustained effort by the CFD community to de-
velop robust high resolution schemes for the simulation of convection dominated flows.
The main ingredients common to these schemes are a higher-order profile for the recon-
struction of the cell face values from cell averages, combined with a monotonicity crite-
rion. To satisfy monotonicity, a number of concepts have been proposed over the years.
In the flux corrected transport (FCT) approach (Zalesak, 1979; Boris & Book, 1973), a
first-order accurate monotone scheme is converted to a high resolution scheme by adding
limiting amounts of anti-diffusive flux. In the monotonous upstream-centered scheme for
conservation laws (MUSCL) of Van Leer (1979), monotonicity is enforced through a lim-
iter function applied to a piecewise polynomial flux reconstruction procedure. The concept
of TVD schemes was introduced by Harten (1983). For certain types of equations, these
algorithms can ensure that the total variation over the whole computational domain does
not increase with time (i.e. total variation diminishing, TVD); thus no spurious numerical
oscillations are generated:
X X n
n+1 n+1 i+1 ni

i+1 i (4.15)
i i

Since by numerical schemes only the value of the cell average is available, with the
concept of TVD the cells are reconstructed in the way that no spurious oscillation is present
near a discontinuity or a zone with steep gradients and high-order accuracy is retained
simultaneously.
In the following discussion, we will consider ui+ 12 > 0 without loss of generality.
The usual methodology to define a second-order TVD discretization scheme is through
4.4. The treatment of convection 81

the introduction of a so-called flux-limiting function controlling the amount of anti-


diffusion on a first-order upwind approximation (Roe, 1981)

1  
i+ 12 = i + ri+ 12 (i+1 i ) (4.16)
2
As wiggles originate from steep changes in gradients of the scalar , it is natural to let
the flux-limiter depend on the local ratio of gradients ( ri+ 21 ):

i i1
ri+ 21 = (4.17)
i+1 i
By using a flux limiter, different first-order and second-order schemes can be written in
the form of equation 4.16. For example, for equal to zero, the first-order upwind scheme
is obtained. Other schemes can be similarly formulated:

First-order downwind scheme: (r) = 2

Second-order central difference scheme: (r) = 1

Second-order upwind scheme: (r) = r


3+r
Second-order QUICK scheme: (r) = 4

Following Sweby (1983), these schemes may be plotted along with the TVD mono-
tonicity region on an r diagram (see Fig. 4.1). Using this diagram, it is easy to
understand the formulation of TVD schemes: any flux-limiter function, formulated to lie
within the TVD monotonicity region yields a TVD scheme. Sweby (1983) has also shown
that to get second-order accuracy in smooth regions, the condition (1) = 1 should be
met. The Superbee flux-limiter, formulated as
h i
(r) = max 0, min[2r, 1], min[r, 2] (4.18)

obeys the two previously mentioned criteria. Wang & Hutter (2001) have used the
Superbee limiter in their work and reported that this limiter is most favorable in treating
convectively dominated problems. Therefore, we have used this limiter in our work for
simulating the scalar transport in high Schmidt number systems.
Usually one restricts to be positive for r 0. In addition, it is set to zero for r < 0
(e.g. the Superbee limiter). Thus, at extrema, the discretization reduces to first-order
upwinding. By fulfilling certain additional restrictions, uniformly accurate limiters can be
constructed which do not degrade to first-order accuracy near extrema (Zijlema, 1996).
PSfrag replacements
82 Chapter 4. Development and validation of a scalar mixing solver

(r)
DW SOU
2
QUICK
1 CD
SUPERBEE

0 UW
1 r

Figure 4.1: TVD monotonicity regions (gray area) on the Sweby (1983) diagram. The
flux-limiters of the upwind scheme (UW), downwind scheme (DW), central
difference scheme (CD), QUICK scheme, second-order upwind scheme (SOU)
and TVD scheme (Superbee) are drawn in the diagram.

We will not consider limiters of this kind in this thesis, as their additional effect was found
to be minimal (Lathouwers, 1999).
TVD schemes can also be formulated in a form slightly different from the one given in
equation 4.16, as in:

1  0 
i+ 21 = i + 0 ri+ 1 (i i1 ) (4.19)
2 2

0 1
where ri+ 1 = r
i+ 1
. The relation between the two formulations is given by
2 2

1    
0 ri+
0
1 = ri+ 21 (4.20)
ri+
0
1 2
2

Taking for example the second-order upwind scheme where (r) = r, the equivalent
flux-limiter function becomes 0 (r0 ) = r 0 r = 1. It can be shown that the Superbee flux-
limiter is the same for both formulations.

4.4.3 TVD multidimensional application


Almost all theory about TVD schemes has been developed for 1-D flow. According to
Goodman & LeVeque (1985), a straightforward extension of TVD definitions to multi-
ple dimensions implies that a conservative TVD scheme is first-order accurate at most.
Clearly, a less strict criterion is needed for monotonicity than TVD in multiple dimen-
sions. On the other hand, experience is that plain extensions of the 1-D TVD concepts
4.5. The fully discretized convection diffusion equation 83

work well and lead to accurate oscillation-free solutions in multiple dimensions (Hirsch,
1990).
The extension of equation 4.16 in three dimensions and considering arbitrary convec-
tion speeds reads

 
+
i,j,k + 21 ri+
1 (i+1,j,k i,j,k ) , ui+ 12 ,j,k 0
2 ,j,k
i+ 21 ,j,k =   (4.21)
i+1,j,k 1 r 1

(i+1,j,k i,j,k ) , ui+ 12 ,j,k < 0
2 i+ ,j,k
2

where the limiter is controlled by the upwind and local gradients of the solution:

+ i i1 i+2 i+1
ri+ 1 = ; ri+ 1 = (4.22)
2 i+1 i 2 i+1 i
Other face values are approximated by similar expressions, found by shifting and trans-
posing indices.

4.5 The fully discretized convection diffusion


equation
The starting point for the derivation of the fully discretized convection-diffusion equation
is equation 4.3. For simplicity reasons, we only consider the 1-D version of this equation.
Extensions to multiple dimensions is straightforward. The 1-D version of equation 4.3,
after handling the diffusive fluxes by second-order central approximations (equation 4.5),
reads

d (i i )
V = (FC )i+ 12 + (FC )i 12
dt (4.23)
+ FD;i+ 21 (i+1 i ) FD;i 12 (i i1 )

where FD is a diffusive prefactor, e.g.


 
A
FD;i+ 12 = (4.24)
i+ 21

If we now assume first-order Euler forward time integration (equation 4.6) and consider
an explicit formulation ( = 0 in equation 4.7) equation 4.23 reads
84 Chapter 4. Development and validation of a scalar mixing solver

V
(i i )n+1 (i i )n = (FC )ni+ 1 + (FC )ni 1

t 2 2
n n n (4.25)
+ (FD )i+ 1 (i+1 i )
2
n
(FD )i 1 (ni ni1 )
2

Finally the convective fluxes are approximated by means of TVD discretization with
the Superbee flux-limiter. After some rearrangements the fully discretized convection dif-
fusion equation is given by:

(bi i )n+1 = (ai i )n + (ai1 i1 )n + (ai+1 i+1 )n + cni (4.26)

where the coefficients bi ,ai ,ai1 ,ai+1 are given by:

V
bi = i (4.27)
t h i
ai1 = FD;i 21 + max FC;i 21 , 0 (4.28)
h i
ai+1 = FD;i+ 12 + max FC;i+ 12 , 0 (4.29)
 
ai = bi ai1 ai+1 FC;i+ 21 FC;i 21 (4.30)

The coefficient ci in equation 4.26 gives the amount of anti diffusion given by the TVD
scheme as

1 h i 
+
 h i 


ci = max FC;i+ 12 , 0 ri+ 1 max FC;i+ 12 , 0 ri+ 1 (i+1 i )+
2 2 2

1  h i 
+
 h i 


max FC;i 21 , 0 ri 1 max FC;i 21 , 0 ri 1 (i i1 )
2 2 2

(4.31)

If we set (r) = 0 (upwind flux limiter), equation 4.26 reduces to a first-order accurate
(spatial upwinding and temporal Euler forward) explicit finite volume formulation of the
convection diffusion equation.
If we consider second-order Adams-Bashford time integration, the scalar update is
through equation 4.8. The fully discretized Qi reads
4.6. Boundary conditions 85

Qi = V 1 (ai i + ai1 i1 + ai+1 i+1 + ci ) (4.32)

where ai1 and ai+1 are given by equations 4.28 and 4.29, respectively, and ai is given
by
 
ai = ai1 ai+1 FC;i+ 21 FC;i 12 (4.33)

4.6 Boundary conditions


The equations described in the previous sections require boundary conditions on the com-
plete boundary of the domain. Several boundary types may be distinguished. In the first
place, inflow boundaries where e.g. mass fluxes, phase fractions, and scalar concentration
need to be specified. If available, these values are obtained from interpolation of experi-
mental data. Specification of scalar variables at inflow boundaries in a numerical method
is done via a Dirichlet boundary condition:

= I (4.34)

where I is the scalar value at the inflow boundary. On the other hand, at outflow
boundaries usually no information about the local flow variable is available at all. In this
case, zero-gradient conditions (i.e. a Neumann boundary condition) for scalars are usually
adopted, assuming fully developed flow:


=0 (4.35)
n
were n is the outward pointing normal. The Neumann boundary condition may be
applied for scalars at solid walls as well, since there is no scalar transport through these
walls.
In a Cartesian grid, complex boundaries generally lie off-grid. Various approaches
have been proposed in the past to model these boundaries accurately (Tucker & Pan, 1999;
Calhoun & LeVeque, 2000). With the cut cell technique, the cells are cut by the boundary
leaving fragments with a volume smaller than the control volume. In the explicit formula-
tion of the convection diffusion equation, the time step scales with the volume of the cell.
In order to guarantee stability of the scheme, an infinitesimal time step is required for an
infinitesimal control volume. Therefore, the cut cell approach is not appropriate here for
modeling (moving) boundaries that are not aligned with the mesh.
86 Chapter 4. Development and validation of a scalar mixing solver

A simple approach for modeling off-grid boundaries is by approximating the boundary


with staircases. The boundary technique based on the staircase approximation is applica-
ble in the explicit finite volume discretization, and it is easy to implement. However, at a
coarse grid resolution with respect to the dimension of the boundary, the staircase bound-
ary is an inaccurate representation of the real boundary. Furthermore, the flow is solved in
a lattice-Boltzmann framework, in which the Dirichlet boundary conditions for the veloc-
ity components at the off-grid boundaries are imposed by an immersed boundary technique
(i.e. an adaptive force field technique developed by Derksen & Van den Akker (1999)).
As a result, by making use of the staircase technique to impose the Neumann boundary
constraint at the off-grid boundaries in the finite volume framework, mass conservation is
not guaranteed.
In line with the immersed boundary technique in the lattice-Boltzmann framework to
impose the Dirichlet boundary conditions for the velocity components at the off-grid walls,
we propose an immersed boundary technique in the finite volume framework to impose a
Neumann boundary condition for the scalar. This technique makes use of ghost cells, and
is consequently referred to as the ghost cell technique. In the next subsections, the ghost
cell technique is explained.

4.6.1 Neumann boundary condition implementation in 1-D


Consider three cells in a 1-D domain (see Figure 4.2a). The spacing between the cell
nodes is . The cell with index i = 1 is the ghost cell, the cells with indices i = 0, 1
are flow cells. The physical wall is in between the ghost node and the first flow node
( xw < 0). We consider a zero-gradient constraint (Neumann condition) at the wall.
Via second-order interpolation the scalar value at the ghost node reads
1

2x0 0 x0w + 1
1 = w 2
(4.36)
x0w 12
where x0w = xw /.
In the special case where the boundary is positioned halfway between two grid nodes
(xw = 0.5), the ghost cell value equals

1 = 0 (4.37)

Here, we only need to copy the near wall scalar values to the ghost nodes (in this case
so-called mirror nodes) to satisfy the zero-gradient constraint. As a result, representing a
physical boundary located halfway two grid nodes by means of mirror nodes results in a
4.6. Boundary conditions 87

1 d
d dn =0 1
dn =0
PSfrag replacements 1 PSfrag replacements
0 0

1
-1 xw 0 1

(a) (b)

Figure 4.2: Neumann boundary condition implementation in 1-D (a). Schematic represen-
tation of the determination of the scalar ghost cell value in 2-D (b).

second-order approximation of the Neumann constraint.

4.6.2 Extension to multiple dimensions


For the determination of the scalar ghost cell value (resulting from a Neumann constraint)
at walls that are skewed with respect to the mesh we consider for simplicity a 2-D prob-
lem. Extension to three dimensions is straightforward. Figure 4.2b shows a curved wall
immersed on a Cartesian grid. The gray cells represent the ghost cells as their cell cen-
ters lie outside the flow domain. Black cells are exterior cells as they do not occupy flow
domain.
We reduce the determination of the scalar ghost cell values in a multi dimensional
problem to a one dimensional problem described in the previous section. For that purpose
we determine the positions of the scalar values 0 and 1 through a fictitious line perpen-
dicular to the wall. The spacing between 1 , 0 and 1 is chosen equal to the lattice
spacing. The procedure for the determination of the scalar ghost cell value 1 is shown in
Figure 4.2b. The scalar values 0 and 1 are determined by bi-linear interpolation of the
surrounding cell nodes. For the determination of 0 at least one of the surrounding nodes
is the ghost node. We can either use bi-linear interpolation by taking the old values of the
surrounding nodes (explicit) or solve the scalar ghost cell value implicitly. Test calcula-
tions showed no significant change by solving the scalar ghost node value implicitly, and
therefore the scalar ghost cell value determination was treated explicitly.
88 Chapter 4. Development and validation of a scalar mixing solver

For the scalar concentration, the Neumann boundary condition is imposed at the walls.
In this case, equation 4.36 is used for the determination of the scalar concentration ghost
cell value. This method is unconditionally stable as 1 x0w < 0.

4.7 Validations
With the help of three benchmark cases, the performance of the scalar mixing solver and
the novel immersed boundary technique will be assessed. These benchmark cases are
laminar cavity flow in a square box, laminar cavity flow in a box with inclined side walls,
and turbulent flow in a cylindrical tank with a side-entry mixer.

4.7.1 Laminar cavity flow in a square box


In this section, we study the mixing behavior of two miscible liquids having the same
(Newtonian) viscosity in a cavity flow geometry. The focus is on the performance of
the different discretization schemes with respect to numerical diffusion. The reference
geometry studied here is based on a mixing experiment reported by Hoefsloot et al. (2000).
The experiments were performed in a rectangular cavity of plexiglass. The experimental
set-up with the relevant dimensions is shown in Figure 4.3a. In the experiments the two
liquids have different rheological parameters. The rheology of the dye solution is similar
to that of a Newtonian liquid, whereas the polyacrylamide in glycerin with 2w% water
solution behaves like a non-Newtonian liquid. Based on the kinematic viscosity of the dye
solution (4.78104 m2 /s), the Reynolds number (Re=Ulid L/, where Ulid is the velocity
of the band and L is the width/height of the cavity) equals 7.5. In addition to experiments,
Hoefsloot et al. (2000) report results obtained by simulations with Fluent 5.1.1. These
simulations are based on the finite volume formulation of the Navier-Stokes equations and
a Volume Of Fluid approach (VOF) for the location of the interface. Both experiments and
simulations started with a zero-velocity field. More details on the simulation procedure
are reported in the paper of Hoefsloot et al. (2000).
The cavity flow was simulated with our code in 2-D, with 256 lattice spacings in the x-
and y-directions. A no-slip boundary condition for the flow and a zero-gradient boundary
condition for the scalar concentration have been imposed at the walls. Since the walls are
aligned with the grid, the no-slip boundary condition for the flow has been applied with
a second-order, half-way bounce-back scheme. The zero-gradient boundary condition has
been applied via mirror nodes (second-order accurate, see equations 4.36 and 4.37). The
velocity of the lid and the viscosity of the fluid have been set such, that the Reynolds
4.7. Validations 89

band Ulid = 0.036 m/s


0.08 m Ulid

symmetry
L = 0.1 m plane L

y z 0.15 m
x
L = 0.1 m L

(a) (b)

Figure 4.3: The experimental set-up, a 3-D cavity (a) and the symmetry plane (b).

number equals 10. At the start of the simulations, fluid 1 (containing a zero scalar con-
centration) and fluid 2 (containing a non-zero scalar concentration) both occupy 50%
of the cavity volume. Fluid 1 (white color) is located on top of fluid 2 (black color),
see Figure 4.3b. Contrary to the experiments and simulations reported by Hoefsloot et al.
(2000), the mixing calculations were started in the fully developed flow field and both liq-
uids have the same kinematic viscosity. In order to test the performance of the different
discretization schemes, the diffusion coefficient is set to zero (i.e. the Schmidt number
equals infinity; Sc= ).
Figure 4.4 shows snapshots of the concentration in the cavity. The flow rotation is
clock-wise as the band is moving from the left to the right. As a result, fluid 1 is lifted
upward at the left side of the cavity. Figure 4.4a shows the experimental mixing pattern
after 10 seconds (Hoefsloot et al., 2000), and Figure 4.4b is their simulated result (Hoef-
sloot et al., 2000). The agreement between the experiment and simulation is good. Due
to a bad illumination in the left top corner, the small white region observed in Figure 4.4b
cannot be observed in the experimental result.
Figures 4.4c, 4.4d and 4.4e show the mixing pattern after 8.5 seconds obtained with our
simulations with three different discretization schemes. Although the Reynolds number is
somewhat higher (10 instead of 7.5) and the simulations were started in a fully developed
flow field, the results look qualitatively the same compared to Figures 4.4a and 4.4b. The
white region at the top left corner is accurately resolved. The present simulations are a
little ahead compared to the results of Hoefsloot et al. (2000). Furthermore, the black
tongue observed at the top of the cavity is thicker.
90 Chapter 4. Development and validation of a scalar mixing solver

(a) (b)

(c) (d) (e)

Figure 4.4: The mixing pattern in the symmetry plane after 10 seconds: Hoefsloot et al.
(2000) experiment (a), Hoefsloot et al. (2000) simulation. The mixing pat-
tern in the symmetry plane after 8.5 seconds: Power-law scheme (c), QUICK
scheme (d) and TVD scheme (e). The dashed arrows correspond with the pro-
files in Figure 4.5.

The differences mentioned may be caused by the following discrepancies between


the settings in the present simulations, and the experiment/simulation of Hoefsloot et al.
(2000). In the first place, the present simulations were started in a fully developed flow
field, whereas the experiment/simulation of Hoefsloot et al. (2000) was started in a zero-
velocity field. Secondly, the Reynolds number in the present simulation is somewhat
4.7. Validations 91

2.4 2.4 2.4

2 2 2

1.6 1.6 1.6

1.2 1.2 1.2


c/c (-)
c/c (-)

c/c (-)
0.8 0.8 0.8

0.4 0.4 0.4

0 0 0
-0.4 -0.4 -0.4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
y/L (-) y/L (-) y/L (-)

(a) (b) (c)

Figure 4.5: Profiles of the concentration (normalized with the final concentration c ) after
8.5 seconds at x/L = 0.5 (i.e dashed arrows in Figure 4.4). Power-law scheme
(a), QUICK scheme (b) and TVD scheme (c).

higher compared to that in the experiment/simulation of Hoefsloot et al. (2000). Thirdly,


in the present simulations the rheology of the fluids is the same, which is not the case in
the experiment/simulation of Hoefsloot et al. (2000).
Figure 4.4c shows the mixing pattern after 8.5 seconds obtained with the power-law
discretization scheme. Although the diffusion coefficient is set to zero, the interface is
smeared out. Since the streamlines are not aligned with the Cartesian mesh, the upwind
values of the concentration are not accurately obtained. As a result, the interface smears
out due to numerical diffusion.
Figures 4.4d and 4.4e show the mixing pattern under the same conditions obtained with
the QUICK and TVD schemes. These higher-order schemes significantly reduce the effect
of numerical diffusion; the interface remains sharp.
Figure 4.5 shows profiles of the concentration at position x/L = 0.5, which corre-
sponds with the dashed arrows in Figures 4.4c, 4.4d and 4.4e (the direction of the arrow is
from y/L = 0 to y/L = 1). The profile shown in Figure 4.5a shows a smooth transition
between low and high scalar concentrations, which is attributed to the numerical diffusion
associated with the power-law scheme.
We have seen in Section 4.4.2 that the QUICK scheme is conditionally stable (the
QUICK flux-limiter is not located solely in the TVD monotonicity region, see Figure 4.1),
92 Chapter 4. Development and validation of a scalar mixing solver

Ulid

0
76

Figure 4.6: Cross-section of the inclined cavity at 76o with respect to the horizontal axis.

whereas the TVD scheme is unconditionally stable. This is clearly demonstrated in Fig-
ures 4.5b and 4.5c. The profile obtained by the QUICK scheme shows wiggles and neg-
ative values of the concentration at the extrema, whereas in the profile obtained by the
TVD scheme only sharp step changes (no wiggles) are observed. As a result, the TVD
scheme is favorable compared to the QUICK and power-law schemes and it is used for the
simulations in the following chapters.
Since the mass change in the simulations is smaller than 0.1%, the finite volume
scheme may be regarded as mass conservative.

4.7.2 Laminar cavity flow in a box with inclined side walls


In this section, the performance of the developed immersed boundary technique is assessed
against the performance of the staircase technique. In this case, we again study the lid
driven cavity flow. The side walls of the cavity are inclined at 76o with respect to the
horizontal axis, see Figure 4.6. Fluid 1 (represented by the white color) contains a zero
scalar concentration, and fluid 2 (represented by the black color) contains a non-zero
scalar concentration. Both fluids occupy 50% of the cavity volume, and fluid 1 is located
on top of fluid 2.
The cavity flow was simulated in 2-D. A no-slip boundary condition for the flow and
a zero-gradient boundary condition for the scalar concentration have been imposed at the
walls. The no-slip boundary constraint at the top and bottom walls are imposed by the half-
way bounce-back boundary condition, and the side walls are modeled with the immersed
boundary technique proposed by Derksen & Van den Akker (1999). The zero-gradient
constraint for the concentration is enforced by means of the newly developed immersed
boundary technique (see section 4.6).
4.7. Validations 93

1.10 1.10
L/ 25 L/ 25
1.08 L / 50 1.08 L / 50
L/ 100 L/ 100
1.06 L / 200 1.06 L / 200
M/M0 (-)

M/M0 (-)
1.04 1.04
1.02 1.02

1 1

0.98 0.98
0 10 20 30 40 0 10 20 30 40
(Ulid /L)t (-) (Ulid /L)t (-)

(a) Staircase technique. (b) Ghost cell technique.

Figure 4.7: Simulated time traces of the total mass in the system normalized with the total
mass at t = 0 (M0 ). The lines represent simulations at four different grid
resolutions.

The velocity of the lid and the viscosity of the fluid have been set such, that the
Reynolds number equals 10. The scalar mixing in the inclined cavity has been simu-
lated at four different grid resolutions in order to check its effect on the mass change and
the near-wall scalar concentration distribution. The different grid resolutions adopted, are
L/25, L/50, L/100 and L/200, respectively. The mixing calculations were started in the
fully developed flow field. The Schmidt number equals 1000, which is a typical value in
liquid systems.
Figure 4.7 shows simulated time series of the normalized total mass at four grid resolu-
tions. Both the staircase and ghost cell techniques (Figures 4.7a and 4.7b, respectively) do
not conserve mass. With increasing grid resolution (from L/25 to L/200) the total (posi-
tive) mass change decreases (from 9% to 4% at (Ulid /L)t = 40). Based on the time series
of the mass change, the ghost cell technique performs similar to the staircase technique.
In Figure 4.8, snapshots of the scalar concentration field are shown for the coarse
(L/25) and fine (L/200) grid resolutions at (Ulid /L)t = 5. The walls represented by
staircases and by the ghost cell technique are shown by means of the solid lines. With
the staircase technique used at the coarse grid (Figure 4.8a), the zero-gradient constraint
is not satisfied at the side walls. Application of the ghost cell technique (Figure 4.7b)
results in a more accurate representation of the zero-gradient constraint at the side walls
(the concentration iso-lines tend to end perpendicular to the inclined side walls). Both
techniques, however, have a problem in recovering the low concentration region in the top
94 Chapter 4. Development and validation of a scalar mixing solver

(a) Resolution L/25, staircase technique. (b) Resolution L/25, ghost cell technique.

(c) Resolution L/200, staircase technique. (d) Resolution L/200, ghost cell technique.

Figure 4.8: Instantaneous realizations of the scalar concentration distribution at


(Ulid /L)t = 5. The walls are represented with the solid lines.

left corner of the cavity due to a lack of resolution.


At a fine grid resolution (see Figures 4.8c and 4.8d), both techniques perform similar. A
close inspection of the concentration iso-lines at the right wall reveals the level of accuracy
of the zero-gradient boundary constraint. This constraint is more accurately recovered by
means of the ghost cell technique compared to the staircase technique.
In conclusion, the ghost cell technique does not guarantee mass conservation. The
4.7. Validations 95

change in mass with the ghost cell technique is comparable to that of the staircase tech-
nique for the grid resolutions tested. Grid refinement is a remedy to resolve this problem,
but this implies higher computational effort. With the ghost cell method, the zero-gradient
constraint is applied at a number of points on the boundary which scales with the grid
resolution. A coarse grid resolution implies a lower number of points on the boundary
compared to the number at a high grid resolution, and consequently a less accurate zero-
gradient boundary constraint at the wall. Nevertheless, in contrast to the staircase method,
the ghost cell method takes into account the shape of the boundary regardless of the grid
resolution.
With a view to future calculations in a stirred tank geometry, it is expected that the
tanks outer wall will be sufficiently accurately represented (in terms of mass conservation
and accuracy of the zero-gradient constraint) by the ghost cell method, as, with respect
to the tanks outer wall, the grid resolution is comparable to L/200. With respect to the
impeller blade length, the grid resolution is comparable to L/25. As a result, the perfor-
mance of the ghost cell method with respect to an impeller blade (not to speak about the
blade edges) may be comparable to the performance observed in the results of the above
simulation at grid resolution L/25. The change of mass should therefore be critically
checked in future calculations. The performance of the ghost cell method in a cylindrical
tank geometry is checked in the next section.

4.7.3 Mixing behavior in a cylindrical tank with a side-entry


mixer
The third validation case focuses on the turbulent flow in a cylindrical tank with a side-
entry mixer. Side-entry mixers are often used to mix the contents of large cylindrical
vertical storage tanks. The mixing time of these tanks is large as the diameter of the
impeller is two orders of magnitude smaller than the tank diameter. This validation case is
based on the geometry presented in the paper by Van Looy et al. (2004). A typical layout
of the storage tank is shown in Figure 4.9.
An extensive experimental program was carried out at Shell Research in 1969-1970 in
order to determine the mixing time in storage tanks. Based on the experimental results, an
empirical correlation was derived that can be used to calculate the mixing time for various
tank sizes and density differences of the two miscible liquids. Recently, CFD has been
employed to model this problem (Van Looy et al., 2004).
In this section, results of the mixing performance of the turbulent flow generated by a
side-entry mixer in a storage tank are presented, based on an LES including passive scalar
transport. This validation case is of help to address issues such as the implementation of
96 Chapter 4. Development and validation of a scalar mixing solver

8
o
D = 0.053T

fluid 1
0.0417T
D D fluid 2
T T

(a) Top view. (b) Cross-section.

Figure 4.9: Schematic representation of the storage tank with a side-entry mixer. T and D
are the tank and impeller diameters, respectively.

subgrid-scale diffusion and assessment of mass conservation in the finite volume scheme
(the tanks outer wall is modeled with the newly developed ghost cell technique). Further-
more, the performance of the scalar mixing solver is compared to a LES using Fluent.
The Schmidt number in liquid systems is of the order of 1000. As a result, the Batch-
elor scale (i.e. the smallest length scale of the scalar field) is a factor 30 smaller than the
Kolmogorov scale. As a result, the effect of the scalar subgrid-scale fluctuations on the
resolved-scale scalar field may be larger than that of the subgrid-scale velocity fluctuations
on the resolved-scale velocity field. From this perspective, it seems reasonable to take the
turbulent Schmidt number (Sct ; relating the eddy diffusion coefficient (e ) to the eddy
viscosity via e = e /Sct ) to be smaller than unity. We have chosen a turbulent Schmidt
number of 0.7. A total diffusion coefficient ( = mol + e , with mol the molecular dif-
fusion coefficient) is calculated and applied in the diffusive prefactor (see equation 4.24).
Fluid 1 contains a zero scalar concentration, and fluid 2 contains a non-zero scalar
concentration. Both fluids occupy 50% of the tank volume, and fluid 1 is located on top
of fluid 2 (see Figure 4.9b). As buoyancy effects are not included in the finite volume
N 2D
scheme yet, a simulation has been carried out at infinite Froude number (Fr= g , with
g the magnitude of the gravitational acceleration vector and the density difference
which equals zero).
The simulation starts with a zero-velocity field. For the LES, the time step needed
can be estimated by realizing that the large scale structures of the impeller jet have to be
resolved. The impeller stream velocity Ujet is typically 5 m/s and the impeller diameter
0.125 m. At the very least, the temporal resolution should therefore be D/Ujet = 0.025 s.
4.7. Validations 97

With an impeller speed (N ) of 50 s1 , the dimensionless temporal resolution equals 1.25.


In the LES, the dimensionless temporal resolution is set at 0.01. The Reynolds number,
defined as N D 2 /, is equal to 8.75 105 .
The impeller is modeled as a swirling jet in the simulation. The profiles of the tangen-
tial and axial velocity components (vtan and vax , respectively) inside a cylindrical region
with diameter and length equal to D read

vtan = 1.02N rj ; 0 rj D/2 (4.38)

vax = 0.77N D (4.39)


with rj the radial distance. This schematic approach for the mixer modeling is justified
since the mixing time is much larger than the impeller revolution time, and there is no
interest in the flow in the impeller region. Only the net effect of the impeller has been
taken into consideration.
The numerical grid used has 106 grid nodes, which is a factor of 20 higher than the
number of nodes used by Van Looy et al. (2004). The simulation was conducted on 4
parallel CPUs, instead of 8 CPUs by Van Looy et al. (2004). While the number of
grid nodes per CPU is much larger in the present simulation, the calculation per impeller
revolution is similar.
Figure 4.10 shows a qualitative picture of the scalar mixing in the storage tank in terms
of instantaneous realizations of the scalar concentration field in a vertical symmetry plane.
The interface between low and high scalar concentration is first sucked into the impeller. A
jet issuing from the side-entry mixer in the bottom left corner appears to cross the domain
and hits the opposing wall, where it is deflected upward. Subsequently, the jet hits the top
wall and is directed to the left. At N t = 300 mixing is taking place in a large part (about
70%) of the domain. The snapshots reveal the mixing time in the storage tank being of the
order of 1000 impeller revolutions.
Following Van Looy et al. (2004), the mixing time is calculated based on the coefficient
of mixing, defined as:
v
u P c c 2 !
u
i
i
c Vi
cmix = t P (4.40)
i Vi

where index i stands for the location of the monitoring node with volume V , and
i = 1, ..., NV , where NV is the total number of control volumes in the flow domain. The
volume averaged mean concentration, c, is calculated as:
98 Chapter 4. Development and validation of a scalar mixing solver

Nt = 100

c / c
2

Nt = 200

1
Nt = 300

Nt = 400 0

Nt = 500

Figure 4.10: Snapshots of the concentration field in a vertical symmetry plane.


4.7. Validations 99

1
Fluent
0.8 LBFV

0.6
cmix (-)

0.4

0.2

0
0 400 800 1200 1600
Nt (-)

Figure 4.11: The coefficient of mixing as a function of dimensionless time. The symbols
represent simulated data points by Fluent (not shown in the paper of Van Looy
et al. (2004)). The solid line is the simulated time trace of the coefficient of
mixing obtained with our lattice-Boltzmann and finite volume (i.e. LBFV)
solvers.

P
i (ci Vi )
c= P (4.41)
i Vi

For industrial applications, a value of cmix = 0.01 is typically used as a criterion


for well-mixed conditions. In the work of Van Looy et al. (2004), the mixing time is
defined as the time needed for reaching the above specified value of cmix . However, for
the simulation at infinite Froude number (not shown in Van Looy et al. (2004)) a value of
0.05 has been used.
In Figure 4.11, a simulated time trace of the coefficient of mixing is shown (solid line).
The symbols represent the simulated coefficient of mixing by Fluent. The results agree
very well. Our simulation estimates a mixing time of about 1600 impeller revolutions,
which is 35% higher than the Fluent prediction. As a result, the mixing performance
predicted by Fluent and our code compare well.
The experimental value of the mixing time at infinite Froude number corresponds to
about 4500 impeller revolutions (see Figure 5 in the paper of Van Looy et al. (2004)).
The predicted mixing times are significantly below this experimental value (up to 75%).
However, it is unclear which criterion was used to determine the mixing time in the ex-
periments. Furthermore, the measurement uncertainty was not documented. As a result,
the predicted mixing times cannot be equally compared with the experimentally measured
mixing time.
100 Chapter 4. Development and validation of a scalar mixing solver

1.03

1.02
M/M0 (-)

1.01

0.99
0 400 800 1200 1600
Nt (-)

Figure 4.12: The evolution of the total mass in the system in time.

Nevertheless, the underprediction of the mixing time is significant. A possible expla-


nation on the numerical side may be the Smagorinsky (1963) subgrid-scale model being
too dissipative in the first stage of the simulation. While initially the flow field is fully
resolved in at least large parts of the tank (the simulation starts with a zero-velocity field),
the Smagorinsky (1963) subgrid-scale model artificially generates extensive values of the
eddy viscosity, resulting in more mixing than experimentally observed. In this respect,
it would be interesting to study the performance of the Voke (1996) subgrid-scale model
(see chapter 2). This model significantly suppresses the eddy viscosity in regions where
the flow is fully resolved.
The performance of the ghost cell technique, applied for modeling the tanks outer
wall, is checked by monitoring the total mass (M ) in the system as a function of time.
The time trace of the total mass normalized with the total mass at t = 0 (M0 ) is shown
in Figure 4.12. Till N t = 180, the total mass in the system remains nearly constant, for
N t > 180 the total mass increases gradually. At N t = 1600 the total mass has increased
by 2.5%, and the average mass increase per impeller revolution equals 0.0016%.
The reason for the mass increase may be understood from a close inspection of the
concentration snapshots. Till N t = 180 the interface between high and low scalar con-
centration near the walls moves slowly downward at the left side and upward at the right
side of the tank (see e.g. the upper snapshot in Figure 4.10). At N t = 180 the jet of high
concentration collides on the wall, and is forced upward (not shown). From this point in
time, the total mass starts to increase gradually. These observations may indicate that the
ghost cell technique has problems with keeping up the fast changes in scalar concentration
near the wall which occur for N t > 180.
4.8. Conclusions 101

4.8 Conclusions
In this chapter, we have developed a scalar mixing solver which can be used as a tool
for studying blending and two-phase mixing processes including mass transport (such as
particle coating, crystallization and dissolution processes). For reasons of computational
memory usage, the scalar mixing solver is based on the finite volume method. The dis-
cretization of the convection-diffusion equation is treated fully explicitly, which makes
implementation relatively simple.
The discretization of the temporal and spatial terms in the convection-diffusion equa-
tion is second-order accurate. Time discretization is performed with the Adams-Bashford
scheme. The diffusion and convection terms are discretized with the central differencing
and flux-limiting (TVD) schemes, respectively. The latter scheme significantly suppresses
numerical diffusion, and is bounded (in contrast to the QUICK scheme). The effect of
subgrid-scale diffusion is taken into account with an eddy diffusion coefficient, which
scales with the eddy viscosity.
The zero-gradient boundary constraint for a scalar at off-grid walls is imposed by
means of a newly developed immersed boundary technique, which is in line with the im-
mersed boundary technique developed by Derksen & Van den Akker (1999) for imposing
the no-slip velocity constraint. In contrast to the commonly applied staircase technique in
a Cartesian grid environment, the immersed boundary technique provides a more accurate
representation of the off-grid walls.
Simulations of the scalar mixing in a square cavity have demonstrated the potential of
the flux-limited scheme (unconditionally stable, negligible numerical diffusion) in favor of
the upwind and QUICK schemes. Furthermore, it is proven that the scheme itself is mass
conservative.
The scalar mixing in an inclined cavity has been simulated by making use of the stair-
case and the newly developed immersed boundary technique. Both techniques are not mass
conserving. At various grid resolutions, the total change in mass is for both techniques
comparable. However, the ghost cell technique provides a more accurate representation of
the boundary compared to the staircase technique.
An LES of the mixing performance of a turbulent flow in a cylindrical tank with a side-
entry mixer has successfully reproduced the mixing time obtained by a LES with Fluent.
Furthermore, the mass change resulting from the application of the immersed boundary
technique was not significant because of the fine grid resolution adopted.
The errors induced by the ghost cell technique may become more significant if the size
of an object (e.g. an impeller blade) is not accurately covered by the grid: a coarse grid
resolution results in less ghost nodes for imposing the zero-gradient constraint compared
102 Chapter 4. Development and validation of a scalar mixing solver

to a fine grid. Therefore, the ghost cell technique should be critically assessed in future
simulations.
4.8. Conclusions 103

Nomenclature

Roman Description Unit


ai , bi coefficients in finite volume discretization scheme kg.s1
ci coefficient in finite volume discretization scheme rep- kg.s1 .[]
resenting the amount of anti diffusion
A surface area m2
c concentration kg.m3
cmix mixing index
c spatial average of the concentration kg.m3
c final concentration kg.m3
D diameter propeller m
Fr Froude number
FC convective prefactor in equation 4.3 kg.s1
FD diffusive prefactor in equation 4.3 kg.s1
g magnitude of the gravitational acceleration vector m.s2
L width/height of the lid driven cavity m
M total mass kg
M0 total mass at t = 0 kg
n normal vector
ni normal vector component i
N impeller speed s1
NV number of control volumes
Pe Peclet number
Q variable defined in equation 4.8 kg.m3 .s1 .[]
rj radial distance m
r, r+ , r ratio of consecutive scalar gradients
r0 reciprocal of r
Re Reynolds number
S surface of total control volume m2
S source term in equation 4.1 kg.m3 .s1 .[]
Sc Schmidt number
Sct turbulent Schmidt number
t time s
T diameter of the cylindrical tank m
u velocity m.s1
104 Chapter 4. Development and validation of a scalar mixing solver

ui velocity component i m.s1


Ujet impeller stream velocity m.s1
Ulid velocity of the band driving the flow in the cavity m.s1
vax axial velocity component m.s1
vtan tangential velocity component m.s1
x, y, z coordinate directions
xi coordinate i m
xw wall position m
x0w wall position normalized with the lattice spacing

Greek Description Unit


scalar diffusion coefficient
e eddy diffusion coefficient m2 .s1
mol molecular diffusion coefficient m2 .s1
lattice spacing
t time step s
V control volume m3
density difference kg.m3
weighting parameter in equation 4.7
kinematic viscosity m2 .s1
e eddy viscosity m2 .s1
density kg.m3
general scalar variable []
I value of a general scalar variable at an inlet []
1 , 0 , 1 values of a general scalar at the ghost node, and two []
points in the flow domain
flux-limiting function
total control volume m3
5 A parameter study of the mixing
time in a turbulent stirred tank by
means of LES

The mixing performance of the turbulent flow generated by a Rushton tur-


bine has been studied numerically for a wide range of conditions by means of
large eddy simulations (LES) including passive scalar transport. The starting
point is a simulation of an experimental benchmark case of a Rushton turbine
stirred tank. The lattice-Boltzmann Navier-Stokes solver and the Smagorin-
sky subgrid-scale model have been adopted for solving the turbulent stirred
tank flow. A finite volume scheme for solving the convection-diffusion equa-
tion for a passive scalar has been coupled to the LES. The simulation results
in this benchmark case agree very well with the experimental data from the
literature. Subsequently, a parameter study has been carried out varying im-
peller size and injection position of the passive scalar. All mixing times in
these simulations differ from various experimental data from the literature by
less than 30% only. The empirical mixing time correlation due to Ruszkowski
is evaluated in greater detail.
Key words: stirred tank, turbulence, mixing time, LES, scalar transport
This chapter has been submitted to AIChE J.

5.1 Introduction
5.1.1 Experimental research on turbulent scalar mixing
Stirred tanks play an important role in the chemical, pharmaceutical, food and water treat-
ment industries. The quality of paints, polymers, detergents, drugs and foodstuffs depend
106 Chapter 5. A parameter study of the mixing time by means of LES

on the geometry and operating conditions of the stirred tank. As a result, the mixing
performance is of high importance to achieve process optimization.
The emphasis of past experimental research for the flow configurations and impellers
has been directed to understanding the velocity characteristics of the turbulent flow (e.g.
Yianneskis et al., 1987; Schfer et al., 1998) together with the power required to drive the
stirrer (e.g. Rushton et al., 1950; Holmes et al., 1964). The essential requirement of mixers
in liquid systems is, however, to bring together two or more fluids, which are initially
separate and this implies the need for information of the scalar mixing characteristics of
the stirrers.
One of the most crucial parameters is the mixing time, which is the time to achieve
complete homogenization of inserted passive scalars. A large number of experimental
studies focused on mixing performance in terms of the mixing time for different tank and
impeller configurations. Kramers et al. (1953) were among the first to report on mixing
time in a propeller agitated tank as a function of the baffle position and impeller rota-
tional speed. Moo-Young et al. (1972) investigated the influence of Newtonian and non-
Newtonian fluids in different flow configurations on the mixing time, and other researchers
focused on the impeller configurations and/or operating conditions in transitional and tur-
bulent flow regions (Hoogendoorn & Den Hartog, 1967; Shiue & Wong, 1984; Sano &
Usui, 1985; Bouwmans et al., 1997; Distelhoff et al., 1997). Results of several studies
have been combined into empirical correlations, which can be of use for industrial appli-
cations (Prochazka & Landau, 1961; Sano & Usui, 1985; Ruszkowski, 1994).
The most common methods for mixing time measurements are the conductivity probes
(e.g. Shiue & Wong, 1984; Sano & Usui, 1985; Ruszkowski, 1994), or decolorization
techniques (Moo-Young et al., 1972). The disadvantage of the conductivity probe is that
it disturbs the flow field (intrusive). The decolorization technique is based upon visual
observation and relies on a subjective decision by the worker. Another technique more
recently used by Distelhoff et al. (1997) is based upon laser induced fluorescence (LIF) for
measuring the scalar concentration in continuously operated stirred tanks. The advantages
of this technique are that the flow remains undisturbed, the measurement volume is an
order of magnitude smaller compared to conductivity probes, and the ability of LIF to
measure time traces of the concentration throughout the tank.
In literature, there is no standardization of a mixing time experiment. In the first place,
mixing time experiments are executed in tanks of different geometries and operating con-
ditions. Secondly, the passive tracer is injected at various injection speeds, thereby chang-
ing the influence of jet mixing and stirred tank mixing effects, and positions in the tank.
Thirdly, there is no agreement on the position and number of measurement points. And
finally, the definition of the mixing time varies from study to study. Most of the studies
5.1. Introduction 107

report the mixing time based on a concentration % (e.g. Bouwmans et al. (1997) and
Distelhoff et al. (1997)), others define mixing time by the variance of concentration fluc-
tuations (e.g. Ruszkowski (1994)).
In particular the variation of injection speed, number and position of the measurement
points, and the definition of the mixing time in the experimental studies result in unequal
comparisons between a measured and/or simulated mixing time, and a mixing time or a
correlation reported in literature. As a result, a measured and/or simulated mixing time
should be viewed with great care. This also applies to mixing time correlations in the
literature.

5.1.2 Potential of CFD on scalar mixing


Another way of describing and understanding the complex turbulent flow in stirred tanks
is through computational modeling. Simulations based on the Reynolds-Averaged Navier-
Stokes (RANS) equations (contained in various commercial software packages) provide
a reasonable picture of the overall mixing patterns in the whole tank. The mean flow is
directly resolved, whereas the Reynolds stresses (related to turbulent velocity fluctuations)
are modeled with a turbulence model. The k- eddy-viscosity model is often used, which
assumes isotropic turbulent transport. It is however inappropriate in rotating and/or highly
three dimensional flows (Wilcox, 1993). While RANS models predict the mean flow rea-
sonably, it invariably underpredicts the turbulent kinetic energy in the impeller discharge
stream by about 50% (e.g. Hartmann et al., 2004a). It is expected that this will certainly
affect the predictions of the mixing patterns and mixing time in stirred tanks.
There have been many modeling studies of turbulent flow characteristics in stirred
tanks, but only a few investigations of mixing patterns have appeared in the literature (e.g.
Osman & Varley, 1999; Jaworski et al., 2000; Bujalski et al., 2002). Bouwmans et al.
(1997) used a particle tracking technique in a RANS flow field for blending liquids with
different densities. Most of these studies have used the RANS approach within a sliding
mesh or multiple frames of reference (MFR) framework, which is believed to provide fully
predictive simulations of the mixing time. The former (sliding mesh) approach is a fully
transient approach, while the latter provides a steady-state approximation of the flow field.
It is evident that the sliding mesh approach is more accurate but it it also much more time
consuming than the multiple frames of reference approach.
In spite of the use of the sliding mesh approach, the predicted mixing times were found,
in general, 2 3 times higher than the measured values (e.g. Osman & Varley, 1999;
Jaworski et al., 2000; Bujalski et al., 2002), which is in line with the underprediction of
the turbulence levels (Ng et al., 1998; Hartmann et al., 2004a). While in case of Rushton
108 Chapter 5. A parameter study of the mixing time by means of LES

turbines the mixing time is usually overpredicted due to poor mixing across the central
plane of the discharge stream, RANS simulations of pitched blade turbines sometimes
yield underpredicted mixing times. This may be due to concentration fluctuations being
not really included.
The main difficulty with the RANS approach in combination with a closure model is
that there is no clear distinction between the part of the turbulent fluctuations that is directly
resolved and the part which is represented by the Reynolds stresses. While the mean flow
is predicted reasonably well, the turbulence levels are underpredicted by 50%, and this has
a negative effect on the mixing characteristics and time. In contrast, an approach that is
capable of resolving the turbulent nature of the flow is the large eddy simulation (LES),
which is getting more attention in the last years. In LES there is a clear distinction between
resolved and unresolved scales, as the range of resolved scales is limited by the numerical
grid.
LES has shown in the past that it is a powerful tool to study stirred tank flow (Eggels,
1996; Derksen & Van den Akker, 1999; Hartmann et al., 2004a,b; Yeoh et al., 2004;
Bakker & Oshinowo, 2004), because it accounts for the unsteady behavior of these flows
and it can be effectively employed to explicitly resolve the phenomena directly related to
the unsteady boundaries. The LES methodology has been recently applied for a mixing
time simulation in the study reported by Yeoh et al. (2005). Their simulation was de-
signed to match the experimental set-up of Lee (1995). The results focus on the mixing
patterns, time traces of the concentration and comparison of the predicted mixing time
with correlations presented in the literature.
This work aims at a thorough analysis (by means of LES) of the mixing time in a
stirred tank geometry as a function of impeller size and injection position of the passive
scalar. In contrast to the work of Yeoh et al. (2005), the passive scalar is injected at zero-
speed to avoid the influence of jet mixing effects on the mixing time. The starting point
is a numerical reproduction of the mixing time experiment of Distelhoff et al. (1997).
Subsequently, large eddy simulations have been performed with different impeller sizes
and injection positions. With a view to the various definitions of the mixing time found
in the literature, careful comparisons are made between the simulated mixing times in this
work and those obtained by two experimentally obtained mixing time correlations reported
by Sano & Usui (1985) and Ruszkowski (1994).
The Navier-Stokes equations are solved with a lattice-Boltzmann discretization scheme
(Somers, 1993; Derksen & Van den Akker, 1999). This scheme is easy to implement and
parallelize, and one does not need to solve the Poisson equation for pressure as in finite
volume solvers. Furthermore, we have coupled to the Navier-Stokes solver a finite volume
discretization scheme for the scalar transport, which is less memory intensive compared
5.2. Flow system 109

T/10
D = T/3
r 3D/4

H=T 0.04D D/5


0.04D
0.16D
T/3
z
D/4
T

Figure 5.1: Cross-section of the tank (left). Plan view and cross-section of the impeller
(right). At the top level there is a lid. The impeller is a Rushton turbine
mounted at height T /3 and has a diameter T /3. In the simulations presented
in this work, impeller sizes of T /4 and T /2 have been used as well.

to a lattice-Boltzmann discretization. In this way we make use of the advantages of both


schemes.

5.2 Flow system


The stirred tank used in this work was a standard configuration cylindrical tank of diameter
T = 147 mm, with four equi-spaced baffles of width 0.1T mounted along the perimeter of
the tank (Distelhoff et al., 1997). The liquid height was set equal to the tank diameter, H =
T . The impeller was a six-bladed Rushton turbine with standard dimensions, mounted at
height T /3. A schematic representation of the flow system is shown in Figure 5.1. The
flow system can be fully characterized by the flow Reynolds number (Re=N D 2 /) if
geometric similarity is maintained. With an impeller speed (N ) of 10 s1 , T /D = 3,
and a kinematic viscosity of 1.0 106 m2 /s (tap water at room temperature) the Reynolds
number yields 24, 000 (as also used in Distelhoff et al., 1997).

5.3 Simulation procedure


5.3.1 Large eddy simulation
For the simulation of flow at industrially relevant Reynolds numbers (i.e. at strongly tur-
bulent conditions), direct simulation of the flow is not feasible and turbulence modeling
110 Chapter 5. A parameter study of the mixing time by means of LES

is required. In a Large Eddy Simulation (LES), the small scales in the flow are assumed
to be universal and isotropic, and the effect the small scales have on the larger scales is
modeled with a subgrid-scale model. The filtering of the small-scale motion is based on
the assumption that the motion of the smallest scales is isotropic in nature and that the
subgrid-scale energy is dissipated via an inertial subrange that has a geometry indepen-
dent character. The LES model applied in this research is a standard Smagorinsky model
(Smagorinsky, 1963). For more details on the LES methodology used we refer to Hart-
mann et al. (2004a,b) and Derksen & Van den Akker (1999).
A lattice-Boltzmann method (Chen & Doolen, 1998) was used for solving the filtered
momentum equations. The specific scheme we used was introduced by Somers (1993),
and is described in more detail in Derksen & Van den Akker (1999).
The entire tank was simulated on a uniform, cubic computational grid. Inside the com-
putational domain, the no-slip boundary conditions at the cylindrical tank wall, the baffles,
the impeller, and the impeller shaft were imposed by an adaptive force-field technique
(Derksen & Van den Akker, 1999).

5.3.2 Scalar transport


In order to describe scalar transport, the convection diffusion equation needs to be solved.
Eggels & Somers (1995) have performed scalar transport calculations on free convective
cavity flow with the lattice-Boltzmann discretization scheme. This scheme, however, is
more memory intensive than a finite volume formulation of the convection-diffusion equa-
tion. In a finite volume discretization we only need to store two or three (depending on the
time integrator; Euler-forward or Adams-Bashford, respectively) double precision concen-
tration fields, whereas in the lattice-Boltzmann discretization typically 18 single-precision
variables need to be stored. Therefore, we have coupled to the lattice-Boltzmann flow
solver a finite volume discretization scheme for the scalar transport.
The convection-diffusion equation in compressible form reads
 
uj
+ = + S (5.1)
t xj xj xj
where is the density of the continuous phase, is a general scalar, uj is velocity com-
ponent j, is the diffusion coefficient (i.e. the sum of the molecular and eddy diffusion
coefficients: = mol + e ) and S is a source term. Since the lattice-Boltzmann scheme
is a compressible scheme, we have implemented the discretized form of the compress-
ible convection-diffusion equation. In this context it should be noted that the maximum
5.3. Simulation procedure 111

velocity (which is approximately the tip speed) is set sufficiently low for meeting the in-
compressibility limit in the lattice-Boltzmann scheme.
For reasons of simplicity, we consider in the following discussion the 1-D version of
the above equation. Extension to multiple dimensions is straightforward. Integration over
a finite volume V and over time t, and using the Gauss theorem leads to

V
(i i )n+1 (i i )n = (FC )ni+ 1 + (FC )ni 1

t 2 2
n n n (5.2)
+ (FD )i+ 1 (i+1 i )
2
n
(FD )i 1 (ni ni1 )
2

where FC and FD are convective and diffusive prefactors, e.g.

FC;i+ 12 = (uA)i+ 1 (5.3)


2

and
 
A
FD;i+ 12 = (5.4)
i+ 21

where u is the normal velocity through the cell face with surface A.
The density, normal velocity components and the diffusion coefficient at the cell faces
are approximated by linear interpolation between the surrounding cell nodes values. For
reasons of simplicity, the time discretization in equation 5.2 is first-order Euler forward.
We have implemented a second-order accurate Adams-Bashford time discretization scheme.
The convection-diffusion equation is solved on the LES grid, and the time step is the
same as the LES time step. The finite volume formulation presented is fully explicit;
the update of the scalar value at cell node i is determined by the scalar values, velocity
components and density at cell node i and its surrounding nodes at time instant n 1.
An explicit scheme has severe restrictions on the time step in order to guarantee stability.
However, the time step used for the LES is very small, and no stability problems were
encountered in the scalar transport calculations.
The molecular Schmidt number (Sc=/mol ) is typically of the order of 1000, which
means that the smallest scale of mass transfer (i.e the Batchelor scale) is a factor 30 smaller
than the smallest scale of turbulence (i.e. the Kolmogorov scale). Hence, it is expected
that the influence of the scalar subgrid-scale fluctuations on the resolved-scale scalar field
is larger than the influence of the subgrid-scale velocity fluctuations on the resolved-scale
112 Chapter 5. A parameter study of the mixing time by means of LES

velocity field. From this, it is reasonable to take a turbulent Schmidt number (Sc t =
e /e ) to be smaller than unity (we chose 0.7).
The cell face values of the general scalar , needed for the convection terms (the first
two terms on the right hand side of equation 5.2), have been approximated with a second-
order TVD (Total Variation Diminishing) scheme, that has been introduced by Harten
(1983). This scheme belongs to the family of high resolution schemes and does not suffer
from numerical diffusion (in contrast to first-order upwind, power-law and hybrid schemes
(Patankar, 1980)), and it is unconditionally stable (contrary to central differences, QUICK
(Leonard, 1979)). Following Roe (1981), the face value i+ 21 is written as the sum of a
diffusive first-order upwind term and an anti-diffusive one.

1  
i+ 21 = i + ri+ 12 (i+1 i ) (5.5)
2
The anti-diffusive part is multiplied by the flux limiter function (r), which is a non-
linear function of r, which is the upwind ratio of consecutive gradients of the solution.
According to Wang & Hutter (2001), the so-called Superbee flux limiter is the least diffu-
sive of all acceptable limiters. Therefore, we have used the Superbee limiter throughout
our simulations. The Superbee limiter is defined as:
h i
(r) = max 0, min[1, 2r], min[r, 2] (5.6)

5.3.3 Immersed boundary technique


The Neumann boundary condition at the non-square or moving (off-grid) objects in the
flow domain are imposed by means of a newly developed immersed boundary technique,
which is applicable in the explicit formulation of the finite volume scheme. With our
novel technique, the off-grid walls are more accurately represented (especially the impeller
walls) compared to the well-known technique approximating the walls by staircases. The
algorithm makes use of so-called ghost cells (i.e. boundary cells with their cell centers
outside the flow domain) in order to impose a zero-gradient scalar constraint (i.e. Neumann
boundary condition) at the walls that are off-grid.
The scalar update is executed only for the cells with their centers inside the flow do-
main. In order to update the scalar in a cell near a wall, the scalar value in the neighboring
ghost cell is needed. The procedure for the determination of the scalar ghost cell value is
shown in Figure 5.2. We consider here a curved wall immersed in a Cartesian grid. Two
scalar values on a line through the ghost cell center and perpendicular to the wall are esti-
mated through bi-linear interpolation of the surrounding cell nodes. The spacing between
5.3. Simulation procedure 113

Figure 5.2: Schematic representation of the determination of the scalar ghost cell value.
White cells are flow cells, gray cells are the ghost cells, black cells are exterior
cells.

the positions of the two estimated scalar values and the ghost cell center is taken equal
to the lattice spacing. The scalar value at the ghost cell is determined via second-order
extrapolation with the two scalar values and the zero-gradient constraint at the wall.
The ghost cell method is unconditionally stable, which is desirable in the explicit finite
volume formulation. Other techniques that make use of body-fitted cells (so-called cut
cells, Tucker & Pan (1999); Calhoun & LeVeque (2000)) will certainly fail, as the time
step scales with the cell volume. A drawback of the proposed technique is that it does not
automatically guarantee scalar mass conservation, and therefore this needs to be checked
during the simulations.

5.3.4 Simulation aspects


Mixing time simulations have been performed in a Rushton turbine stirred tank. The ex-
perimental benchmark case studied, is based on the work of Distelhoff et al. (1997). They
observed for Re> 104 only a slight increase of the dimensionless mixing time with in-
creasing Reynolds number. A key parameter dominating the mixing time is the impeller
diameter according to the correlation proposed by Sano & Usui (1985) and Ruszkowski
(1994). We have investigated the influence of the impeller diameter on the mixing time by
means of three flow simulations, each with a different impeller diameter. The tank over
impeller diameter ratios equal T /D = 2, T /D = 3 and T /D = 4. The Reynolds number
in each of the simulations equals 24, 000. Furthermore, we have varied the injection posi-
tion of the passive scalar. Four injection positions have been chosen, and as a result, four
scalar mixing calculations have been performed simultaneously per flow simulation. The
different cases studied are listed in Table 5.1. Please note that increases in the direction
114 Chapter 5. A parameter study of the mixing time by means of LES

Table 5.1: Numerical setup. The parameters varied are the impeller diameter and the feed
location. Please note that increases in the direction of the impeller rotation,
and = 0o is a mid-way baffle plane. rp is the feed pipe radius.
Case T /D Feed: r/T Feed: z/T Feed: Feed: rp /T
2A 2 0.17 1 0o 0.0238
2B 2 0.1875 0.483 0o 0.0125
2C 2 0.35 0.333 0o 0.0125
2D 2 0.475 0.8 320o 0.0125
3A 3 0.17 1 0o 0.0238
3B 3 0.125 0.467 0o 0.0125
3C 3 0.211 0.333 0o 0.0125
3D 3 0.475 0.8 320o 0.0125
4A 4 0.17 1 0o 0.0238
4B 4 0.09375 0.458 0o 0.0125
4C 4 0.225 0.333 0o 0.0125
4D 4 0.475 0.8 320o 0.0125

of the impeller rotation.


Case 3A resembles the Distelhoff et al. (1997) experiment. The feed position is located
at the top of the tank in a mid-way baffle plane (i.e. = 0o ) at 0.17T from the tank
centerline. Cases 2A and 4A have the same feed location and dimension of the feed pipe,
but with tank over impeller ratios T /D = 2 and T /D = 4, respectively.
Cases 2B, 3B and 4B have the feed pipe location at 0.1T above the top of the impeller
blade, and at a radial position halfway along the impeller blade. Cases 2C, 3C and 4C have
the feed pipe located at disk height at 0.1T from the impeller tip. In cases 2D, 3D and 4D
the passive scalar is injected in the wake of a baffle (i.e. at 5o angle with respect to the
baffle) at 0.8T axial height.
The tracer injection time in the mixing time experiment was chosen to be less than 1%
of a typical mixing time (Distelhoff et al., 1997). As a result, the injection time was set at
half an impeller revolution (i.e. N t = 0.5). This injection time was set in all simulations
performed.
Time traces of the scalar concentration were recorded at various monitoring points
positioned in accordance with Distelhoff et al. (1997). A total of 32 monitoring points
were set in the four vertical planes mid-way between two baffles. The points were located
at 0.19T and 0.67T axial heights at r/T = 0.126, 0.252, 0.361, 0.469, respectively.
The code runs on a parallel computer platform by means of domain decomposition:
the computational domain was horizontally (i.e. perpendicular to the tank centerline) split
5.4. Results 115

in eight equally-sized subdomains.


For the coupled LES/scalar mixing simulations, a cubic, Cartesian grid of 240 3 lattice
cells was defined. The diameter of the tank equals 240 lattice spacings, and hence the
spatial resolution equals T /240 which corresponds to 0.6125 mm in the experiment. The
temporal resolution is limited by the lattice-Boltzmann method. In order to meet the in-
compressibility limit in the lattice-Boltzmann discretization scheme, the tip speed of the
impeller should approximately be set at 0.1 lattice spacings per time step. As a result,
the impeller speed is different for the three impeller types. For the T /D = 2, 3, 4 cases
the impeller makes a full revolution in 3600, 2400 and 1800 time steps, respectively. The
corresponding temporal resolutions in physical units are 62.6 s, 41.7 s and 31.3 s (at
N = 4.44 s1 , 10 s1 and 17.77 s1 , respectively).
For the LES, 21 (18 directions for the LB-particles and 3 force components) single-
precision, real values need to be stored. For the scalar mixing simulation twelve (four
concentration fields at the time instants n + 1, n and n 1) double-precision, real scalar
values needs to be stored. The memory requirements of the simulation are proportional to
the grid size, resulting in an executable of about 2 GByte. The simulations were performed
on an in-house PC cluster with eight Athlon 1, 800+ MHz processors using an MPI mes-
sage passing tool for communication within the parallel code. To calculate one time step
takes about 58 seconds wall-clock time, hence an impeller revolution takes 1 2 days.
The scalar was injected in a quasi steady state flow field that was obtained from a zero-
velocity flow field. The quasi steady state was reached after 10 40 impeller revolutions
(depending on the impeller size), which has been checked by monitoring the total kinetic
energy of the flow as a function of time.

5.4 Results
5.4.1 Snapshots of the scalar concentration field
Figure 5.3 gives an impression of the scalar mixing process in the Rushton turbine stirred
tank (Case 3A; T /D = 3) during the course of the simulation. In the first stage of the
mixing process, high concentration, macroscopic structures are identified that are advected
by the flow from the tracer injection point (at the top of the tank) toward the impeller
region. These structures are vigorously mixed by the turbulence generated by the revolving
impeller. After about 30 impeller revolutions, the scalar concentration is heading toward a
homogeneous distribution.
The impact of the impeller diameter on the duration of the mixing process is illustrated
116 Chapter 5. A parameter study of the mixing time by means of LES

c/c
5

4
Nt = 3 Nt = 9 Nt = 15
3

Nt = 21 Nt = 27 Nt = 33

Figure 5.3: Case 3A; T /D = 3. Six instantaneous realizations of the scalar field in a mid-
way baffle plane. The scalar is injected at the top of the tank (black dot in the
graphs).

in Figure 5.4. In this case, the passive scalar is injected 0.1T above the impeller disk.
Again, high concentration macroscopic structures are identified in all cases presented. The
impeller diameter significantly influences the mixing process: the mixing time decreases
at increasing impeller diameter. While in case 4B (T /D = 4) more than 32 revolutions
are needed to reach a more or less homogeneous scalar distribution, in case 2B (T /D = 2)
only 8 revolutions are necessary to reach a similar situation.

5.4.2 Time series


Case 3A: comparison simulation and experiment

Simulated and experimentally measured time series are shown in Figure 5.5 at four mon-
itoring points (i.e. two radial positions and two axial heights) in a vertical mid-way baffle
plane at = 90o . The experimentally measured time series are taken from Distelhoff
et al. (1997). A qualitative comparison between the simulated and experimentally mea-
5.4. Results 117

Nt = 8 Nt = 16 Nt = 32
c/c
5

1
Nt = 6 Nt = 12 Nt = 24
0

Nt = 4 Nt = 8 Nt = 16

Figure 5.4: Cases 4B (T /D = 4); upper graphs, 3B (T /D = 3); middle graphs and 2B
(T /D = 2); lower graphs. Three instantaneous realizations of the scalar field
in a mid-way baffle plane. The scalar is injected 0.1T above the impeller (black
dot in the graphs).

sured time series results in the conclusion that the time series compare well in terms of the
path of the curves and the mixing time scale. The time step used in the simulations is a
factor of 20 smaller than the sampling time in the experiments (Distelhoff et al., 1997). As
a result, the simulated time series show more concentration fluctuations compared to the
experimentally measured time series.
The simulated time series at z/T = 0.19 show a delayed response on the tracer injec-
118 Chapter 5. A parameter study of the mixing time by means of LES

5 5
r/T = 0.252 r/T = 0.252
4 0.469 4 0.469

3 3
c/c (-)

c/c (-)
2 2

1 1

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Nt (-) Nt (-)

(a) Simulation, z/T = 0.67 (b) Experiment, z/T = 0.67

5 5
r/T = 0.252 r/T = 0.252
4 0.469 4 0.469

3 3
c/c (-)

c/c (-)

2 2

1 1

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Nt (-) Nt (-)

(c) Simulation, z/T = 0.19 (d) Experiment, z/T = 0.19

Figure 5.5: Case 3A, tracer injection point at the top of the tank. Simulated (a,c) and
experimental (b,d) time series at various monitoring points in a vertical mid-
way baffle plane ( = 90o ). The experimentally measured time series are taken
from Distelhoff et al. (1997).

tion (0 N t 0.5) compared to the time series at z/T = 0.67, which is to be expected
as the tracer is injected at the top of the tank. The experimental time series show an earlier
response on the tracer injection compared to the simulated time series.
5.4. Results 119

5 5
z/T = 0.67 z/T = 0.67
4 0.19 4 0.19

c/c (-)
3
c/c (-)

2 2

1 1

0 0
0 4 8 12 16 20 0 10 20 30 40 50 60
Nt (-) Nt (-)

(a) Case 2B, T /D = 2, r/T = 0.126. Data for (b) Case 3B, T /D = 3, r/T = 0.126.
N t < 2 is missing.

5
z/T = 0.67
4 0.19

3
c/c (-)

0
0 20 40 60 80 100
Nt (-)

(c) Case 4B, T /D = 4, r/T = 0.126.

Figure 5.6: Cases 2B, 3B and 4B, tracer injection point at 0.1T above impeller blade.
Simulated time series at various monitoring points in a vertical mid-way baffle
plane ( = 90o ).

Cases 2B, 3B, 4B: influence impeller size

In Figure 5.6 simulated time series are shown for the three different impeller dimensions
at two monitoring points. In this case, the tracer is injected at 0.1T above the impeller
blade. Due to a processing error, the data for N t < 2 is missing in Figure 5.6a. According
120 Chapter 5. A parameter study of the mixing time by means of LES

5 5
z/T = 0.67 z/T = 0.67
4 0.19 4 0.19

3 3
c/c (-)

c/c (-)
2 2

1 1

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Nt (-) Nt (-)

(a) Case 3A, r/T = 0.126. (b) Case 3B, r/T = 0.126.

5 5
z/T = 0.67 z/T = 0.67
4 0.19 4 0.19

3 3
c/c (-)

c/c (-)

2 2

1 1

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Nt (-) Nt (-)

(c) Case 3C, r/T = 0.126. (d) Case 3D, r/T = 0.126.

Figure 5.7: Cases 3A, 3B, 3C and 3D (i.e. four different feed point positions). Simulated
time series at various monitoring points in a vertical mid-way baffle plane ( =
180o ).

to Sano & Usui (1985) and Ruszkowski (1994), mixing time strongly depends on the
impeller size. This is confirmed by the simulated time series. For T /D = 2 only 14
impeller revolutions are needed to reach more or less the final concentration. About 45
and 100 impeller revolutions are needed for T /D = 3 and T /D = 4, respectively, to
reach a similar stage.
5.4. Results 121

Cases 3A, 3B, 3C, 3D: influence injection position

Figure 5.7 shows simulated time series with the T /D = 3 impeller type. The position
of the tracer injection point has a small effect on the mixing time scale. The time traces
corresponding to the cases 3A and 3D show that after roughly 45 impeller revolutions
the concentration approximates the final level. In cases 3B and 3C a similar situation is
reached after roughly 35 impeller revolutions.

5.4.3 Mixing time and coefficient of mixing


The mixing time, m , is the time required to mix the added passive tracer with the contents
of the tank until a certain degree of uniformity is achieved. The location in the tank where
the passive tracer concentration is measured, has to represent the state of mixing of the
entire vessel (i.e. the mixing time measured at any point in the tank should be the same).
Distelhoff et al. (1997) measured mixing times where the concentration variations were
smaller than 10%, 5% and 1% of the fully mixed concentration. These concentrations are
called the 90%, 95% and 99% concentration. The mixing times defined by the 90% and
95% concentration varied by up to 27% and 21%, respectively, between different regions
of the tank. These uncertainties can be reduced by averaging the measurements obtained
simultaneously at several locations. The variation between the mixing times for the 99%
concentration was found substantially smaller. Based on these observations, Distelhoff
et al. (1997) based their mixing time on the 99% concentration.
Another way of defining a degree of mixing is through the variance of the concentration
measured at various positions in the tank. In order to have a precise mathematical formu-
lation of mixing criteria that can be used for estimation of mixing times, the coefficient of
mixing is introduced. The coefficient of mixing is defined as:
v
u P c c 2 !
u
i
i
V i
cmix =t Pc (5.7)
i Vi

where index i stands for the location of the monitoring point with volume V , and
i = 1, ..., K, where K is the total number of monitoring points. The volume averaged
mean concentration, c, is calculated as:
P
i (ci Vi )
c= P (5.8)
i Vi

The method provides the concentration variance of points in different regions of the
122 Chapter 5. A parameter study of the mixing time by means of LES

0.25 4000
fit 1
0.2 fit 2
3000

number (-)
0.15
cmix (-)

2000
0.1
1000
0.05

0 0
70 75 80 85 90 95 100 0.004 0.005 0.006 0.007
% concentration cmix (-)

(a) (b)

Figure 5.8: Calibration graph (a) of coefficient of mixing vs. concentration %. The his-
togram of the coefficient of mixing at 99% concentration is shown in (b), to-
gether with two Gaussian fits.

tank. In order to link the coefficient of mixing to a concentration % (e.g. the 99% concen-
tration) we perform the following numerical experiment. Consider a set of K statistically
independent, dimensionless concentrations ranging between 0.99 and 1.01. The dimen-
sionless concentrations are obtained at the monitoring points. With these numbers, a co-
efficient of mixing can be calculated. If we repeat this experiment L times (i.e inclusion
of time as a variable), we obtain a distribution function of the coefficient of mixing for the
99% concentration. The experiment is repeated for other concentration percentages.
Figure 5.8a shows a calibration graph of the coefficient of mixing vs. the concentration
%. The total number of monitoring points (K) equals 32 and the number of repetition (L)
was set at 105 , which resembles a time interval comparable to the length of the simulated
time series. The graph shows that with increasing concentration %, the average coefficient
of mixing decreases, and the distribution becomes narrower as expected.
Figure 5.8b shows the histogram of the coefficient of mixing for the 99% concentration.
The distribution function is nearly Gaussian, as shown by the two Gaussian fits. Gaussian
fit 1 is fitted in the range 0.0045 0.006 and fit 2 in the range 0.005 0.0065. The average
coefficient of mixing that can be linked to the 99% concentration equals 0.00566. For the
95% and 90% concentrations, the average coefficient of mixing equals 0.0283 and 0.0566,
respectively.
The coefficient of mixing as a function of time calculated from the concentrations at
5.4. Results 123

101 101
2A 3A
0
2B 0
3B
10 2C 10 3C
2D 3D

cmix (-)
cmix (-)

10
-1
10-1
90% 90%
95% -2
95%
10-2 10
99% 99%
-3 -3
10 10
0 5 10 15 20 25 0 10 20 30 40 50 60 70 80
Nt (-) Nt (-)

(a) Cases 2A, 2B, 2C and 2D; T /D = 2. Lines (b) Cases 3A, 3B, 3C and 3D; T /D = 3. Lines
fitted in the range 6 N t 14. fitted in the range 20 N t 35.

101
4A
0
4B
10 4C
4D
cmix (-)

10-1
90%
-2
95%
10
99%
-3
10
0 20 40 60 80 100 120 140
Nt (-)

(c) Cases 4A, 4B, 4C and 4D; T /D = 4. Lines


fitted in the range 40 N t 80.

Figure 5.9: The coefficient of mixing as a function of the dimensionless time. The horizon-
tal lines labeled 90%, 95% and 99% represent the corresponding concentration
percentage.

the 32 monitoring points is shown in Figure 5.9, where all cases are summarized in three
graphs. In Figure 5.9a the data is missing for N t < 2, due to a processing error. The
three graphs show similarities. In the first stage of the mixing process, the macroscopic
124 Chapter 5. A parameter study of the mixing time by means of LES

Table 5.2: The mixing times for the 90%, 95% and 99% concentrations, respectively.
Case N m,90% N m,95% N m,99%
2A 13.6 16.0 21.3
2B 13.0 15.1 20.2
2C 12.1 14.2 19.3
2D 16.4 18.5 23.5
3A 46.8 54.7 73.0
3B 38.1 44.2 58.5
3C 33.0 39.8 55.7
3D 44.5 51.1 66.5
4A 85.0 100.2 135.4
4B 84.5 98.7 131.9
4C 81.8 96.5 130.8
4D 81.4 96.6 131.9

high-concentration structures are broken up and mixed into smaller structures with lower
concentration. The coefficient of mixing strongly fluctuates with a value higher than one,
and no clear trend is observed. In the second stage, all three graphs reveal an exponential
decay of the coefficient of mixing. This has also been stated by Ruszkowski (1994). He
based, similarly to our work, the degree of mixedness on the variance of concentration
fluctuations. In the final stage of the simulations, the coefficient of mixing stabilizes at the
value of approximately 0.02. This value corresponds with the 96% 97% concentration,
which means that concentration fluctuations of 3%4% remain. This is clearly unphysical,
and these fluctuations are attributed to the numerics (e.g. the compressibility of the lattice-
Boltzmann and finite volume schemes used, the introduction of numerical errors at the
walls by the our novel immersed boundary technique).
Table 5.2 lists the value of the 90%, 95% and 99% mixing times, respectively for
all cases. These mixing times are evaluated by intersection of the fitted straight lines
(implying exponential decay) and the horizontal lines corresponding to the 90%, 95%
and 99% concentration, respectively (see Figure 5.9). For the mixing case studied by
Distelhoff et al. (1997), it took 57 impeller revolutions to reach the 99% concentration.
The simulation of case 3A predicts that about 47, 55 and 73 impeller revolutions that
are needed to obtain the 90%, 95% and 99% concentrations, respectively. As a result,
the simulation overestimates the 99% mixing time by 28%. The spread of the mixing
times with respect to the injection points for the T /D = 3 cases, however, is relatively
large compared to the T /D = 2 and T /D = 4 cases. The 99% mixing time for the other
5.4. Results 125

injection positions all lie close to the 99% mixing time reported by Distelhoff et al. (1997).
The mixing time in a stirred tank geometry reported by Yeoh et al. (2005) resembles
the 95% mixing time, and equals roughly 33 impeller revolutions. Our results are 65%,
33%, 20% and 54% higher for the cases 3A, 3B, 3C and 3D, respectively. While case 3A
is most similar to the case studied by Yeoh et al. (2005) (injection position at r/T = 0.17
instead of 0.25), the difference between the predicted mixing times is the largest of all
cases (i.e. 65%). This observation can be explained as follows. In the first place, Yeoh
et al. (2005) used only typically 500k cells for their LES. As a result, their flow and scalar
fields are much less resolved compared to those presented in this work. Less resolution
leads to a reduction of the predicted mixing time. In the second place, in the simulation
of Yeoh et al. (2005) the tracer is injected at an injection speed of 15.9 m/s for about 50%
of their predicted mixing time, whereas in our work the tracer is injected at zero-speed
in accordance to the experiment by Distelhoff et al. (1997). A higher injection speed
introduces jet mixing effects, the passive tracer reaches the impeller region at an earlier
time, and the mixing time is reduced.

5.4.4 Mixing time correlations


Numerical simulations, such as presented in this work, contain all information needed to
have an equal comparison between the simulated mixing times and experimentally ob-
tained mixing times and correlations from the literature. This has been illustrated in this
work by repeating the mixing time experiment of Distelhoff et al. (1997), and it will be
demonstrated in this section.
Sano & Usui (1985) measured the mixing times under turbulent conditions in agitated
tanks equipped with either a paddle or turbine impeller. The operating parameters they
varied were the impeller diameter, the height of the impeller blade (w) and the number of
impeller blades (nb ). The mixing time was determined by the point on a recorder chart at
which the signal reaches the final concentration after the tracer liquid is poured into the
tank, with a degree of relative deviation of concentration < 1% (i.e. the 99% mixing time).
From the measurements the following correlation was proposed:

 w 0.1  T 2.5
N m = 12.2P o3/4 nb (5.9)
T D
In the geometry under consideration in this work, w = 0.2D and nb = 6. The power
number (Po) has been measured by Distelhoff et al. (1997) for the particular case of interest
and equals 4.5. As a result, the Sano & Usui (1985) correlation reduces to
126 Chapter 5. A parameter study of the mixing time by means of LES

150
99%
Sano & Usui (1985)
120

90
Nqm (-)

60

30

0
0 1 2 3 4 5
T/D (-)

Figure 5.10: Simulated 99% mixing times as a function of T /D. The symbols refer to the
cases mentioned in table 5.1. The thick, solid line resembles the Sano & Usui
(1985) mixing time correlation.

 2.4
T
N m = 4.02 (5.10)
D
It is noted that the power number is not independent of T /D according to e.g. Shiue &
Wong (1984). In the measured range 2.2 < T /D < 3, the power number varies between
5.5 (at T /D = 2.2) and 4.3 (at T /D = 3). With respect to the mixing time, this effect
will be less significant (i.e a deviation of 10% 20%) compared to the effect of T /D. As
a result, the relationship between power number and T /D will be ignored in the following
discussions.
In Figure 5.10 the simulation results of the 99% mixing time and the Sano & Usui
(1985) correlation (equation 5.10) are shown as a function of T /D. The mixing time
calculated of all cases with T /D = 2 and the cases 3B and 3C (T /D = 3) compare
very well with the correlation proposed by Sano & Usui (1985). The mixing time for case
3A is overestimated by 30%, and for the cases 3D, 4A-D by 18%. The overestimation is
relatively small compared with the deviations (100% and more) reported in earlier CFD
work based on the RANS approach (e.g. Osman & Varley, 1999; Jaworski et al., 2000;
Bujalski et al., 2002). Furthermore, it should be noticed that
the correlation proposed by
Sano & Usui (1985) is based on measurements in the range 2 T /D 2.5, which only
partly overlaps the T /D range simulated in this work.
5.4. Results 127

Table 5.3: The fit parameters a0 and a1 of the mixing time correlation given in equa-
tion 5.11. The fit parameters are given for the 90%, 95% and 99% mixing
times, respectively.
Concentration % a0 a1
90% 2.46 2.54
95% 2.80 2.56
99% 3.60 2.60

The simulation results are fitted based on a similar correlation:


 a1
T
N m = a 0 (5.11)
D
where a0 and a1 are the fit parameters. The values of the fit parameters are listed in
Table 5.3. The value of the fit parameter a1 shows an increased dependency of T /D with
increasing concentration %.
The 99% mixing time obtained with our corresponding correlation lies about 9% above
the experimental value measured by Distelhoff et al. (1997). Next to the Distelhoff et al.
(1997) mixing time experiment, the experimentally obtained mixing times of Hoogen-
doorn & Den Hartog (1967) and Bouwmans et al. (1997) are compared with the mixing
time obtained from our correlations. The 90% mixing time has been measured at various
Reynolds numbers (up to 104 ) by Hoogendoorn & Den Hartog (1967). At high Reynolds
numbers the mixing time becomes Reynolds number independent and requires about 30
impeller revolutions. The mixing time obtained with our 90% correlation equals 40 im-
peller revolutions, which is an overestimation of about 30%. Bouwmans et al. (1997)
measured a 95% mixing time of 25 impeller revolutions in a Rushton turbine stirred tank
with T /D = 2.5. The value of the 95% mixing time (i.e 29 impeller revolutions) obtained
with our correlation lies 17% above the experimental value of Bouwmans et al. (1997).

5.4.5 The Ruszkowski (1994) mixing time correlation


Ruszkowski (1994) measured the mixing times under turbulent conditions for a Rushton
turbine and other impellers and suggested the following correlation:
 2  2
1/3 T T
N m = 5.3P o = 3.21 (5.12)
D D
Our 90%, 95% and 99% mixing times listed in Table 5.2 have been plotted, together
128 Chapter 5. A parameter study of the mixing time by means of LES

150
90%
95%
120 99%
fit 90%
90 fit 95%
Nqm (-)

fit 99%
60 Ruszkowski
(1994)
30

0
0 1 2 3 4 5
T/D (-)

Figure 5.11: Simulated 90%, 95% and 99% mixing times as a function of T /D, together
with the fitted lines based on equation 5.11. The symbols refer to the cases
mentioned in table 5.1. The thick, solid line resembles the Ruszkowski (1994)
mixing time correlation.

with the fit lines (equation 5.11), as a function of T /D in Figure 5.11. For comparison the
Ruszkowski (1994) correlation (equation 5.12) is shown in the figure. The simulation re-
sults for the 90%, 95% and 99% mixing times significantly deviate from the Ruszkowski
(1994) correlation. If the Ruszkowski (1994) correlation would correspond to e.g. the
90% mixing time, the power number should equal 1 at T /D = 4. This value is signifi-
cantly lower than the range of experimentally measured power numbers by Shiue & Wong
(1984). Based on this remark and the results shown in Figure 5.11, the Ruszkowski (1994)
correlation is likely to correspond to a concentration % lower than 90%.

We will now investigate this possibility in greater detail. The Ruszkowski (1994) cor-
relation is based on a mixing index defined by

v
u 4n+7 
u 1 X ci,rms c 2
IM =1t (5.13)
8 i=4n c

where n is a time index, c is the final concentration and ci,rms is defined by


5.4. Results 129

2 2
10 10
1 3A 1 3A
10 3B 10 3B
0 0
10 3C 10 3C
10
-1 3D -1 3D
10
1-IM (-)

1-IM (-)
-2 -2
10 10
-3 -3
10 10
-4 -4
10 10
-5 -5
10 10
-6 -6
10 10
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Nt (-) Nt (-)

(a) Time window 8 points, in accordance with (b) Time window 400 points, i.e. one blade pas-
equation 5.13. sage period.

Figure 5.12: The value of 1 IM as a function of time for the T /D = 3 cases. The dashed
line represents IM = 0.95.

v
u
M
u 1 X 2
u
ci,rms =t c (5.14)
N j=1 i,j

where ci,j is the concentration at a monitoring point j at time t = it, with t the
sample time in the experiment. The mixing index expresses the root mean square con-
centration fluctuation as a fraction of the mean concentration in the tank after addition of
the tracer. IM can vary from for an infinitely unmixed system to 1.0 for a perfectly
mixed system. The mixing index is calculated for one section with eight points, the section
is moved four data points along the time history and recalculated, and so on. The mixing
time was defined as the time for IM to reach 0.95.
Figure 5.12a shows the simulated time trace of 1IM for the cases 3A, 3B, 3C and 3D.
The value of IM = 0.95 is represented by a dashed line. All four cases show that the mix-
ing time equals approximately 29.5 impeller revolutions (i.e. average over the four cases).
Increasing the width of the time window to a blade passage period (Figure 5.12b) results
in less fluctuations of IM , but it has only a marginal influence on the path of the curves,
and consequently the mixing time. Despite the mixing vessel geometry of Ruszkowski
(1994) was a bit different compared to our system (e.g. dished bottom instead of a flat
130 Chapter 5. A parameter study of the mixing time by means of LES

150
70%
fit 70%
120 Ruszkowski (1994)

90
Nqm (-)

60

30

0
0 1 2 3 4 5
T/D (-)

Figure 5.13: 70% Mixing time as a function of T /D. The Ruszkowski (1994) correlation
and the fitted correlation (i.e. equation 5.11) are shown for comparison. The
symbols refer to the cases mentioned in table 5.1.

bottom, baffle diameter T /12 instead of T /10), the mixing time predicted by our simula-
tions compares very well with the value of 31 impeller revolutions reported by Ruszkowski
(1994).

The mixing time based on IM = 0.95 for the cases 3A, 3B, 3C and 3D equal 32, 30,
24 and 32 impeller revolutions, respectively. With the help of Figure 5.9b, these mixing
times correspond with a coefficient of mixing ranging between 0.12 and 0.22. Figure 5.8a
shows that this range of the coefficient of mixing corresponds with a concentration % of
about 70%. As a result, the definition of the mixing time by Ruszkowski (1994), based on
IM = 0.95, corresponds to the time required for the concentration fluctuations to become
less than 30% of the fully mixed concentration (i.e. the 70% mixing time).

In order to check whether this statement also holds for the T /D = 2 and T /D = 4
cases, the 70% mixing time has been calculated based on the average coefficient of mixing
of 0.17. The results are shown in Figure 5.13. The Ruszkowski (1994) correlation now
fits the data for T /D = 2 and T /D = 3. For T /D = 4 the simulations overestimate
the mixing time by about 15%. The data have been fitted with the correlation given in
equation 5.11, resulting in the fit parameters a0 = 1.92 and a1 = 2.48.
5.4. Results 131

c/c
5

(a) Complete concentration field.

(b) Enlarged view of the concentration (c) Enlarged view of the concentration
field near the impeller. field near the upper left baffle in (a).

Figure 5.14: Performance immersed boundary technique. Snapshot of the concentration


in a horizontal plane at z/T = 0.33 (i.e. disk height) and N t = 15. The dots
represent the reconstructed position of the zero-gradient boundary condition,
the lines represent the walls of the geometry.

5.4.6 Performance immersed boundary technique, mass


conservation
In this section we check the performance of the immersed boundary technique which im-
poses a zero-gradient at walls that are off-grid (see Section 5.3.3). It is, however, not
132 Chapter 5. A parameter study of the mixing time by means of LES

1.1 1.1

1.05 1.05
M/M0 5.

M/M0 5.
1 1
3A
0.95 3B 0.95 2B
3C 3B
3D 4B
0.9 0.9
0 10 20 30 40 50 60 0 20 40 60 80 100
Nt (-) Nt (-)

(a) Cases 3A, 3B, 3C and 3D (b) Cases 2B, 3B and 4B

Figure 5.15: The dimensionless total mass as a function of time. M0.5 is the total mass
after half an impeller revolution (i.e. after the addition of the full amount of
tracer).

expected that a violation of the mass conservation constraint by the immersed boundary
technique will have a significant impact on the mixing time. Possible mass leakage occurs
only at the walls, and a local increase or decrease of mass will not affect the decay rate of
the coefficient of mixing (observed in Figure 5.9), and consequently the mixing time.
In order to check the performance of the immersed boundary technique, the position
of the zero-gradient constraint has been reconstructed with the ghost cell values and two
estimated scalar values in the concentration field (i.e. similar to the routine described in
Section 5.3.3.
Figure 5.14a shows a snapshot of the concentration field at z/T = 0.33 (i.e. disk
height) after 15 impeller revolutions. The reconstructed zero-gradient positions are repre-
sented by (overlapped) dots. Figure 5.14b shows an enlarged view of the impeller region.
The dots are clearly identified. The lines represent the walls of the impeller geometry.
The dots and the lines overlap, which means that the zero-gradient boundary constraint
coincides with the geometry walls, as should be expected. The same conclusion is drawn
based on Figure 5.14c that shows an enlarged view near the baffle.
The total mass in the system has been monitored in order to check the mass conserva-
tion constraint. The total mass, M , is calculated as follows:
5.4. Results 133

Vtot
X
M= ci Vi (5.15)
i=1

where Vtot is the amount of control volumes in the flow domain. The dimensionless
total mass (with M0.5 the total mass after half an impeller revolution, i.e. after stopping
the injection) as a function of the dimensionless time for the cases with the T /D = 3
impeller is shown in Figure 5.15a. The time traces of the total mass show after some time
a more or less linear increase of the total mass. At the end of the simulations, the total
mass increase has increased by 5% 8%.
The slope of the mass increase depends on the impeller size as can be seen in Fig-
ure 5.15b. An increase of impeller size results in a steeper slope of the mass as a function
of time. This observation implies that mass conservation at the impeller blades is a prob-
lem; larger blades result in a larger mass increase.
The increase of mass may be attributed to the blade thickness (i.e. three lattice spac-
ings). The blades are thin with respect to the resolution of the grid, and consequently the
blade edges are not accurately represented by the immersed boundary technique. As a re-
sult, the blade edges may be considered as the main sources of mass leakage. Furthermore,
increasing the impeller size results in a larger surface of the impeller blade edges, which
explains the steeper increase of the total mass observed in Figure 5.15b. Also the edges of
the baffles and impeller disk may contribute to the mass increase, as their edges have the
same thickness as the impeller blade edge.
That the blade and baffle thickness with respect to the grid resolution is a problem in
conserving the total mass is observed in the time traces shown in Figure 5.15a. The mass
increase of case 3D starts in the early stage of the mixing process, because of the vicinity
of a baffle. In case 3C the injection point is in the impeller stream, where the concentration
is strongly advected radially outward toward the baffles. In case 3B the injection point is
0.1T above the impeller and the mass increase starts at roughly N t = 2. In case 3A, the
mass injection point is at the top of the tank, and it takes some time for the concentration
being advected toward the impeller region. As a result, the mass increase is expected at a
later moment in time (i.e. N t = 10) compared to the other cases (see e.g. the instantaneous
concentration fields in Figure 5.3).
In summary, the immersed boundary technique developed positions accurately the
zero-gradient constraint at the geometry walls (Figure 5.14), but the mass conservation
constraint is not satisfied (Figure 5.15). The mass increase is significant (5% 8%), and
in future research the current technique needs to be improved. It is expected that an in-
crease of the grid resolution or the usage of a cylindrical grid in the scalar mixing solver
134 Chapter 5. A parameter study of the mixing time by means of LES

will result in a better approximation of the mass conservation.

5.5 Conclusions
In this chapter, we have investigated the influence of injection position and impeller size
on the mixing time in a Rushton turbine stirred tank by means of a LES including scalar
mixing. In literature, there is no standardization of a mixing time experiment. Mixing
times are reported for tanks of different layouts and operating conditions, tracer injection
positions and injection speeds, number and positions of measurement points. Furthermore,
there is no agreement on the definition of the mixing time. As a result, a measured and/or
simulated mixing time should be compared with great care as is done in this work. This
also applies to mixing time correlations in the literature.
We have coupled the mixing time to a coefficient of mixing, that provides a concentra-
tion variance of points in different regions in the tank. Simulated time traces of the coeffi-
cient of mixing revealed after some impeller revolutions an exponential decay. At the end
of the calculations, the coefficient of mixing stabilized at about 0.02. This is unphysical,
and this effect is attributed to numerical errors, induced by e.g. the compressibility effect
of the schemes used or the scalar concentration errors introduced at the walls by the im-
mersed boundary technique. The exponential decay has been extrapolated through a fitting
procedure. The mixing times were defined by the intersection of the fitted lines and the
calibrated values of the 90%, 95% and 99% concentration criterions, respectively.
The mixing time is significantly influenced by the impeller size. The simulated results
of the 90%, 95% and 99% mixing times show an impeller size dependency of typically
(T /D)2.5 . The simulated mixing times agree within 30% with experimentally obtained
values. Furthermore, the simulated 99% mixing times compared very well with the Sano
& Usui (1985) mixing time correlation, within an overestimation up to 30%. Our work
has indicated that the Ruszkowski (1994) mixing time correlation corresponds to the 70%
concentration criterion.
The position of the tracer injection point does not significantly affect the mixing time.
While a spread of the simulated mixing time was observed in the T /D = 3 cases, it was
found to be clearly less significant in the T /D = 2 and T /D = 4 cases. The predicted
99% mixing time in the case similar to the experiment of Distelhoff et al. (1997) overesti-
mates the experimentally obtained mixing time with 26%.
The scalar transport has been solved with the finite volume approach. In contrast with a
lattice-Boltzmann discretization of the convection diffusion equation, the former approach
yields a less memory intensive scheme. False diffusion effects in the finite volume scheme
5.5. Conclusions 135

are minimized through the use of a high-resolution TVD scheme. Furthermore, an im-
mersed boundary technique is developed that imposes a zero-gradient boundary constraint
for the scalar at the walls that are off-grid.
Our novel immersed boundary technique does not guarantee mass conservation. The
average mass increase per impeller revolution depends on the impeller size being 0.25%,
0.125% and 0.05% for the impeller sizes T /D = 2, T /D = 3 and T /D = 4, respectively.
The unphysical mass increase is attributed to the mass leakage at the edges of the impeller
blades, impeller disk and baffles as they are not accurately represented by the grid reso-
lution. It is expected that (local) grid refinement or the usage of a cylindrical grid in the
scalar mixing solver will result in less leakage of mass.
The coupled LES and finite volume scalar mixing simulations have provided a detailed
insight in the scalar mixing characteristics of the stirred tank. The detailed simulations
presented in this work enabled a careful comparison of the simulated mixing times with
those experimentally and/or numerically measured, and with those obtained via correla-
tions in the literature. Furthermore, this research opens a promising perspective for the
coupling of the developed code to a particle transport code developed by Derksen (2003)
in order to study either crystallization or dissolution processes.
136 Chapter 5. A parameter study of the mixing time by means of LES

Nomenclature

Roman Description Unit


a0 ,a1 fit parameters of the mixing time correlation in equa- -
tion 5.11
A surface area m2
A+ constant in the Van Driest (1956) wall damping func- -
tion
c concentration kg.m3
ci,rms the rms concentration at time it -
cmix coefficient of mixing -
c spatial average of the concentration kg.m3
c final concentration kg.m3
cs Smagorinsky constant -
D impeller diameter m
FC convective prefactor in equation 5.2 kg.s1
FD diffusive prefactor in equation 5.2 kg.s1
H height of the tank m
IM mixing index defined by Ruszkowski (1994) -
K number of monitoring points -
L value of repetition of experiment described in sec- -
tion 5.4.3
n time index -
nb number of impeller blades -
N impeller speed s1
M total mass kg
M0.5 injected mass after half an impeller revolution kg
Po Power number -
r radial coordinate m
r upwind ratio of consecutive gradients -
rp radius of the feed pipe m
Re Reynolds number -
S resolved deformation rate s1
S source term in convection-diffusion equation kg 2 .m6 .s1
Sc molecular Schmidt number -
Sct turbulent Schmidt number -
5.5. Conclusions 137

t time s
T tank diameter m
u velocity m.s1
ui velocity component i m.s1
Vtot amount of control volumes in flow domain m3
w height of the impeller blade m
xi coordinate i m
y+ distance from wall in viscous wall units -
z axial coordinate m

Greek Description Unit


diffusion coefficient m2 .s1
e eddy diffusion coefficient m2 .s1
mol molecular diffusion coefficient m2 .s1
lattice spacing -
t time step s
V finite volume m3
o
angle
m mixing time s
mix mixing length m
kinematic viscosity m2 .s1
e Smagorinsky eddy viscosity m2 .s1
density kg.m3
general scalar variable -
flux limiter -
6 Numerical simulation of a
dissolution process in a stirred
tank reactor

A dissolution process of solid particles suspended in a turbulent flow of a


Rushton turbine stirred tank is studied numerically by large eddy simula-
tions including passive scalar transport and particle tracking. The lattice-
Boltzmann flow solver and the Smagorinsky subgrid-scale model have been
adopted for solving the stirred tank flow. To the large eddy simulation (LES), a
finite volume scheme is coupled that solves the convection-diffusion equation
for the solute. The solid particles are tracked in the Eulerian flow field through
solving their dynamic equations of linear and rotational motion. Particle-
particle and particle-wall collisions are included, and the particle transport
and fluid flow are two-way coupled. The simulation has been restricted to a
lab-scale tank with a volume equal to 102 m3 . A set of 7 106 spherical
particles 0.3 mm in diameter is released in the top part of the tank, resulting
in a local initial solids volume fraction of 10%. The particle properties are
such that they resemble calcium-chloride beads. The focus is on solids and
scalar concentration distributions, particle size distributions, and the dissolu-
tion time. For the particular process considered, the dissolution time is found
to be at most one order of magnitude larger than the time needed to fully
disperse the solids throughout the tank.

Key words: stirred tank, turbulence, dissolution, suspension, simulation, par-


ticle transport

Parts of this chapter have been submitted to Chem. Eng. Sci.


140 Chapter 6. Numerical simulation of a dissolution process

6.1 Introduction
Processes in which turbulently agitated solid-liquid suspensions are involved, have a large
share in various industrial applications. Examples are crystallization (Hollander et al.,
2001), suspension polymerization (initially a liquid-liquid dispersion that in the course of
the process turns into a solid-liquid mixture, see e.g. Kiparissides (1996)), particle coating
and catalytic slurries. Such processes are very complex multi-phase. In order to improve
competitiveness, there is an industrial drive for research on the hydrodynamic phenomena,
and their coupling with chemistry and heat and mass transfer.
One of the key aspects in the (dynamic) behavior of the processes mentioned is the
role of hydrodynamics. On a macroscopic scale, the hydrodynamic conditions control e.g.
residence time and circulation time in the specific flow system (e.g., a stirred tank). On
a microscopic scale, rate-limiting processes such as mass transfer (needed for nucleation
and growth in crystallization), agglomeration and attrition (a major source for secondary
nucleation), collisions, and the yield of a chemical product (in case of competitive chemi-
cal reactions) are largely determined by the smallest-scale flow phenomena. It is difficult
to determine a priori the impact of the smallest-scale flow phenomena on the overall out-
come of the particular process (e.g. the layer thickness in a particle coating process, the
final particle size distribution in a crystallization or suspension polymerization process,
dissolution time of salts in a dissolution process).
Next to the role of hydrodynamics, there are many unresolved issues with respect to
what is actually going on at the particle scale. One may think of heat and/or mass transfer
between the particle and the continuous phase, mechanical load on the particle (which
determines the occurrence of an attrition event) as a result of particle-particle or particle-
impeller collisions, and the influence of the presence of particles on the local and global
flow features in the flow system. Furthermore, physical mechanisms are often controlled
by non-linearities. For instance, in crystallization processes the collision rate depends on
the square (i.e. non-linear) of the particle number concentration (Hollander et al., 2001),
and in chemical reactive flows the reaction rate constants depend in a non-linear way on the
concentrations of the reactants. Next to various experimental and theoretical investigations
reported in the literature, numerical modeling is an alternative route to investigate the
issues mentioned.
In this work, the focus is on a dissolution process under strongly turbulent conditions
induced by a Rushton turbine in a stirred tank. Because of limited computational re-
sources, the strongly turbulent flow cannot be fully captured by the computational grid.
Consequently, direct solving the flow system under consideration goes beyond the present
and future foreseeable computational possibilities. As a result, we need to revert to model-
6.1. Introduction 141

ing. The first step in realistic modeling of solids suspension in turbulently stirred tanks is
an accurate representation of the single-phase flow field. An approach widely used today
for solving the turbulent flow is based on the solution of the Reynolds-averaged Navier-
Stokes (RANS) equations. Although various investigations (Ng et al., 1998; Hartmann
et al., 2004b) have shown that a simulation based on a RANS approach is able to rea-
sonably represent the average flow field, the turbulence levels are underpredicted by 50%.
This certainly affects the mixing patterns and consequently solid-liquid mixing.

As stated above, the motion of particles, mass transfer, and collisions are determined
by the small-scale flow phenomena, and consequently small-scale flow information com-
parable to the particle size is needed. Therefore, the basis of the simulation discussed here
is a representation of the continuous flow by means of large eddy simulation (LES). Com-
pared to a RANS approach, turbulence modeling by means of LES leaves less room for
speculation in modeling the turbulence and the motion of the solids immersed in the flow.
Various numerical investigations have demonstrated that LES can accurately represent the
single-phase flow in a stirred tank, including the turbulence levels (Eggels, 1996; Derksen
& Van den Akker, 1999; Hartmann et al., 2004a,b; Yeoh et al., 2004; Bakker & Oshinowo,
2004).

The confidence achieved in using the flow solver based on the LES methodology has set
a solid basis for setting up a simulation on solid-liquid processes including mass transport,
which is the focus of this work. In the LES flow field, spherical particles are released in
the top part of the tank. The motion of the particles and collisions are handled by a solver
developed by Derksen (2003). For the inter-phase mass transfer between the disperse and
continuous phases, a single-particle correlation of Ranz & Marshall (1952) is applied.
Information on the local scalar concentration field, needed for the determination of the
mass transfer rate, is obtained by solving the convection-diffusion equation for the species
involved. Exactly as in a LES, the latter equation is solved in an Eulerian framework
through a finite volume discretization. The finite volume scheme used in this work has
been described in Chapter 5.

The aim of this work is to provide insight in the physical mechanisms occurring in
a dissolution process by means of detailed LES including passive scalar transport and
particle tracking. In this work, we focus on the different stages occurring in the dissolution
process. Furthermore, we pay attention to the evolution of the particle size distribution
in time. Through monitoring the total amount of particles in the suspension in time, a
dissolution time is determined.
142 Chapter 6. Numerical simulation of a dissolution process

T/10
D = T/3
3D/4

H=T 0.04D D/5


0.04D
0.16D
T/3
D/4
T

Figure 6.1: Cross-section of the tank (left). Plan view and cross-section of the impeller
(right). At the top level there is a lid. The impeller is a Rushton turbine
mounted at height T /3 and has a diameter T /3.

6.2 Flow system


The stirred tank used in this work was a standard configuration cylindrical tank of di-
ameter T , with four equi-spaced baffles of width 0.1T mounted along the perimeter of
the tank. The liquid height was set equal to the tank diameter, H = T . The impeller
was a six-bladed Rushton turbine with standard dimensions, mounted at height T /3. A
schematic representation of the flow system is shown in Figure 6.1. If geometric similarity
is maintained, the single-phase flow can be fully characterized by the Reynolds number
(Re=N D 2 /, with N the impeller speed, D the impeller diameter and the viscosity of
the continuous phase). In this work, the Reynolds number amounts to 105 .
The tank volume (V ) is set to 102 m3 , which implies an impeller diameter of D =
7.78 102 m. The continuous phase is water (with viscosity = 106 m2 /s and density
l = 103 kg/m3 ). A set of Np0 = 7 106 mono-disperse spherical particles with diameter
dp0 = 0.3 mm, and density p = 2150 kg/m3 is released uniformly distributed over in
the upper part (0.9T T ) of the tank. The saturation concentration (csat ) is set to 600
kg/m3 , and the molecular diffusion coefficient (mol ) equals 0.7 109 m2 /s, which yields
a Schmidt number of about 1400. The settings of the particle diameter, the saturation
concentration and the diffusion coefficient are typical for calcium-chloride beads in water.
Since calcium ions are larger than chloride ions, they have a lower diffusion coefficient.
As a result, the diffusion of calcium ions is rate-limiting, and therefore the diffusion coeffi-
cient in the simulation is that of calcium ions. The initial solids volume and mass fractions
of the set of particles released in the top part of the tank amount to 10% and 21.5%, re-
spectively (the tank average volume and mass fractions are 1% and 2.15%, respectively).
6.3. Simulation procedure 143

The initial macroscopic Stokes number (i.e. the ratio of the Stokesian particle relaxation
2
d N
p0
time and the time of one impeller revolution; Stk= pl 18 ) yields Stk=0.1774. The initial
microscopic Stokes number which relates the Stokesian particle relaxation time to the time
scale of the smallest turbulent fluctuations (i.e. the Kolmogorov time scale based on the
tank-averaged energy dissipation rate) equals 600. As a result, the particles will initially
follow the large-scale turbulent fluctuations. During the course of the simulation, the par-
ticle size decreases and the motion of the particle will be influenced by an increasing part
of the turbulence spectrum (i.e. the particle will behave more and more as a tracer).
The impeller speed is chosen to be above the (initial) just-suspended impeller speed
(Njs ) according to the Zwietering (1958) correlation:
0.45
d0.2
p0
0.1
(g) 0.13
m0
Njs = s (6.1)
0.45
l D0.85
with g the magnitude of the gravitational acceleration vector, = p l , m0
the initial solids mass fraction in %, and s is a constant that equals 8 for the particular
configuration. The impeller speed is set to 16.5 rev/s, whereas the (initial) just-suspended
speed is 11.4 rev/s. Since the particles dissolve during the course of the simulation, particle
size and the solids mass fraction decreases in time, and consequently the just-suspended
speed decreases.

6.3 Simulation procedure


6.3.1 Flow solver
The Reynolds number presented in the previous section (Re=105 ) implies a strongly tur-
bulent flow that cannot be fully resolved by the computational grid. As a result, turbulence
modeling is required. We apply large eddy simulation (LES), in which the small scales in
the flow are assumed to be universal and isotropic. The flow resolved by the grid can be
interpreted as a low-pass filtered representation of the true flow. The effect the small scales
(i.e. the part that has been filtered out) have on the larger scales is modeled with a subgrid-
scale model. The LES-model applied here is the Smagorinsky (1963) subgrid-scale model,
where the action of the subgrid-scale motion is considered to be purely diffusive through
an eddy viscosity. For more details on the LES methodology used we refer to Hartmann
et al. (2004a,b) and Derksen & Van den Akker (1999).
A lattice-Boltzmann method (Chen & Doolen, 1998) was used for solving the filtered
momentum equations. The specific scheme we used was introduced by Somers (1993),
144 Chapter 6. Numerical simulation of a dissolution process

and is described by Derksen & Van den Akker (1999).


The entire tank was simulated on a uniform, cubic computational grid. Inside the com-
putational domain, the no-slip boundary conditions at the cylindrical tank wall, the baffles,
the impeller, and the impeller shaft were imposed by an adaptive force-field technique
(Derksen & Van den Akker, 1999).

6.3.2 Particle transport solver


The particle transport solver used for the simulation of the dissolution process makes use
of the Eulerian-Lagrangian approach and has been developed by Derksen (2003). Next to
the solution of the solid particles dynamics, particle-wall, particle-impeller and particle-
particle collisions are considered in the solver. The latter mechanism proved to be crucial
in order to obtain a realistic particle distribution throughout the tank (Derksen, 2003). This
improvement was primarily caused by the exclusion effect brought about by the collision
algorithm. We will describe here the global features of the solver, the details can be found
in the paper of Derksen (2003).
The motion of the particles is controlled by the drag, added mass, gravity, lift and stress
gradients forces. The inclusion of the lift force requires knowledge of the vorticity of the
fluid and the particle rotation. The latter is obtained from the solution of the dynamic
equation for the particle angular velocity. The time step in the discrete version of the
dynamic equations equals the time step in the LES.
The LES provides the resolved part of the fluid motion. Solid particles released in the
LES flow field, will also feel the subgrid-scale part of the velocity fluctuations. Derksen
(2003) has shown that for the particular case studied here, the resolved flow dominates
the particle motion. In the first place, the time step is able to keep up with the particle
relaxation time for a large range of particle diameters (i.e 0.05 dp /dp0 1). In the
second place, the resolved velocity fluctuations are an order of magnitude larger than the
(estimated) subgrid-scale fluctuations (Derksen, 2003). Only for the determination of the
drag force, a subgrid-scale part has been included.
The initial solids volume fractions are such that two-way coupling effects may be im-
portant (Elgobashi, 1994). As a result, they have been included in the solver.
In the particle collision algorithm, particle-wall, particle-impeller and particle-particle
collisions are considered to be fully elastic and frictionless. The efficiency of the collision
algorithm has been improved by limiting the computational effort in handling particle-
particle collisions. This has been achieved by grouping the particles in a so-called link-list
(see Chen et al., 1998; Derksen, 2003). A proper functioning of the collision algorithm is
limited by the time step, or the particle volume fraction. Derksen (2003) has indicated that
6.3. Simulation procedure 145

the number of missed collisions increases with increasing solids loading. The increase is
most pronounced in highly turbulent regions (i.e. in the impeller region). As in our case
the initial solids volume fraction equals 10% in the top part of the tank, the number of
missed collisions may not be significant compared with the number of correctly handled
collisions. In the first place, the turbulence intensity is low in the top part of the tank.
Secondly, the local solids volume fraction decreases rapidly as the particles are dispersed
throughout the tank.

6.3.3 Mass transfer


The particles lose/gain weight due to a continuous process of inter-phase mass transfer be-
tween the particles and the liquid (i.e. the continuous phase) or vice-versa. Mass transfer
is controlled by a mass transfer coefficient (k) and a driving force, which is the differ-
ence between the mass saturation concentration (csat ) and the mass concentration of the
surroundings (c):

mint = kA (csat c) (6.2)

where mint is the inter-phase mass transfer rate between the particle and liquid, and
A = d2p is the surface of the spherical particle. The mass transfer coefficient is defined
as:

mol
k = Sh (6.3)
dp
where Sh is the Sherwood number.
The concentration of the surroundings in equation 6.2 is assumed to be the average
concentration in the control volume of the scalar transport solver. The effect of concentra-
tion fluctuations at the subgrid-scale level on the mass transfer rate has been ignored, since
the mass transfer rate linearly scales with the concentration (see equation 6.2), and the ef-
fect of fluctuations tends to average out. The cell average concentration is determined via
the discrete solution of the convection-diffusion equation. This is achieved by coupling a
scalar mixing solver (described in the next subsection) to the LES and particle transport
solvers.
Mass transfer between solid particles and continuous phase depends on the motion
of the solid surface relative to the liquid. Apart from linear velocities, particle rotation
might play a role in mass exchange. Derksen (2003) has shown the spatial distribution
of linear and rotational slip velocities in terms of the respective Reynolds numbers. For
146 Chapter 6. Numerical simulation of a dissolution process

particles with a diameter equal to 0.3 mm, the rotational Reynolds numbers are one order
of magnitude smaller than translational Reynolds numbers. As a result, it is expected that
mass transfer is likely to be controlled by translation of the particles, and the contribution
of rotation to the overall mass transfer rate is small and can be neglected. In this work, the
Ranz & Marshall (1952) correlation is used for the calculation of the Sherwood number:
1 1
Sh = 2.0 + 0.6Rep2 Sc 3 (6.4)

where Rep is the particle Reynolds number, defined as:

vsl dp
Rep = (6.5)

where vsl is the slip velocity (i.e. the magnitude of the velocity of the particle relative
to the fluid).

6.3.4 Scalar mixing solver


In the simulation of passive scalar transport, we assume one-way interaction with the flow
field, i.e. the instantaneous velocity affects the scalar evolution, while the passive scalar
does not change the flow characteristics. In this way, the scalar field is mathematically
decoupled from the dynamical equations that govern the flow field. Thus, the solution of
the flow field is a prerequisite to the solution of the scalar field.
Eggels & Somers (1995) have performed scalar transport calculations on free convec-
tion cavity flow with the lattice-Boltzmann discretization scheme. This scheme, however,
is more memory intensive than a finite volume formulation of the convection-diffusion
equation. Therefore, we have coupled a compressible finite volume discretization scheme
for the scalar transport to the lattice-Boltzmann flow solver. The convection-diffusion
equation in compressible form reads
 
l c ui l c c
+ = l + Sc (6.6)
t xi xi xi
where l is the density of the continuous phase, c is the scalar concentration, ui is
velocity component i, is the diffusion coefficient and Sc is a source term. Note the sum-
mation over the repeated index i. The diffusion coefficient is the sum of the molecular and
eddy diffusion coefficients: = mol +e . The eddy diffusion coefficient is related to the
eddy viscosity via a turbulent Schmidt number (Sct ) through e = e /Sct . For the turbu-
lent Schmidt number we have chosen a value of 0.7. Since the lattice-Boltzmann scheme
6.3. Simulation procedure 147

is a compressible scheme, we have implemented the discretized form of the compress-


ible convection-diffusion equation. In this context it should be noted that the maximum
velocity (which is approximately the tip speed) is set sufficiently low for meeting the in-
compressibility limit in the lattice-Boltzmann scheme. The finite volume scheme used in
the context of this work has been described in detail in Chapters 4 and 5.
The source term Sc in equation 6.6 describes the mass transfer between the particles
and the continuous phase. The discretized source term Sc,i reads

Np,i
X
Sc,i = l V 1 mint,j (6.7)
j=1

where V is the volume of cell i (which is the same for all cells in the uniform, cubic
grid), Np,i is the number of particles with their center of mass in cell i and mint,j is the
inter-phase mass transfer rate (see equation 6.2) between particle j and the surrounding
continuous phase. The diameter of particle j is adjusted each time step as follows:
r
3 6 1
dn+1 dnp,j
3
p,j = p mint,j t (6.8)

where t is the time step.
The zero-gradient constraint for the scalar concentration at the off-grid (moving) walls
has been imposed by the novel immersed boundary technique described in sections 4.6
and 5.3.2.

6.3.5 Simulation aspects


The particles are released in a quasi steady-state flow field. The scalar concentration at
the start of the simulation equals zero throughout the tank. The set-up of the dissolution
process has been described in section 6.2. The code runs on the parallel computer platform
Aster located at SARA in Amsterdam. Aster is an SGI Altrix 3700 system, consisting of
416 CPUs (Intel Itanium 2, 1.3 GHz, 3 Mbyte cache each), 832 Gbyte of memory and 2.8
Tbyte of scratch disk space. The total peak performance is 2.2 Teraflop/sec. Every node in
Aster is a CC-NUMA machine (Cache-Coherent Non Uniform Memory Access). All pro-
cessors are integrated into one machine, which significantly speeds-up the communication
between the processors.
Within the parallel environment of Aster the code makes use of domain decomposi-
tion: the computational domain was horizontally split in 30 equally-sized subdomains (i.e.
normal to the tank centerline). The presence of particles in a stirred tank significantly com-
148 Chapter 6. Numerical simulation of a dissolution process

plicates the parallelization of the computer code. In the first place, load-balancing needs to
be reconsidered since the computational load of a sub-domain now depends on the number
of particles inside the sub-domain in a quite unpredictable manner. Furthermore, there is
much more communication between subdomains compared to a single-phase simulation,
as particles cross sub-domain borders. Dynamic load-balancing was not implemented in
the code. This initially results in a poor parallel efficiency and a low calculation speed,
as all particles are located in the upper subdomains. Gradually, the calculation speed
increases as the particles travel toward the impeller region and become more or less homo-
geneously distributed throughout the tank (bearing in mind that particles dissolve during
simulation, and are consequently more easily resuspended from the bottom). The MPI
message passing tool has been used for communication between the subdomains.
The simulation has been executed on a cubic, Cartesian grid of 2403 lattice cells.
The diameter of the tank equals 240 lattice spacings, resulting in a spatial resolution of
T /240 = 0.973 mm. The temporal resolution is limited by the lattice-Boltzmann method.
The tip speed of the impeller is set to 0.1 lattice spacings per time step in order to meet the
incompressibility limit. This results in a temporal resolution of 1/(2400N ) = 25 s.
For the LES, 21 (18 directions and 3 force components) single-precision, real values
need to be stored. For the scalar transport, 3 (concentration fields at time instants n + 1, n
and n1) double-precision, real values need to be stored. Finally for the particle properties
a memory space for 16 (i.e. particle properties; particle coordinates, velocity and angular
velocity components, diameter, etc) times the number of particles (7106 ) double precision,
real values is needed. In total, the memory requirements of the simulation result in an
executable of about 2.5 GByte. The full simulation of 100 impeller revolutions took about
6 weeks.

6.4 Results
6.4.1 Snapshots of the particle distributions and
concentration fields
Figures 6.2 and 6.3 give an impression of the dissolution process during the course of
the simulation. The scalar concentrations are normalized with the final concentration c
defined as:
 3
Mp0 2 dp0
c = = Np0 p (6.9)
V 3 T
6.4. Results 149

Nt=2 Nt=5

Nt=7 Nt=10

Nt=20 Nt=40

Figure 6.2: Snapshots of the particle distribution in a vertical plane mid-way between two
baffles. In the graphs, the particles in a slice with thickness T /240 have been
displayed. The diameter of the particles is 3 times enlarged for clarity. The
respective concentration distributions are shown in Figure 6.3.
150 Chapter 6. Numerical simulation of a dissolution process

Nt=2 Nt=5

c/c
Nt=7 Nt=10 1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
Nt=20 Nt=40 0.2

Figure 6.3: Snapshots of the concentration distribution in a vertical plane mid-way be-
tween two baffles. The respective particle distributions are shown in Figure 6.2.
6.4. Results 151

where Mp0 = Np0 p 6 d3p0 is the total particle mass at N t = 0.


During the first five impeller revolutions, the particles are transported toward the im-
peller region. High particle concentrations are accompanied by high scalar concentrations,
and a clear interaction between the particles and turbulence is identified. The particles are
swept radially outward by the revolving impeller, and are subsequently entrained in the
lower recirculation loop. In the next 5 15 impeller revolutions, the particles are dis-
tributed throughout the tank. During the first 10 20 impeller revolutions (till the point
that the particles are distributed throughout the tank), clear macroscopic scalar concentra-
tion structures are identified. The scalar transport matches the particle transport.
As the dissolution process proceeds, a number of features are identified from the snap-
shots shown in Figure 6.2. In the first place, the snapshots of the particle distributions at
N t = 10 and N t = 20 show high particle concentrations at the bottom and outer walls.
Secondly, the particles are organized in streaky patterns. Because of the particle inertia
and the density of the particles being higher than that of the liquid, the particles are swept
out of the high-vorticity regions. Thirdly, the size of the particles becomes smaller (com-
pare the snapshots at N t = 20 and N t = 40). As the solid particle mass decreases during
the process, the particle inertia decreases. As a result, the streaky patterns die out. Fur-
thermore, smaller particles are more easily picked up from the bottom and resuspended
in the flow. The high particle concentration at the bottom, observed in the snapshots at
N t = 10 and N t = 20, has disappeared in the snapshot at N t = 40. A decrease of
the particle mass, accompanied by a decrease of the particle inertia and resuspension of
particles, results in the particle distribution to become more and more homogeneous.
From the point the particles are distributed throughout the tank, the macroscopic struc-
tures in the scalar concentration have disappeared (see Figure 6.3, the snapshots at N t =
20 and N t = 40). A decrease of the particle mass observed in the snapshots in Figure 6.2
is associated with an increase of scalar concentration. At the wall separation points and at
the bottom of the tank spots of high concentration are observed that correspond with local
high solids concentrations.
In Figure 6.4 two snapshots of the particle distribution at N t = 26.5 and N t = 60
are shown together with the overall particle size distributions. The particles are 10 times
enlarged and colored by their size. The particle distribution at N t = 26.5 shows an axial
gradient in the average particle size. The particles transported in the lower recirculation
zones are generally smaller than the particles transported in the upper recirculation zone.
The reason for this effect is that compared with the upper recirculation loop, the lower
recirculation loop is stronger and this induces an increased mass transfer rate. A compari-
son between the upper and lower graphs clearly shows the difference in particle size. The
spatial particle distribution at N t = 60 is nearly homogeneous.
152 Chapter 6. Numerical simulation of a dissolution process

0.01
dp /dp0
0.0075
Np /Np0(-)

0.72
0.005 0.7
0.0025
0.66
0
0 0.25 0.5 0.75 1
dp /dp0 (-) 0.62

0.58

0.54

0.5
Nt = 26.5
0.01
dp /dp0
0.0075
Np /Np0(-)

0.5
0.005
0.45
0.0025 0.4
0 0.35
0 0.25 0.5 0.75 1
dp /dp0 (-) 0.3
0.25
0.2
0.15
0.1
0.05
Nt = 60
Figure 6.4: Snapshots of the particle distribution in a vertical mid-way baffle plane accom-
panied by the respective overall particle size distributions. In the graphs, the
particles in a slice with thickness T /240 have been displayed. The diameter
of the particles is 10 times enlarged for clarity, and the colors represent the
dimensionless particle diameter.
6.4. Results 153

dp /dp0
0.82

0.79

0.76

0.73

0.7

0.67

c/c
0.8

0.72

0.64

0.56

0.48

0.4

Figure 6.5: Snapshot of the particle distribution at N t = 15 (upper graph) in a horizontal


plane at disk height accompanied by the respective concentration distribution
(lower graph). In the upper graph, the particles in a slice with thickness T /240
have been displayed. The diameter of the particles is 10 times enlarged for
clarity, and the colors represent the dimensionless particle diameter.
154 Chapter 6. Numerical simulation of a dissolution process

The snapshot of the particle distribution at N t = 26.5 shows a region void of particles
extending from the bottom to closely underneath the impeller, slightly right from the tank
centerline. In this region there is a manifestation of a slowly precessing vortex that crosses
the cross-section at N t = 26.5. In this region of high vorticity, the particles having a larger
density than that of the liquid are swept out due to their inertia. This precessing vortex (a
so-called macro-instability), here in interaction with the particles, was studied in detail in
Chapter 3 (Hartmann et al., 2004a).
Figure 6.5 shows instantaneous realizations of the particle distribution and the scalar
concentration in a horizontal plane at disk height at N t = 15. In the horizontal cross-
section of the particle distribution, large voids behind, and high solids concentrations in
front of the blades are observed (in accordance with the simulation results of Derksen
(2003)). The streaks of particles keep their identity of quite a long radial distance. Also
high solids concentrations near the walls are observed, as a result of the centrifugal forces.
Spots of high concentration are observed near the walls that are associated with locally
high solids concentration. Furthermore, the scalar concentration distribution shows in-
creased concentration levels in front of the blades as a result of the high solids concentra-
tions.

6.4.2 Stages in the dissolution process


Figure 6.6a shows time-traces of the number of particles in ten axial slices of height 0.1T ,
with a focus on the top and bottom slices. From the start of the process, the particle
number in the upper slice decreases rapidly as the particles are transported toward the
impeller region. After approximately 10 impeller revolutions, the first particles reach the
bottom of the tank. Because of a density ratio of the disperse and the continuous phases
being larger than unity, the largest number of particles is observed in the bottom slice for
N t > 15. Based on the observations in the previous subsection and the path of the curves
shown in Figure 6.6a, five stages can be identified:

Stage I (0 N t < 12): Mixing and dispersing

Stage II (12 N t < 24): Quasi steady-state

Stage III (24 N t < 42): Resuspension

Stage IV (N t 42): Dissolution

Stage V (N t 58): Homogeneous suspension


6.4. Results 155

1
I II III IV
0.8
Np /Np0 (-)

0.6 0.9T - T

0.4
0 - 0.1T
0.2 V
0
0 20 40 60 80 100
Nt (-)

(a)

1
a = 0.9T - T
b = 0.8T - 0.9T
0.8 c = 0.7T - 0.8T
a d = 0.6T - 0.7T
Np /Np0 (-)

0.6 e = 0.5T - 0.6T


f = 0.4T - 0.5T
g = 0.3T - 0.4T
0.4 h = 0.2T - 0.3T
b i = 0.1T - 0.2T
j = 0 - 0.1T
0.2
c d j
e f
g h i
0
0 3 6 9 12 15
Nt (-)

(b)

Figure 6.6: Evolution of particle number in ten axial slices of height 0.1T . In (a) the fo-
cus is on the top and bottom slices. Five stages in the dissolution process can
be defined: Mixing and dispersing (I), Quasi-steady-state (II), Resuspension
(III), and Dissolution (IV). At time instant N t = 58 a more or less homoge-
neous suspension is reached (Stage V). In (b) the focus is on the first stage (i.e.
mixing and dispersing).
156 Chapter 6. Numerical simulation of a dissolution process

In stage I (Mixing and dispersing), the solids are distributed throughout the tank till a
quasi steady-state situation is reached (stage II). During stage II, the dissolution process
proceeds, but the number of particles in all slices remains approximately constant. The
number of particles in the bottom slice is about 1.6 2.3 times higher than the number of
particles in the other slices. At N t = 24 the number of particles in the bottom slice starts
to decrease significantly (stage III). At this point in time, the mass of the particles residing
at the bottom has decreased to a level where the turbulent motions are strong enough
to resuspend the particles. After about 42 impeller revolutions, the first particle is fully
dissolved (the onset of stage IV). Stage V is a part of stage IV and starts at the time instant
N t = 58 when a fully homogeneous suspension is reached. Based on Figure 6.6a it can be
concluded that (for the particular case studied here), the time to dissolve all particles (i.e.
the dissolution time) is at most one order of magnitude larger than the time to distribute
the particles throughout the tank.
Figure 6.6b presents an enlarged view of the first stage. The number of particles nor-
malized with the total amount of particles in slice a (i.e. the top slice) decreases rapidly
from 100% to about 11%. Subsequently, the normalized number of particles in slices b to
i increase successively to 8% 11%. The normalized number of particles in the bottom
slice (j) is significantly higher (i.e. 18%) representing the high solids concentration at the
bottom.

6.4.3 Evolution of particle size distribution in time


In Figure 6.7 the particle size distributions are shown at nine instants in time. The first
six distributions correspond to the particle distributions shown in Figure 6.2. During the
course of the simulation, the particle size distribution broadens and shifts toward smaller
particle sizes. At N t = 60, 80, 100 a peak is observed in the first bin, representing the
number of dissolved particles (i.e. 16.8%, 97.3%, 99.9% of the total amount of particles
released).
A peculiar feature of the particle size distributions at N t = 60, N t = 80 and N t = 100
is its steep slope at the smallest particle sizes. This can be explained by equations 6.2 and
6.3. As dp 0 the Sherwood number approaches 2. Consequently, the mass flux scales
as d1p , and the mass transfer rate as dp . A mass conservation equation describing the
inter-phase mass transport results in the time derivative of the particle diameter to scale
with d1p . As a result, the particle size distribution collapses faster at low particle sizes
than at higher particle sizes.
Figure 6.8 shows the time trace of the normalized Sauter mean diameter (d32 =
P Np 3 P Np 2
( i=1 dp,i )/( i=1 dp,i )). The Sauter mean diameter decreases as a function of time till
6.4. Results 157

0
10
Nt = 2 Nt = 5 Nt = 7
Np /Np0 (-)

10-2
10-4
10-6
100
Nt = 10 Nt = 20 Nt = 40
Np /Np0 (-)

10-2
10-4

10-6
100
Nt = 60 Nt = 80 Nt = 100
Np /Np0 (-)

10-2
10-4
10-6
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
dp /dp 0 (-) dp /dp 0 (-) dp /dp 0 (-)

Figure 6.7: Instantaneous realizations of the particle size distribution (lin-log plots).
1

0.8
d32/dp0 (-)

0.6

0.4

0.2

0
0 20 40 60 80 100
Nt (-)

Figure 6.8: Evolution of the Sauter mean diameter in time. The crosses correspond with
the particle size distributions shown in Figure 6.6.
158 Chapter 6. Numerical simulation of a dissolution process

0.8
Np /Np0 (-)

0.6

0.4

0.2 Nqs,99%

0
50 60 70 80 90 100
Nt (-)

Figure 6.9: Evolution of number of particles throughout the tank in time, starting at N t =
50.

N t = 70. In the final stage of the dissolution process, the Sauter mean diameter increases
slightly. This effect is explained by the fact that the particle size distribution collapses
faster at low particle sizes than at high particle sizes.

6.4.4 Dissolution time


Compared to the various definitions in the literature of the mixing time in e.g. blending
processes (see Chapter 5), there is not much debate on the definition of a dissolution time.
The dissolution time is the time needed to fully dissolve all the solids. In a numerical
simulation including particle tracking, the total number of solid particles is monitored,
and hence the dissolution time is the time when the total number of particles has reached
the zero value.
In Figure 6.9a the evolution of the total number of particles normalized with the number
of released particles as a function of time is shown. The dissolution process starts at
N t = 42.5 (not shown), and from N t = 50 a steep decrease in the number of particles is
observed till N t = 70. In the final stage of the dissolution process a long tail is observed
in the time series of the particle number. The simulation was stopped at N t = 100 when
less than 0.1% of the total amount of released particles was left. With a view to a sensible
use of computational effort, it is not believed that any valuable information apart from the
dissolution time is to be found for N t > 100. The time instant where 99% of the particles
6.4. Results 159

2 2

1.6 1.6

1.2 1.2
c/c (-)

c/c (-)
0.8 0.8

0.4 0.4

0 0
0 20 40 60 80 100 0 20 40 60 80 100
Nt (-) Nt (-)

(a) z/T = 0.67; r/T = 0.25 (b) z/T = 0.19; r/T = 0.25

Figure 6.10: Simulated time traces of the concentration normalized with the final concen-
tration c .

is dissolved is s,99% = 84N 1 .

6.4.5 Time series of concentration

In Figure 6.10 time series of the scalar concentration are shown, recorded at two moni-
toring points. One monitoring point is located in the upper recirculation loop (i.e. in a
mid-way baffle plane at z/T = 0.67 and r/T = 0.25), and the other in the lower recircu-
lation loop (i.e. in a mid-way baffle plane at z/T = 0.19 and r/T = 0.25). In the first
12 15 impeller revolutions strong concentration fluctuations are observed, which corre-
spond with the macroscopic structures seen in Figure 6.3. The concentration fluctuations
registered at the monitoring point at z/T = 0.67 (Figure 6.10a) are stronger compared to
those at the monitoring point at z/T = 0.19 (Figure 6.10b). This is because the particles
were released in the top part of the tank. The strong concentration fluctuations only occur
in the mixing and dispersing stage (I).
Subsequently for N t > 15, the concentration recorded at both monitoring points grad-
ually increases to a final level of about 1.12c . The reason for the significant deviation of
12% is subject of the next subsection.
160 Chapter 6. Numerical simulation of a dissolution process

1.2
M/Mp0 (-) 1

0.8

0.6

0.4
Mp 0 - Mp
0.2
Mc
0
0 20 40 80 100
Nt (-)

Figure 6.11: The dissolved particle mass (M = Mp0 Mp ; solid line), and the tracer
mass (M = Mc ; dashed line), made dimensionless with the total particle
mass Mp0 as a function of time.

6.4.6 Mass conservation finite volume scheme


We have used the newly developed immersed boundary technique for the scalar concen-
tration to impose a zero-gradient constraint at the walls that are off-grid (see Sections 4.6
and 5.3.2). This technique is unconditionally stable in the explicit formulation of the finite
volume scheme used. However, mass conservation is not guaranteed.
The total mass in the system has been monitored in order to check the mass conserva-
tion constraint. The total mass of the solids (Mp ) is evaluated through:

Np
X 3
Mp = p d (6.10)
6 i=1 p,i

The total mass dissolved equals (Mp0 Mp ). In theory, the total mass dissolved equals
the total scalar mass (Mc ), which reads

NV
X
Mc = ci Vi (6.11)
i=1

where V is a control volume and NV is the number of control volumes.


In Figure 6.11 the time trace of the total mass dissolved (i.e. solid line) and the total
scalar mass divided by the fraction of dissolved particle mass (dashed line) normalized
6.5. Conclusions 161

with the total mass of the solids are shown. The total mass dissolved increases rapidly
in the first part of the dissolution process. At the end of stage I (Mixing and dispersing;
N t = 12) approximately 50% of the total solids mass is already dissolved, and at the start
of stage IV (Dissolution; N t = 42) 90% of the solids mass is dissolved.
At the end of the simulation, the total scalar mass equals 1.12 times the total dissolved
mass. The significant deviation of 12% has already been observed in the time series of the
scalar concentration in the previous subsection. The average mass increase per impeller
revolution equals 0.12% which corresponds to the average mass increase observed in the
simulations of the blending process (Chapter 5).
In Chapter 5 the increase of mass has been attributed to the lack of resolution to accu-
rately represent the edges of the impeller blades, impeller disk and baffles. Consequently,
the mass conservation constraint is not satisfied, and in future research the current tech-
nique needs to be improved. Possible solutions for better conserving scalar mass are (local)
grid refinement, and the use of a cylindrical grid by the scalar transport solver.

6.5 Conclusions
A simulation of a dissolution process (i.e. a solid-liquid suspension) according to an
Eulerian-Lagrangian approach has been presented in this chapter. In this approach, 7 mil-
lion particles were tracked in an unsaturated liquid flow field that was represented by a
large eddy simulation (LES). The impeller speed exceeds the just-suspended speed (Zwi-
etering, 1958) to keep the particles suspended in the liquid. Under the action of a driving
force (i.e the difference between the saturation concentration and the concentration of the
surroundings) and a mass transfer coefficient (depending on the particle slip velocity) the
particles slowly dissolved during the course of the simulation. The effects of particle ro-
tation and subgrid-scale concentration fluctuations on the mass transfer rate have been
ignored.
Typical features with respect to the solids distribution and the associated concentration
distribution have been observed. In the first stage of the process (typically the first 12 im-
peller revolutions), the particles are distributed throughout the tank. The particles organize
in streaky patterns, by leaving high-vorticity regions. High solids concentrations are found
below the impeller, at the bottom of the tank, in front of the impeller blades, at the outer
walls, and at the wall separation points. High solids concentrations are accompanied by
high scalar concentration (i.e. macroscopic structures). Large regions void of particles are
observed behind the impeller blades. The decrease of the particle mass in the course of the
process has two important implications. In the first place, the high particle concentration
162 Chapter 6. Numerical simulation of a dissolution process

at the bottom of the tank disappears as the particles get resuspended. Secondly, the streaky
patterns disappear due to decreasing inertia of the particles.
Through monitoring the number of particles in axial slices of width 0.1T , various
stages of the dissolution process were identified. In the first stage (0 12 impeller rev-
olutions), the particles are dispersed throughout the tank. Subsequently, the system is in
a quasi steady-state situation (i.e. the second stage, 12 24 impeller revolutions). A
significant decrease of the number of particles in the bottom axial slice indicates that the
particles at the bottom of the tank get resuspended (i.e. the third stage, 24 42 impeller
revolutions). The fourth and final stage covers the dissolution of the particles. At roughly
58 impeller revolutions a homogeneous solids suspension is reached.
The non-homogeneous mixing and consequently mass transfer in a stirred tank reactor
is expressed in the evolution of the particle size distribution. During the course of the
simulation, the particle size distribution broadens and moves to lower particle diameters.
The peculiar steep slope observed at the smallest particle sizes is due to the mass transfer
rate increasing at decreasing particle diameter for very small particles. This latter effect
causes the slight increase of the Sauter diameter in the final stage of the process: the
particle size distribution collapses at a higher rate for small particle sizes compared to
larger particle sizes.
For computational and physical reasons the simulation was stopped at N t = 100,
covering 99.9% of the dissolution process. The 99% dissolution time has been determined
by the time series of the particle number (i.e. the number of particles with dp > 0) and
equals 84 impeller revolutions. Relatively more particles are fully dissolved in the impeller
region and below.
Time traces of the concentration show in the first 15 impeller revolutions strong con-
centration fluctuations associated with the macro-scopic structures observed in the con-
centration distribution. Subsequently, the concentration measured at the monitoring points
increases gradually to 1.12c . The significant 12% deviation observed, corresponds to
the deviation between the total scalar mass and the total dissolved mass. The reason for
this unphysical disagreement is the newly developed immersed boundary technique for
imposing the zero-gradient boundary constraint at the off-grid walls. The lack of resolu-
tion in recovering the edges of the impeller blades, the impeller disk and the baffle causes
an inaccurate representation inducing leakage of mass into the system.
The artificially higher concentration (due to mass leakage) provides a reduced mass
transfer rate, and consequently an overprediction of the dissolution time. In the present
case studied, the final concentration is a factor 30 lower than the saturation concentration.
As a result, the overprediction of the dissolution time is expected to be marginal (only
0.5% based on a back-of-the-envelope calculation).
6.5. Conclusions 163

There are still quite some issues open for improvement with respect to the modeling
attempt here. The volume fraction occupied by the particles, which is significant in the
first stages of the dispersion, has not been taken into account. For further issues related
to the particle transport solver we refer to the paper of Derksen (2003). The mass leakage
into the system induced by our (novel) immersed boundary technique may be repaired by
(local) grid refinement or the usage of a cylindrical grid in the case of the scalar transport
solver.
The simulation of a dissolution process through LES including passive scalar transport
and particle tracking gave a detailed insight in the complex phenomena (i.e. the intricate
interplay of turbulence, mass transfer, motion of particles, collisions, etc) occurring during
the process. For the process considered in this work, it has been shown that the dissolution
time is at most one order of magnitude longer than the dispersion time scale. About 50% of
the total particle mass dissolves in the first stage (i.e. the mixing and dispersing stage; 0
N t 12) of the process. The current successful application of the coupled LES/particle
transport/scalar mixing solvers opens worthwhile future directions of numerical research
like crystallization processes.
164 Chapter 6. Numerical simulation of a dissolution process

Nomenclature

Roman Description Unit


A surface of the spherical particle m2
c concentration kg.m3
cs Smagorinsky constant
csat saturation concentration kg.m3
c final concentration kg.m3
dp particle diameter m
dp0 initial particle diameter m
d32 Sauter mean diameter m
D impeller diameter m
g magnitude of the gravitational acceleration vector m.s2
H height of the tank m
k mass transfer coefficient m.s1
mint inter-phase mass transfer rate kg.s1
M total mass kg
Mc total scalar mass kg
Mp total particle mass kg
Mp0 initial total particle mass kg
N impeller speed s1
Np number of particles
Np0 initial number of particles
NV total amount of control volumes
r radial coordinate m
Re Reynolds number
Rep particle Reynolds number
s constant in equation 6.1
Sc Schmidt number
Sct turbulent Schmidt number
Sc source term in equation 6.6 kg 2 .m6 .s1
Stk Stokes number
Sh Sherwood number
t time s
T tank diameter m
ui velocity component i m.s1
6.5. Conclusions 165

vsl slip velocity m.s1


V tank volume m3
xi coordinate i m
z axial coordinate m

Greek Description Unit


diffusion coefficient m2 .s1
e eddy diffusion coefficient m2 .s1
mol molecular diffusion coefficient m2 .s1
lattice spacing
t time step s
V control volume m3
difference between the densities of the particle and kg.m3
continuous phase
s dissolution time s
kinematic viscosity m2 .s1
e Smagorinsky eddy viscosity m2 .s1
l density of the continuous phase kg.m3
p density of the particle phase kg.m3
m0 initial mass fraction
7 Conclusions and perspectives

7.1 General discussion


The goal of the present work was to contribute to reliable and accurate numerical predic-
tions of complex, multi-phase processes occurring in stirred tank configurations. In order
to achieve this ambitious goal, three issues were treated successively:
Turbulent flow phenomena
Scalar mixing
Solid-liquid mixing including mass transfer
In order to improve our understanding of the complex, multi-phase processes, the first
ingredient needed is detailed information of the turbulent flow phenomena encountered in
a stirred tank. For that purpose, the large eddy simulation (LES) was chosen. A thorough
assessment of the LES flow solver based on the lattice-Boltzmann discretization scheme
has been presented in chapters 2 and 3. The main findings have been summarized in
section 7.2.
The description of inter-phase mass transfer requires information of the local species
concentrations. In order to describe passive scalar transport, a scalar mixing solver (based
on the finite volume approach) has been developed. The solver has been described and
assessed in chapter 4. An LES including passive scalar mixing is well suited to determine
the mixing performance of the turbulent flow in a stirred tank, which was the focus of
chapter 5. The main conclusions regarding scalar mixing are summarized in section 7.3.
The motion of suspended particles is handled with a particle transport solver, that was
already available. The coupled flow, scalar mixing and particle transport solvers allow for
detailed simulations of processes dealing with solid-liquid mixing including mass transfer
in accordance with the ultimate goal described above. We decided to study a dissolu-
tion process, which is the subject of chapter 6. The main conclusions based on the final
simulation are summarized in section 7.4.
168 Chapter 7. Conclusions and perspectives

7.2 Turbulent flow phenomena


The lattice-Boltzmann Navier-Stokes solver based on the LES methodology has been crit-
ically assessed in order to gain confidence in the method. Two subgrid-scale models have
been attempted. The focus is on the accuracy of the LES flow field predictions (in terms
of velocity, turbulent kinetic energy, energy dissipation, and turbulence anisotropy), and
the recovery of coherent fluctuations induced by large-scale precessing vortices (i.e. a
so-called macro-instability). For the LES assessment, results of LDA experiments and a
transient RANS simulation were used. The most important conclusions are as follows:

Both the LES and RANS models predict the average velocity field to a high level of
accuracy. Furthermore, the LES and RANS results show, in the absence of ex-
perimental results, an inhomogeneous distribution of the energy dissipation rate
throughout the tank. While a LES resolves the turbulence levels to a high level
of accuracy, in a RANS simulation they are underestimated by some 50%. The high
levels of turbulence anisotropy (e.g. in the impeller stream, and in the boundary
layers along the walls) observed in a LES suggest the need of a more sophisticated
closure model than a k- type of model in the RANS approach.

Next to the widely used Smagorinsky subgrid-scale model, a more sophisticated


subgrid-scale model (the so-called Voke model) has been tested. The latter model is
a blend between a DNS in fully resolved regions and a LES with the Smagorinsky
subgrid-scale model in regions where the flow is not fully resolved. Despite the
potential of the Voke model, the results did not show any significant improvements.

Large eddy simulations have successfully recovered coherent velocity fluctuations


caused by vortices precessing about the tank centerline. The frequencies of these
fluctuations at various values of the Reynolds number agree within 30% with fre-
quencies found experimentally in a similar stirred tank geometry. Furthermore, the
LES results have shown that the coherent fluctuations induced by the precessing
vortices contribute significantly to the turbulence levels (up to 44%) in the top and
bottom parts of the tank near the tank centerline, which is in accordance with exper-
imental findings.

7.3 Scalar mixing


The scalar transport is modeled by means of the finite volume approach. Within this
approach, an immersed boundary technique was developed to impose a zero-gradient con-
7.3. Scalar mixing 169

straint at off-grid boundaries. The scalar mixing solver has been assessed in two laminar
and one turbulent flow case. The focus is on the performance of the immersed boundary
technique, and the mixing time (in case of the blending process). The main conclusions of
the assessment are summarized below.

A description of scalar transport by means of the finite volume approach is less


memory intensive than a lattice-Boltzmann approach. The time step used for the
LES including passive scalar transport was found to be small enough to attain a
stable algorithm in the explicit formulation of the finite volume scheme.

Direct numerical simulations of the scalar mixing by a laminar flow in a square


cavity have lead to the conclusion that the flux-limited TVD scheme performs best
compared to the first-order upwind and second-order QUICK schemes. The TVD
scheme significantly reduces numerical diffusion and is unconditionally stable.

In case of the scalar mixing in an inclined cavity, the newly developed immersed
boundary technique to impose a zero-gradient constraint at off-grid walls performs
similar to the existing staircase technique in terms of mass conservation. Both tech-
niques do not conserve mass. In contrast to the staircase technique, the immersed
boundary technique takes into account the shape of the boundary regardless of the
grid resolution.

According to the results of the scalar mixing in an inclined cavity, the grid resolution
has a significant effect on the scalar mass change. This is confirmed by the blending
cases in the storage and the Rushton turbine stirred tanks. In the former case, the
average scalar mass increase amounts to 0.0016% per impeller revolution. Since the
cylindrical wall is accurately represented by the grid, the mass leakage is not sig-
nificant. In the latter case, the average scalar mass increase per impeller revolution
equals 0.05%, 0.125% and 0.25% of the amount of scalar mass injected at impeller
sizes equal to one-fourth, one-third and one-half of the tank diameter, respectively.
The significant scalar mass increase is ascribed to the lack of resolution to accurately
recover the thickness of the impeller blades, impeller disk and baffles.

In the literature there is no standardization of mixing time experiments. Mixing


times are reported for tanks of different layouts and operating conditions, tracer in-
jection positions and injection speeds, and number and positions of the measurement
points. Furthermore, there is no agreement on the definition of the mixing time. As
a result, simulated and/or measured mixing times should be compared with great
care. This also applies to mixing time correlations in the literature.
170 Chapter 7. Conclusions and perspectives

An LES of the scalar mixing in a cylindrical tank equipped with a side-entry mixer
resulted in a mixing time that agrees within 35% with the mixing time predicted by a
LES using Fluent. The mixing time in a Rushton turbine stirred tank, obtained by a
LES including passive scalar transport, agrees within 30% with experimental results
from the literature. Furthermore, the mixing time obtained at different impeller
sizes and injection positions also are within 30% of values calculated from two
correlations presented in the literature. The mixing time is not significantly affected
by the tracer injection position. On the other hand, impeller size has a significant
effect on the mixing time. According to the simulations, the mixing time scales with
the tank over impeller diameter ratio to the power 2.5.

7.4 Solid-liquid mixing including mass transfer


A dissolution process has been simulated in detail by means of a LES including passive
scalar transport, coupled to a Lagrangian description of spherical, solid particles immersed
in the turbulent flow driven by a Rushton turbine. The focus is on the transient phenomena
observed, in terms of solids and scalar concentrations, particle size distributions, and the
dissolution time. The main conclusions are formulated below.

The dissolution process has been successfully simulated until the point that 99.9%
of all particles have been fully dissolved. For the particular dissolution process
simulated, the dissolution time was found to be at most one order of magnitude
larger than the time needed to fully disperse the solids throughout the system.

With settings used in the simulation, four stages have been identified. In the first
stage the particles are dispersed throughout the system. In the second stage, the
distribution of solid particles throughout the tank is more or less steady. In the third
stage, the high solids concentration at the bottom of the tank decreases significantly
due to resuspension of particles. During the fourth stage, the particles are fully
dissolved, and the solids distribution is heading toward a homogeneous distribution.

Initially, the particles organize in streaky patterns due to the solids density being
higher than the density of the liquid, and the inertia of the particles. The particles
are swept out the high vorticity regions, which is in particular nicely illustrated by
the presence of the slowly precessing vortex below the impeller. During the course
of the process, the inertia of the particles decreases in time, and consequently the
particles will behave more like fluid flow tracers. As a result, the streaky patterns
disappear.
7.5. Perspectives and recommendations 171

Until the point the particles are fully distributed throughout the tank, the spatial
solids and scalar distributions are very inhomogeneous. High scalar concentrations
are associated with high solids concentrations. Subsequently, the spatial solids and
scalar concentration distributions become more or less homogeneous. The concen-
tration increases gradually throughout the tank.

The influence of non-homogeneous mixing effects on the mass transfer is expressed


in the development of a particle size distribution. During the course of the simula-
tion, the particle size distribution broadens and shifts toward smaller particle diam-
eters.

The unphysical change of scalar mass due to the newly developed immersed bound-
ary technique, is significant (i.e. an average increase per impeller revolution of
0.12% of the initial solids mass), and resembles that determined in the blending
process in the Rushton turbine stirred tank. The higher levels of scalar concen-
tration due to leakage of scalar mass into the tank result in a lower driving force
of inter-phase mass transfer. This consequently leads to an overestimation of the
dissolution time. However, the impact of the mass leakage on the mixing and disso-
lution times is expected to be marginal. In case of the dissolution process studied,
a back-of-the-envelope calculation resulted in an overestimation of the dissolution
time by about 0.5%.

7.5 Perspectives and recommendations


In view of the work presented in the successive chapters, the ultimate goal of contributing
to reliable and accurate predictions of complex, multi-phase processes has been met. The
basis of a simulation is the solution of the flow field. In order to have an accurate descrip-
tion of the various, complex mechanisms occurring in these processes, their interplay with
the turbulent velocity fluctuations present in the flow field should be represented to a high
level of detail.
An attractive, and widely used approach today (in terms of run times, memory use,
user-friendly interface) is CFD based on the Reynolds-averaged Navier-Stokes equations
(RANS). With a view to the complexity of a multi-phase process, the high level of detail
is not reached by a steady-state approximation. Although a transient RANS simulation is
able to predict the flow field reasonably, we have seen that it fails in providing detailed
information of turbulent length and time scales. A k  type of closure model, which
is widely used today, invariably underpredicts the turbulence levels by 50%, and this will
172 Chapter 7. Conclusions and perspectives

certainly affect the mixing performance and the overall outcome of the particular process
of interest.
In the last decade, large eddy simulations (LES) have demonstrated their ability to
provide accurate details of the flow field and the turbulence levels. Application of LES to
complex, multi-phase processes in industrial scale reactors is not yet within reach, because
of limited computational resources. However, studying such processes in lab-scale reactors
by means of LES in conjunction with e.g. scalar mixing and particle transport solvers (as
in this work) has become a promising possibility. The studies as reported in chapters 5 and
6 provide detailed insight in the mixing performance of the turbulent flow in an agitated
tank, and the physical mechanisms occurring in a dissolution process. Because of the
continuing growth of computer resources, scaling-up studies are a matter of time.
Although we have been able to achieve the goal posed, there is quite some room for im-
provements. When reading the thesis, several weak spots in the modeling attempts can be
discerned. In the first place, the applicability of the particle transport code is limited by the
size of the particles and the time step in relation to the collision algorithm. Secondly, the
hydrodynamic forces and the mass transfer rate are based on single-particle correlations;
the hydrodynamic interactions between particles are not included. Thirdly, the immersed
boundary technique introduced in the finite volume approach is not mass conservative, the
mass leakage is expected to be most pronounced at the edges of the impeller blades.
We propose here some suggestions for improvement. The performance of the collision
algorithm may be improved by reducing the time step. The single-particle correlations
might be adapted such that hydrodynamic interactions are included through the inclusion
of e.g. the local solids volume fraction. In this respect, a more advanced approach based
on direct numerical simulations of a particle suspension in a simple geometry may be used
to derive relations for hydrodynamic interaction forces. Increasing the spatial resolution
(notably near the walls by means of local grid refinement) results in a better performance
of the immersed boundary technique. Other boundary techniques may be explored, such
as the cut-cell technique (see e.g. Tucker & Pan (1999)) in combination with an implicit
formulation of the convection-diffusion equation.
In order to increase the confidence in a multi-phase simulation, detailed experiments
are needed. These experiments should provide averaged as well as angle-resolved data
of the particle concentration and scalar concentration fields, local flow velocities in the
presence of particles, and information on collisions and mass transfer rates.
The numerical results obtained in this thesis demonstrate the capability of the current
solvers to simulate a complex, multi-phase process in detail. As a result, other worth-
while future directions of numerical research, like industrial crystallization processes, are
opened. Various phenomena common in crystallization can already be recovered with
7.5. Perspectives and recommendations 173

the current solvers available. For instance, the mass transfer needed for crystal growth,
brought about by the fluid flow, is determined by the species concentrations (e.g. super-
saturation) which is provided by the scalar transport solver. Also, information on crystal-
crystal collisions that are induced by local velocity gradients is directly available. The
knowledge of the local supersaturation in combination with the collision rate of the crys-
tals enable the description of crystal agglomeration. Furthermore, attrition events (a major
source of secondary nucleation) may be predicted with a relative ease, since the frequency
and intensity of crystal-crystal and crystal-impeller collisions are provided by the particle
transport solver.
We have seen that realistic modeling of solid-liquid suspensions including mass trans-
fer is not an easy task. While it is demonstrated that an accurate representation of the flow
field by means of a LES is the first step toward success, we still need to revert to mod-
eling particle/crystal related issues. For instance, the occurrence of a nucleation event is
highly determined by concentration fluctuations at the subgrid-scale level. Consequently,
these fluctuations should be taken into account by means of a subgrid-scale model. Fur-
thermore, the performance of the current collision algorithm is limited by particle size,
and consequently the simulation of a crystallization process is restricted to low volume
fractions. In addition, the description of chemical reactions is an open problem as well.
In case of reaction schemes with order higher than one, a closure model is needed for the
reaction rate term of the scalar components that arises in the filtered convection-diffusion
equation. Two major approaches may be adopted in handling the closure of the chem-
ical reactions; the mechanistic micro-mixing approach (which has been exploited in the
Kramers laboratory by Bakker (1996)) or the more advanced stochastic approach in terms
of probability density functions (PDFs) through Monte Carlo simulations (exploited by
Van Vliet (2003)). While, in the latter approach the causality problems encountered in the
former approach are avoided, the computational demand is higher.
The work presented in this thesis has set a firm basis for accurate numerical predictions
of complex, multi-phase processes. Hopefully, this thesis will contribute to new develop-
ments and encourages other researchers to tackle the challenges and problems (such as
mentioned above) that are encountered in the future directions of numerical research on
transport phenomena and chemically reacting flows.
Bibliography

A LVAREZ , J., A LVAREZ , J. & H ERNANDEZ , M. 1994 A population balance approach


for the description of particle size distribution in suspension polymerization reactors.
Chemical Engineering Science 49 (1), 99113.

BAKKER , A. & O SHINOWO , L. M. 2004 Modelling of turbulence in stirred vessels using


large eddy simulation. Chemical Engineering Research and Design 82(A9), 11691178.

BAKKER , R. 1996 Micromixing in chemical reactors. Models, experiments and simula-


tions. PhD thesis, Delft University of Technology.

BARTELS , C., B REUER , M., W ECHSLER , K. & D URST, F. 2001 Computational fluid dy-
namics applications on parallel-vector computers: computations of stirred vessel flows.
Computers and Fluids 31 (1), 6997.

B ORIS , J. P. & B OOK , D. L. 1973 Flux-corrected transport I. SHASTA, a transport algo-


rithm that works. Journal of Computational Physics 11, 3869.

B OUWMANS , I., BAKKER , A. & VAN DEN A KKER , H. E. A. 1997 Blending liquids of
different viscosities and densities in stirred vessels. Chemical Engineering Research &
Design 75(A8), 777783.

B UJALSKI , W., JAWORSKI , Z. & N IENOW, A. W. 2002 CFD study of homogenization


with dual Rushton turbines - comparsion with experimental results. Chemical Engineer-
ing Research and Design 80, 97104.

C ALHOUN , D. & L E V EQUE , R. J. 2000 A Cartesian grid finite volume methode for the
advection-diffusion equation in irregular geometries. Journal of Computational Physics
157, 143180.

CFX-5 2002 User manual, solver and solver manager, pp. 355366.
176 Bibliography

C HEN , M., KONTOMARIS , K. & M C L AUGHLIN , J. B. 1998 Direct numerical simulation


of droplet collisions in a turbulent channel flow. Part I: collision algorithm. International
Journal of Multiphase Flow 24, 1079.

C HEN , S. & D OOLEN , G. D. 1998 Lattice-Boltzmann method for fluid flows. Annual
Review of Fluid Mechanics 30, 329364.

D ERKSEN , J. J. 2001 Assessment of large eddy simulations for agitated flows. Transac-
tions of the Institute of Chemical Engineers 79A, 824830.

D ERKSEN , J. J. 2003 Numerical simulation of solids suspension in a stirred tank. AIChE


Journal 49 (11), 27002714.

D ERKSEN , J. J. & VAN DEN A KKER , H. E. A. 1999 Large eddy simulations on the flow
driven by a Rushton turbine. AIChE Journal 45 (2), 209221.

D ERKSEN , J. J. & VAN DEN A KKER , H. E. A. 2000 Simulation of vortex core precession
in a reverse-flow cyclone. AIChE Journal 46 (7), 13171336.

D ERKSEN , J. J., D OELMAN , S. & VAN DEN A KKER , H. E. A. 1999 Three-dimensional


LDA measurements in the impeller region of a turbulently stirred tank. Experiments in
Fluids 27, 522532.

D ISTELHOFF , M. F. W., M ARQUIS , A. J. & N OURI , J. M. AND . W HITELAW, J. H. 1997


Scalar mixing measurements in batch operated stirred tanks. The Canadian Journal of
Chemical Engineering 75, 641652.

E GGELS , J. G. M. 1996 Direct and large eddy simulations of turbulent fluid flow using the
lattice-Boltzmann scheme. International Journal of Heat and Fluid Flow 17, 307323.

E GGELS , J. G. M. & S OMERS , J. A. 1995 Numerical simulation of free convective


flow using the lattice-Boltzmann scheme. International Journal of Heat and Fluid Flow
16 (5), 357364.

E LGOBASHI , S. 1994 On predicting particle-laden turbulent flows. Applied Scientific Re-


search 52, 309.

G ALETTI , C., L EE , K. C., PAGLIANTI , A. & Y IANNESKIS , M. 2004 Reynolds num-


ber and impeller diameter effects on instabilities in stirred vessels. Submitted to AIChE
Journal .
Bibliography 177

G ODUNOV, S. K. 1959 A difference scheme for numerical computation of discontinuous


solution of hydrodynamic equations. Math. Sbornik 47, 271306.

G OODMAN , J. B. & L E V EQUE , R. J. 1985 On the accuracy of stable schemes for 2D


scalar conservation laws. Mathematics of computation 45, 1521.

H AAM , S., B RODKEY, R. S. & FASANO , J. B. 1992 Local heat transfer in a mixing vessel
using heat flux sensors. Industrial Engineering and Chemical Research 31, 13841391.

H ARTEN , A. 1983 High resolution schemes for hyperbolic conservation laws. Journal of
computational physics 49, 357393.

H ARTMANN , H. 2001 Report on heat transfer coefficients and temperature gradients in


the DOW lab and pilot reactor (confidential report).

H ARTMANN , H. 2002a Form factor estimation (confidential report).

H ARTMANN , H. 2002b Population balance modelling; a report produced for the OPTI-
MUM project (confidential report).

H ARTMANN , H., D ERKSEN , J. J. & VAN DEN A KKER , H. E. A. 2004a Macro-instability


uncovered in a Rushton turbine stirred tank. AIChE Journal 50 (10), 23832393.

H ARTMANN , H., D ERKSEN , J. J., M ONTAVON , C., P EARSON , J., H AMILL , I. S. &
VAN DEN A KKER , H. E. A. 2004b Assessment of large eddy and RANS stirred tank
simulations by means of LDA. Chemical Engineering Science 59 (12), 24192432.

H ASAL , P., M ONTES , J. L., B OISSON , H. C. & F ORT, I. 2000 Macro-instabilities of


velocity field in stirred vessel: detection and analysis. Chemical Engineering Science
55, 391401.

H IRSCH , C. 1990 Numerical computation of internal and external flows, volume 2. John
Wiley & Sons, Chichester.

H OEFSLOOT, H. C. J., W ILLEMSEN , S. M., H AMERSMA , P. J. & I EDEMA , P. D. 2000


Mixing of two liquids with different rheological behaviour in a lid driven cavity. In Pro-
ceedings of the 10th European conference on mixing. July 2nd -5th , Delft, The Nether-
lands.

H OLLANDER , E. D., D ERKSEN , J. J., P ORTELA , L. M. & VAN DEN A KKER , H. E. A.


2001 Numerical scale-up study for orthokinetic agglomeration in stirred vessels. AIChE
Journal 47, 24252440.
178 Bibliography

H OLMES , D. B., VONCKEN , R. M. & D EKKER , J. A. 1964 Fluid flow in turbine stirred,
baffled tanks I: circulation time. Chemical Engineering Science 19, 201208.

H OOGENDOORN , C. J. & D EN H ARTOG , A. P. 1967 Model studies on mixers in viscous


flow regions. Chemical Engineering Science 22 (12), 16891699.

H OUCINE , I., P LASARI , E., DAVID , R. & V ILLERMAUX , J. 1999 Feedstream jet inter-
mittency phenomena in continuous stirred tank reactor. Chemical Engineering Journal
72, 1930.

JAWORSKI , Z., B UJALSKI , W., OTOMO , N. & N IENOW, A. W. 2000 CFD stduy of
homogenization with dual Rushton turbines - comparison with experimental results.
part I: initial studies. Chemical Engineering Research and Design 78, 327333.

K ANDHAI , D., D ERKSEN , J. J. & VAN DEN A KKER , H. E. A. 2003 Interphase drag
coefficients in gas-solid flows. AIChE Journal 49 (4), 10601065.

K IPARISSIDES , C. 1996 Polymerization modelling: a review of recent developments and


future directions. Chemical Engineering Science 51, 1637.

K RAMERS , H., BAARS , G. M. & K NOLL , W. H. 1953 A comparative study on the rate
of mixing in stirred tanks. Chemical Engineering Science 2, 3542.

L ATHOUWERS , D. 1999 Modelling and simulation of turbulent bubbly flow. PhD thesis,
Delft University of Technology.

L EE , K. C. 1995 An experimental investigation of the trailing vortex structure and mixing


characteristics of mixing vessels. PhD thesis, Kings College London, University of
London.

L EONARD , B. P. 1979 A stable and accurate convective modelling procedure based on


quadratic upstream interpolation. Computational methods in applied mechanics and en-
gineering 19, 5998.

L ESIEUR , M. & M ETAIS , O. 1996 New trends in large eddy simulations of turbulence.
Annual Review of Fluid Mechanics 28, 4582.

L UMLEY, J. 1978 Computational modelling of turbulent flows. Advances in Applied Me-


chanics 26, 123176.

L UO , H. & S VENDSEN , H. F. 1996 Theoretical model for drop and bubble break-up in
turbulent dispersions. AIChE Journal 42, 12251233.
Bibliography 179

M AGGIORIS , D., G OULAS , A., A LEXOPOULOS , A. H., C HATZI , E. G. & K IPARIS -


SIDES , C. 2000 Prediction of particle size distribution in suspension polymerization re-
actors: effect of turbulence non-homogeneity. Chemical Engineering Science 55, 4611
4627.

M ASON , P. J. & C ALLEN , N. S. 1986 On the magnitude of the subgrid-scale eddy coef-
ficient in large eddy simulations of turbulent channel flow. Journal of Fluid Mechanics
162, 439462.

M ENTER , F. R. 1994 Two-equation eddy-viscosity turbulence models for engineering


applications. AIAA Journal 32 (8), 269289.

M ETAIS , O. & L ESIEUR , M. 1992 Spectral large eddy simulations of isotropic and stably
stratified turbulence. Journal of Fluid Mechanics 239, 157.

M ICHELETTI , M., BALDI , S., Y EOH , S. L., D UCCI , A., PAPADAKIS , G., L EE , K. C.
& Y IANNESKIS , M. 2004 On spatial and temporal variations and estimates of energy
dissipation in stirred reactors. Transactions of the Institute of Chemical Engineers 82A,
11881198.

M OO -YOUNG , M., T ICHAR , K. & TAKAHASHI , F. A. L. 1972 The blending efficiencies


of some impellersin batch mixing. AIChE Journal 18, 178182.

M UMTAZ , H. S., H OUNSLOW, M. J., S EATON , M. J. & PATERSON , W. R. 1997 Or-


thokinetic aggregation during precipitation, a computational model for calcium oxalate
monohydrate. Transactions of the Institute of Chemical Engineers 75, 152159.

M YERS , K. J., WARD , R. W. & BAKKER , A. 1997 A digital particle image velocimetry
investigation of flow field instabilities of axial-flow impellers. ASME Journal of fluids
engineering 119, 623632.

N G , K., F ENTIMAN , N. J., L EE , K. C. & Y IANNESKIS , M. 1998 Assessment of sliding


mesh CFD predictions and LDA measurements of the flow in a tank stirred by a Rushton
impeller. Transactions of the Institute of Chemical Engineers 76A, 737747.

N IKIFORAKI , L., M ONTANTE , G., L EE , K. C. & Y IANNESKIS , M. 2002 On the origin,


frequency and magnitude of macro-instabilities of the flows in stirred vessels. Chemical
Engineering Science 58 (13), 29372949.

O PPENHEIM , A. V., W ILLSKY, A. S. & YOUNG , I. T. 1983 Signals and systems, pp.
388389. Prentice-Hall International Editions.
180 Bibliography

O SMAN , J. J. & VARLEY, J. 1999 The use of computational fluid dynamics (CFD) to
estimate mixing times in a stirred tank. In IChemE Symposium Series 146, pp. 1522.

PAO , Y. H. 1965 Structure of turbulent velocity and scalar fields at large wavenumbers.
Physics of Fluids 8, 10631075.

PATANKAR , S. V. 1980 Numerical heat transfer and fluid flow. Hemisphere Publ. Corp.

P IOMELLI , U., M OIN , P. & F ERZIGER , J. H. 1988 Model consistency in large eddy
simulation of turbulent channel flows. Physics of Fluids 31, 18841891.

P RINCE , M. J. & B LANCH , H. W. 1990 Bubble coalescence and break-up in air-sparged


bubble columns. AIChE Journal 36, 14851499.

P ROCHAZKA , J. & L ANDAU , J. 1961 Studies on mixing XII. homogenation of miscible


liquids in the turbulent region. Coll. Czech. Chem. Commun. 26, 29612973.

Q IAN , Y. H., D H UMIERES , D. & L ALLEMAND , P. 1992 Lattice BGK for the navier-
stokes equation. Europhysics Letters 17 (6), 479484.

R AMKRISHNA , D. 1985 The status of population balances. Review in Chemical Engineer-


ing 3, 4995.

R ANZ , W. E. & M ARSHALL , W. R. 1952 Evaporation from drops. Chemical Engineering


Progress 48, 142180.

ROE , P. L. 1981 Approximate Riemann-solvers, parameter vectors and difference


schemes. Journal of Computational Physics 43, 357372.

ROUSSINOVA , V., G RGIC , B. & K RESTA , S. M. 2000 Study of macro-instabilities in


stirred tanks using a velocity decomposition technique. Transactions of the Institute of
Chemical Engineers 78A, 10401052.

ROUSSINOVA , V., K RESTA , S. M. & W EETMAN , R. 2003 Low frequency macroinstabil-


ities in a stirred tank: scale-up and prediction based on large eddy simulations. Chemical
Engineering Science 58, 22972311.

RUSHTON , J. H., C OSTICH , E. W. & E VERETT, H. J. 1950 Power characteristics of


mixing impeller I and II. Chemical Engineering Progress 46, 395404 and 467476.

RUSZKOWSKI , S. 1994 A rational method for measuring blending performance, and com-
parison of different impeller types. In Proceedings of the 8th European conference on
mixing. September 21st -23rd , Cambride, U.K.
Bibliography 181

S ANO , Y. & U SUI , H. 1985 Interrelations among mixing time, power number and dis-
charge flow rate number in baffled mixing vessels. Journal of Chemical Engineering
Japan pp. 4752.

S CHFER , M. 2001 Charakterisierung, Auslegung und Verbesserung des Makro- und


Mikromischens in gerhrten Behltern. PhD thesis, Technischen Fakultt der Univer-
sitt Erlangen-Nrnberg.

S CHFER , M., H FKEN , M. & D URST, F. 1997 Detailed LDV measurements for visu-
alization of the flow field within a stirred-tank reactor equipped with a Rushton turbine.
Transactions of the Institute of Chemical Engineers 75A, 729736.

S CHFER , M., Y IANNESKIS , M., WCHTER , P. & D URST, F. 1998 Trailing vortices
around a 45o pitched-blade impeller. AIChE Journal 44 (6), 12331246.

S HARP, K. V. & A DRIAN , R. J. 2001 PIV study of small-scale flow structure around a
Rushton turbine. AIChE Journal 47 (4), 766778.

S HIUE , S. W. & W ONG , C. W. 1984 Studies on homogenization efficiency of various


agitators in liquid blending. Canadian Journal of Chemical Engineering 62, 602609.

S MAGORINSKY, J. 1963 General circulation experiments with the primitive equations: 1.


The basic experiment. Monthly Weather Review 91, 99164.

S OMERS , J. A. 1993 Direct simulations of fluid flow with cellular automata and the lattice-
Boltzmann equation. Applied Scientific Research 51, 127133.

S WEBY, P. K. 1983 High resolution schemes using flux-limiters for hyperbolic conserva-
tion laws. SIAM Journal of Numerical Analysis 21, 357393.

TATTERSON , G. 1994 Scale-up and design of industrial mixing processes. McGraw-Hill,


Inc., New York.

T EN C ATE , A., K RAMER , H. J. M., VAN ROSMALEN , G. M. & VAN DEN A KKER ,
H. E. A. 2001 The microscopic modelling of hydrodynamics in industrial crystallizers.
Chemical Engineering Science 56, 2495.

T SOURIS , C. & TAVLARIDES , L. L. 1994 Breakage and coalescence models for drops in
turbulent dispersions. AIChE Journal 40 (3), 395406.

T UCKER , P. G. & PAN , Z. 1999 A Cartesian cut cell method for incompressible viscous
flow. Applied Mathematical modelling 24, 591606.
182 Bibliography

VAN D RIEST, E. R. 1956 On turbulent flow near a wall. Journal of the Aeronautical
Sciences 23, 10071011.

VAN L EER , B. 1979 Towards the ultimate conservative difference scheme. V. A second-
order sequel to Godunovs method. Journal of Computational Physics 32, 101136.

VAN L OOY, R., K ENJERES , S., H ANJALIC , K., A L BATAINEH , K., H OLLANDER ,
E. & C OLENBRANDER , G. 2004 CFD analysis of the mixing behaviour of a sta-
ble stratification in a cylindrical tank with a side-entry mixer. In Proceedings of the
5th International Bi-annual ASME/JSME Symposium on Computational Technology
for Fluid/Thermal/Stressed systems with Industrial Applications. July 25th -29th , San
Diego/La Jolla, CA, USA.

VAN T R IET, K. & S MITH , J. M. 1975 The trailing vortex system produced by Rushton
turbine agitators. Chemical Engineering Science 30, 10931105.

VAN V LIET, E. 2003 Turbulent reactive mixing in process equipment. 3D LIF experiments
and FDF/LES modelling. PhD thesis, Delft University of Technology.

VAN V LIET, E., D ERKSEN , J. J. & VAN DEN A KKER , H. E. A. 2001 Modelling of par-
allel competitive reactions in isotropic homogeneous turbulence using a filtered density
approach for large eddy simulations. In Proceedings of PVP01, 3rd International sympo-
sium on computational technologies for fluid/thermal/chemical systems with industrial
applications. July 22nd -26th , Atlanta, Georgia, USA.

VAN WAGENINGEN , W. F. C., K ANDHAI , D., M UDDE , R. F. & VAN DEN A KKER ,
H. E. A. 2004 Dynamic flow in Kenics static mixer: an assessment of various CFD
methods. AIChE Journal 8, 16841696.

V IVALDO -L IMA , E., W OOD , P. E., H AMIELEC , A. E. & P ENLIDIS , A. 1998 Calcula-
tion of the particle size distribution in suspension polymerization using a compartment-
mixing model. Canadian Journal of Chemical Engineering 76, 495505.

VOKE , P. R. 1996 Subgrid-scale modelling at low mesh Reynolds number. Theoretical


and Computational Fluid Dynamics 8, 131143.

WANG , Y. & H UTTER , K. 2001 Comparisons of numerical methods with respect to con-
vectively dominated flows. International Journal for Numerical Methods in Fluids 37,
721745.
Bibliography 183

W ILCOX , D. C. 1993 In Turbulence modelling for CFD. La Canda (CA), DCW Indus-
tries.

Y EOH , S. L., PAPADAKIS , G., L EE , K. C. & Y IANNESKIS , M. 2004 Large eddy simula-
tion of turbulent flow in a rushton impeller stirred reactor with sliding-deforming mesh
methodology. Chemical Engineering and Technology 27 (3), 257263.

Y EOH , S. L., PAPADAKIS , G. & Y IANNESKIS , M. 2005 Determination of mixing time


and degree of homogenity in stirred vessels with large eddy simulation. Chemical Engi-
neering Science 60, 22932302.

Y IANNESKIS , M. 2004 Private communication.

Y IANNESKIS , M., P OPIOLEK , Z. & W HITELAW, J. H. 1987 An experimental study of the


steady and unsteady flow characteristics of stirred reactors. Journal of Fluid Mechanics
175, 537555.

Z ALESAK , S. T. 1979 Fully multidimensional flux corrected transport algorithms for flu-
ids. Journal of Computational Physics 31, 335362.

Z IJLEMA , M. 1996 Computational modelling of turbulent flow in general domains. PhD


thesis, Delft University of Technology.

Z WIETERING , T. N. 1958 Suspending of solid particles in liquid by agitators. Chemical


Engineering Science 8, 244.
Dankwoord

Het werk is voor mij nu af. Een periode van vier jaar lijkt als je begint erg lang, maar
niets is minder waar. De jaren zijn voorbij gevlogen en nu kan ik terugkijken op een tijd
die voor mij erg leerzaam is geweest. Naast de wetenschappelijke verdieping heb ik veel
geleerd van de samenwerking in het Europese project OPTIMUM en van de contacten
met lotgenoten in de wetenschappelijke wereld tijdens conferenties.
Het resultaat van mijn onderzoekswerk dat opgetekend staat in dit boekwerk is niet
louter door mij alleen verkregen. Alhoewel je geacht wordt zelfstandig onderzoek te kun-
nen verrichten, is het slagen hierin te danken aan de steun van vele mensen in mijn (directe)
omgeving. Ik heb nu de mogelijkheid deze mensen te bedanken en hun namen op de laatste
bladzijden van mijn proefschrift vast te leggen.
In de eerste plaats wil ik mijn promotor Harrie van den Akker bedanken. Ik weet
nog goed dat je in de tijd van mijn afstudeerperiode vertrouwen in mij had om een pro-
motieonderzoek succesvol af te kunnen ronden. Halverwege mijn promotietijd hebben we
samen nog wel eens nagedacht over de richting waarin mijn onderzoek zou gaan. Het
OPTIMUM project gaf mij niet veel vrijheid een eigen smaak te geven aan mijn onder-
zoek. Gelukkig heb ik dat het laatste jaar helemaal goed kunnen maken, mede dankzij
jouw steun en inspiratie in een voor jou moeilijke tijd. Harrie, bedankt voor alle goede
discussies en voor de leuke en leerzame momenten tijdens de conferenties in Indianapolis
en Lake Placid.
Mijn onderzoek was niet van de grond gekomen zonder de steun van mijn begeleider
Jos Derksen. Jos, de door jou ontwikkelde LES code vormde de basis van mijn werk. Ik
ben je dankbaar voor alle input die je mij geleverd hebt. We hebben vele discussies gevoerd
binnen, maar zeker ook buiten het lab. Voor het OPTIMUM project zijn we samen vaak
op pad geweest. Weet je nog dat we in een fantastisch hotel in Erlangen overnacht hebben
en de hachelijke verkeersituaties in Engeland die we overleefd hebben? Op deze plaats
wil ik jou en je gezin ook van harte bedanken voor de week dat ik bij jullie mocht logeren
in Princeton. De discussies die we daar gehad hebben, resulteerden uiteindelijk in een
significant gedeelte van mijn proefschrift.
186 Dankwoord

I will now switch to English for a moment to thank all partners of the OPTIMUM
project. In the next sentences I will thank some of them in particular. Bernd Genenger,
you were the chief in charge. I would like to thank you for the different discussions
we had, in particular about the macro-instabilities. Gerrit Hommersom, thanks for the
discussions on the suspension polymerization process, the Maxwell model, and the nice
conversations we had during our stay in a small hotel in England. Ian Hamill, Christiane
Montavon and Jane Pearson, your help was of high importance in case of the preparation
of chapter 2 and the paper published in Chemical Engineering Science. Christiane, we had
fruitful (e-mail) discussions considering the Maxwell model, thank you for that. Michael
Yianneskis, Loukia Nikiforaki and Martina Micheletti, thank you for the discussions and
the great moments we spent together during the OPTIMUM project. Your input was
valuable in the realization of chapter 3 and the paper published in AIChE Journal. Mike,
I appreciate your effort in reading my dissertation and for taking place in the commission
for my PhD defense.
Weer terug in het Nederlands ben ik aangekomen bij mijn collega-promovendi. On-
danks dat ik vaak weg was voor OPTIMUM vergaderingen en conferenties, heb ik toch
het grootste gedeelte van mijn tijd doorgebracht in het lab. Een aantal mensen wil ik in het
bijzonder bedanken: mijn W220 kamergenoten Andreas ten Cate, Wouter Harteveld, Mar-
tin Rohde, Ruurd Dorsman, Marco Zoeteweij en Jos van t Westende. Wouter en Martin,
met jullie heb ik enorm veel lol gehad! Ik denk met veel plezier terug aan al de momenten
(daar kan je zelfs een proefschrift van schrijven!) die ik met jullie beleefd heb.
Ik kan zeker niet voorbijgaan aan de dames van het secretariaat: Karin van de Graaf
en Thea Miedema. Bedankt voor alle gesprekken die we gehad hebben en de steun die ik
van jullie heb mogen ervaren tijdens de ziekte van Mirjam. Ik mis jullie al vanaf de eerste
dag dat mijn contract verlopen was, onvervangbaar! Ook de drie Japen, Thea, Jan, Wouter
en Ab bedankt voor de leuke gesprekken aan de koffietafel! Ik kan als promovendus
die volledig afhankelijk is van het rekencluster niet voorbijgaan aan de systeembeheerder
(alias computermuis) Peter Bloom: bedankt voor het fungeren als vraagbaak en je bijdrage
aan de significante uitbreiding van het rekencluster!
Nu kom ik aan bij de mensen die mij vanuit de privesfeer enorm tot steun zijn geweest.
Allereerst mijn vrienden Arjan en Mirjam Grevengoed en Han en Ella Knoeff. Jullie zijn
mij heel dierbaar geworden. Meer dan de motivatie tijdens mijn promotiewerk waardeer
ik jullie ondersteuning op geestelijk vlak. Jullie hebben mij elke keer weer laten zien om
Wie het draait in het leven. Heel veel dank ben ik jullie verschuldigd!
Pa en ma, ik wil jullie danken voor wie ik ben zoals ik nu ben. Jullie zorg en moti-
vatie tijdens mijn school- en universiteitsperiodes hebben hiertoe bijgedragen. Ook mijn
schoonouders (laat het woordje schoon maar weg!) wil ik bedanken. In de acht jaar dat
187

ik jullie dochter nu ken, zijn ook jullie voor mij ouders geworden. Bedankt dat ik mijn
frustraties en onzekerheden bij jullie kwijt kon tijdens mijn promotie. Meer dank nog voor
jullie zorg voor Mirjam en Lon tijdens Mirjams ziekteperiodes. Eigenlijk is dit niet in
woorden uit te drukken.
Tot slot mijn vrouw, Mirjam. Wat jij voor mij betekent is onbeschrijflijk. We hebben
hele moeilijke periodes gekend de laatste twee jaar. Tijdens de momenten dat jij zo ziek
was, dacht je nog aan mij en Lon. Jij hebt tijdens mijn afstuderen en promoveren heel
erg veel geduld met mij gehad. Bedankt voor jouw gebeden en dat je mij elke keer er
weer op gewezen hebt dat ik alles bij God neer mag leggen. Zonder jou had ik het niet
  
  
gered!
List of Publications

H. Hartmann, J.J. Derksen and H.E.A. van den Akker (2003). An LES investigation
of the flow macro-instability in a Rushton turbine stirred tank. Proceedings of the
11th European Conference on Mixing, Bamberg, Germany

H. Hartmann, J.J. Derksen, C. Montavon, J. Pearson, I.S. Hamill and H.E.A. van
den Akker (2004). Assessment of large eddy and RANS stirred tank simulations by
means of LDA. Chem. Eng. Sci., 59(12), 2419-2432

H. Hartmann, J.J. Derksen and H.E.A. van den Akker (2004). Macro-instability
uncovered in a Rushton turbine stirred tank. AIChE J., 50(10), 2383-2393

H. Hartmann, J.J. Derksen and H.E.A. van den Akker (2005). LES of the mixing
time in a Rushton turbine stirred tank. Submitted to AIChE J.

H. Hartmann, J.J. Derksen and H.E.A. van den Akker (2005). Numerical simulation
of a solubility process in a stirred tank reactor. Submitted to Chem. Eng. Sci.
About the Author

Hugo Hartmann was born on August 8th 1977 in Zuidelijke IJsselmeerpolders (Lelystad),
the Netherlands. In 1995, he graduated from pre-university education (VWO) at the In-
terconfessionele Scholengemeenschap Arcus in Lelystad. From 1995 to 2001, he studied
Applied Physics at Delft University of Technology. During this period he was a member
of a Christian students association C.S.R. (Civitas Studiosorum Reformatorum). He did
his graduation research at the Kramers Laboratorium voor Fysische Technologie (Delft
University of Technology), and on February 15th 2001 he graduated on his thesis entitled
Investigation of flashing induced instabilities in boiling water reactors. On March 1 st
2001, he started working on his PhD research at the same department. Till December
2003, he took part in a European project called OPTIMUM (Optimization of Industrial
Multiphase Mixing). From June 1st 2005, he is working as a separation technologist at
Frames Process Systems in Zoeterwoude. The author married Mirjam Kolk on May 18 nd
2001, and on April 22nd 2003 he became father of a son called Lon.

You might also like