Download as pdf or txt
Download as pdf or txt
You are on page 1of 158

Design and scale-up of polycondensation reactors :

hydrodynamics in horizontal stirred tanks and pervaporation


membrane modules
Citation for published version (APA):
Gulik, van der, G. J. S. (2002). Design and scale-up of polycondensation reactors : hydrodynamics in horizontal
stirred tanks and pervaporation membrane modules. [Phd Thesis 1 (Research TU/e / Graduation TU/e),
Chemical Engineering and Chemistry]. Technische Universiteit Eindhoven. https://doi.org/10.6100/IR559776

DOI:
10.6100/IR559776

Document status and date:


Published: 01/01/2002

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

Please check the document version of this publication:


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

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

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
• You may not further distribute the material or use it for any profit-making activity or commercial gain
• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please
follow below link for the End User Agreement:
www.tue.nl/taverne

Take down policy


If you believe that this document breaches copyright please contact us at:
openaccess@tue.nl
providing details and we will investigate your claim.

Download date: 26. Apr. 2023


DESIGN AND SCALE-UP OF
POLYCONDENSATION REACTORS

Hydrodynamics in horizontal stirred tanks and


pervaporation membrane modules

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de


Technische Universiteit Eindhoven op gezag van de
Rector Magnificus, prof.dr. R.A. van Santen, voor een
commissie aangewezen door het College voor
Promoties in het openbaar te verdedigen
op dinsdag 10 december 2002 om 14.00 uur

door

Gerardus Johannes Stefanus van der Gulik

geboren te Uithuizen
Dit proefschrift is goedgekeurd door de promotoren:

prof.dr.ir. J.T.F. Keurentjes


en
prof.dr.ir. W.P.M. van Swaaij

CIP-DATA LIBRARY TECHNISCHE UNIVERSITEIT EINDHOVEN

Gulik, Gerardus J.S. van der

Design and scale-up of polycondensation reactors : hydrodynamics in horizontal stirred tanks and
pervaporation membrane modules / by Gerardus J.S. van der Gulik. – Eindhoven : Technische
Universiteit Eindhoven, 2002.

Proefschrift. – ISBN 90-386-2704-1

NUR 913

Trefwoorden: chemische reactoren ; opschaling / hydrodynamica / membraantechnologie ;


pervaporatie / menging / warmteconvectie / numerieke stromingsleer ; CFD / ultrasone
computertomografie

Subject headings: chemical reactors ; scale-up / hydrodynamics / membrane technology ;


pervaporation / mixing / convective heat transfer / computational fluid dynamics ; CFD /
ultrasonic computer tomography

© Copyright 2002, G.J.S. van der Gulik


Omslagontwerp: Paul Verspaget
Druk: Universiteitsdrukkerij TU/e
Voor mijn ouders
Het in dit onderzoek beschreven onderzoek werd financieel gesteund door het Stan
Ackermans Instituut, Akzo Nobel, Tejin Twaron en de NOVEM.
SUMMARY

Condensation polymers are an important class of polymers, and have already


been produced for a long time. For new processes, reactor design is often a time
consuming process of trial and error, resulting in a reactor, for which little is known
about the occurring physical and chemical conditions. Consequently, the conditions
are difficult to reproduce at a different scale, making the quality of the produced
polymer sensitive to the scale of operation. A proper insight into the hydrodynamics
would enlarge the operating window and provides a route to a more straightforward
scale-up. This includes the Residence Time Distribution, flow patterns, fluid
velocities, and the mixing in the reactor under all flow conditions. The removal of
water during polycondensations is also strongly influenced by the hydrodynamics.
In this thesis, hydrodynamic aspects are studied for the design and scale-up of two
types of reactor for the production of condensation polymers. The first type of
reactor is a horizontal stirred tank reactor for which it is not straightforward to keep
mixing conditions constant upon scale-up. In the second type of reactor, pervapo-
ration membranes are used for the removal of water to favor product formation.
Controlling hydrodynamics is important for increasing the heat and mass transfer
between the membrane surface and the bulk liquid.
Hydrodynamic aspects of the horizontal stirred tank reactor have been studied
experimentally and numerically. The experimental characterisation has been
performed using Planar Laser Induced Fluorescence, Pulse Response Measure-
ments, Power measurements and Laser Doppler Anemometry. These techniques
lead to information on overall circulation, poorly mixed zones, macro-mixing times,
power consumption, velocities and turbulent quantities. Under laminar conditions
the mixing in this reactor appears to be chaotic, which identifies the reactor as a
good laminar mixer. Under turbulent conditions, however, the fluid mainly rotates
like a solid body, which classifies the reactor as a moderate turbulent mixer. Macro-
mixing times have been measured as a function of geometrical parameters and
operation conditions. By correlation, equations are provided for scale-up of the
reactor, while keeping the macro-mixing times constant.
In more detail, the hydrodynamics in the horizontal reactor under turbulent
conditions has been studied numerically using Computational Fluid Dynamics. For
the turbulence modelling, the isotropic k-ε model and the anisotropic Differential
Stress Model have been applied. A comparison of fluid velocities and turbulent
quantities has been made with data from Laser Doppler Anemometry. The tur-
bulence models proved to be able to describe the flow properties equally good.
Passive scalar mixing could only be described properly using the anisotropic
Differential Stress Model. Using these numerical tools, new routes come available
for designing an improved reactor.
Computational Fluid Dynamics has also been used for the design of a
pervaporation membrane reactor. In this type of reactor it is important to reduce
concentration and temperature polarization to obtain high water fluxes during
operation. Polarization can effectively be reduced with secondary flow as induced
by density differences. The secondary flow is found to be most effective in
horizontal set-ups increasing water fluxes up to 50%. The temperature segregation
in the system, as induced by the secondary flow, has been measured using
Ultrasonic Computer Tomography. With this experimental technique the pro-
pagation time of sound waves between several transducers is measured. The
average temperature between the transducers can be calculated as the propagation
time depends on temperature. With Computer Tomography, a 2D-distribution can
be constructed from a number of average temperatures. The constructed 2D-
temperature distribution shows temperature segregation, which is a result of the
secondary flow.
This study is a compilation of numerical and experimental work performed on
two reactors for polycondensation processes. From the work presented in this thesis
it can be concluded that hydrodynamics has a major impact on the design and
development of reactors for polycondensation reactions. CFD has been used as a
numerical tool for studying hydrodynamics and designing reactors. The major
challenge for the future will be to combine the insights in hydrodynamics with
kinetic reaction schemes using CFD. This should lead to new reactor designs and
modes of operation, providing reliable and sustainable processes for the future.
SAMENVATTING

Condensatiepolymeren vormen een belangrijke klasse van polymeren die reeds


gedurende lange tijd geproduceerd wordt. Bij het vergroten van de produc-
tieprocessen, moeten de reactoren veelal worden opgeschaald. Het reactorontwerp
is dan echter vaak een tijdrovend proces van trial and error, resulterend in
reactoren waarvan weinig bekend is over de heersende fysische en chemische
condities. Bij schaalvergroting zijn de condities derhalve moeilijk te reproduceren,
wat leidt tot ongewenste variaties in productkwaliteit. Goed inzicht in de
hydrodynamica zou het werkgebied van de reactor kunnen vergroten en zou het
opschalen kunnen vergemakkelijken. Onder de hydrodynamica wordt onder andere
verstaan de verblijftijdspreiding, de mengpatronen, lokale vloeistofsnelheden en de
menging in de reactor bij verschillende stromingsregimes.
In dit proefschrift worden de resultaten beschreven van het onderzoek naar
hydrodynamische aspecten van twee type reactoren waarin polycondensaties
worden uitgevoerd. Het doel is de verkregen inzichten te gebruiken voor het
opschalen en ontwerpen van de reactoren. Het eerste type reactor is een horizontale
geroerde tankreactor waarbij de nadruk ligt op het voorspellen en controleren van
de menging op verschillende schaalgroten. In het tweede type worden pervaporatie
membranen toegepast voor het verwijderen van water uit het reactiemengsel
waardoor productvorming wordt bevorderd. Het kunnen voorspellen en controleren
van de hydrodynamica in dit type reactoren is belangrijk omdat daarmee de
warmte- en stofoverdracht kunnen worden verbeterd tussen het membraanoppervlak
en de vloeistofbulk.
De hydrodynamische aspecten van de horizontaal geroerde reactor zijn zowel
experimenteel als numeriek onderzocht. Experimenteel onderzoek is uitgevoerd met
behulp de experimentele methoden Planar Laser Induced Fluorescence, puls-
responsie-metingen, Laser Doppler Anemometrie en vermogensmetingen. Deze
methoden leveren inzicht in stromingspatronen, slecht gemengde zones,
macromengtijden, vermogensverbruik, lokale vloeistofsnelheden en turbulente
grootheden. Onder laminaire condities heeft de menging in de reactor een chaotisch
karakter waardoor de menging relatief goed en snel verloopt. Onder turbulente
omstandigheden draait de vloeistof onder invloed van de centrifugaalwerking
voornamelijk rond zijn as als een star lichaam. De menging is dan niet erg effectief.
Macromengijden zijn voor verscheidene condities bepaald als functie van geome-
trische variaties. Door correlatie van de data zijn empirische relaties afgeleid die
kunnen worden toegepast om bij schaalvergroting de macromengtijden te kunnen
voorspellen.
Met behulp van de numerieke techniek Computational Fluid Dynamics, is de
hydrodynamica in meer detail bestudeerd. Voor de turbulente modellering zijn het
isotrope k-ε-model en het anisotrope Differential Stress Model gebruikt. Berekende
vloeistofsnelheden en turbulente grootheden zijn vergeleken met data uit de Laser
Doppler Anenometrie metingen. Beide turbulentiemodellen voorspellen de hoofd-
stromingen even goed. Opmenging van passieve scalairen, hetgeen wordt gebruikt
om de mengsnelheid te kunnen kwantificeren, wordt alleen goed voorspeld met
behulp van het Differential Stress Model. Uitgebreidere toepassing van Computa-
tional Fluid Dynamics technieken zou het ontwerpen van dit type reactoren kunnen
vergemakkelijken en versnellen.
Compuational Fluid Dynamics technieken zijn ook toegepast bij het ontwerpen
van het tweede type reactor; een pervaporatie membraanreactor. In dit type
reactoren is het van belang de concentratie- en temperatuurpolarisatie te reduceren
teneinde hoge waterfluxen te verkrijgen tijdens het productieproces. Polarisatie kan
effectief worden gereduceerd door secundaire stroming die onstaat onder invloed
verschillen in dichtheid in de vloeistof. Het is gebleken dat deze secundaire
stroming het meest effectief is in horizontale opstellingen. Een toenamen van 50%
in de waterflux wordt voorspeld.
Aan temperatuurssegregatie in een horizontale modelopstelling zijn metingen
verricht met behulp van Ultrasone Computer Tomografie. Met deze experimentele
techniek wordt de voortplantingssnelheid gemeten van geluidsgolven tussen
verschillende transducers. De gemiddelde temperatuur tussen twee transducers kan
hiermee indirect worden vastgesteld omdat de voortplantingssnelheid afhankelijk is
van de temperatuur. Met Computer Tomografie kan van een groot aantal
gemiddelde temperaturen een 2-dimensionale temperatuursverdeling gereconstru-
eerd worden. Deze verdeling laat zien dat in de onderzochte horizontale
membraanmodules de voorspelde temperatuurssegregatie inderdaad optreedt.
Dit onderzoek is een compilatie van numeriek en experimenteel werk aan twee
typen reactoren voor de productie van condensatiepolymeren. De hydrodynamica is
van eminent belang voor het goed functioneren en kunnen ontwerpen van beide
type reactoren. Grote vooruitgang kan worden geboekt door uitgebreidere
toepassing van Computational Fluid Dynamics technieken. De grote uitdaging is te
vinden in het combineren van de verkregen inzichten in de hydrodynamica en
kinetische reactieschema’s in Computational Fluid Dynamics. Dit biedt perspectief
voor nieuwe revolutionaire reactor ontwerpen en operationele condities, resulterend
in betrouwbare en duurzame productieprocessen voor de toekomst.
CONTENTS

1. THE NEED FOR CONTROLLING HYDRODYNAMICS


IN POLYCONDENSATION REACTORS 1

2. HYDRODYNAMICS IN A HORIZONTAL STIRRED


TANK REACTOR 15

3. HYDRODYNAMICS AND SCALE-UP OF HORIZONTAL


STIRRED REACTORS 37

4. FLUID FLOW AND MIXING IN AN UNBAFFLED


HORIZONTAL STIRRED TANK 59

5. HYDRODYNAMICS IN A CERAMIC PERVAPORATION


MEMBRANE REACTOR FOR RESIN PRODUCTION 87

6. MEASUREMENT OF 2D-TEMPERATURE DISTRIBUTIONS


IN A PERVAPORATION MEMBRANE MODULE USING
ULTRASONIC COMPUTER TOMOGRAPHY 107

7. FUTURE PERSPECTIVES FOR PROCESS ENGINEERING


OF POLYCONDENSATION REACTIONS 131

DANKWOORD
CURRICULUM VITAE
1
THE NEED FOR CONTROLLING
HYDRODYNAMICS IN
POLYCONDENSATION REACTORS
2 Chapter 1

1.1 Condensation polymers


Condensation polymers are an important class of macromolecular products and
include synthetic materials used as high-strength and/or high-toughness plastics and
fibers (e.g., polyamides, polyesters and polycarbonates) as well as almost all hard
resins (e.g., unsaturated polyesters, epoxy resins, urea-, melamine-, and phenol
formaldehyde resins), thus covering a wide range of applications. Additionally,
members of the same product family are used in relatively small but important
sectors as sealants, elastomers, foams, and adhesive coatings (e.g., silicones, alkyd
resins, and polyimides, Parodi et al., 1989).
Condensation polymerization, also called step-growth polymerization,
involves one or more reactants (monomers) possessing at least two reactive
functional groups (Manaresi et al., 1989). In general, a polycondensation reaction
can be described schematically in the following way:

(n+1) A R1 A + n B R2 B → A R1 ( C R2 C R1 )n A + 2 n Q
← (R.1)

End groups A and B react, forming a group C that becomes part of the polymer
chain and a small molecule Q. The monomer molecules will disappear rapidly, and
consequently, after a while only chains of monomers are coupling with other
chains. In principle, the degree of polymerization is determined by a small excess of
one of the monomers. When the relative excess equals 1/n, the average chain will
contain n units1.

n/(n+1)
1000 1.0

0.995
Pn [-]

0.99
100
0.975
0.95
0.925
0.9

10
0.9 0.92 0.94 0.96 0.98 1
Xc [-]

Figure 1.1: Average degree of polymerization as a function of


conversion Xc at different initial stoichiometric ratios n/(n+1).

1
The excess origins from the ratio of reactants in reaction R.1: (n+1)/n = 1 + 1/n
The need for controlling hydrodynamics in polycondensation reactors 3

The conversion, Xc, can be expressed as the degree of conversion of end


groups, either based on A or B. In Figure 1.1, the number-averaged degree of
polymerization 3n has been plotted against the degree of conversion Xc for several
initial stoichiometric ratios n/(n+1). This figure shows that to obtain 100 units per
chain with an initial stoichiometric ratio of 0.99, Xc has to be 0.995. Thus, to obtain
long polymer chains both the stoichiometric ratio and Xc have to be close to unity.
Often such long polymer chains are required to achieve the desired product quality.
4 Chapter 1

1.2 Reaction engineering of polycondensations


The very high conversions that are required often lead to major technical
problems (Thoenes, 1994; Vollbracht, 1989). For example, the small molecule Q
(often water) may have to be removed to force the reaction to completion
(Carothers, 1936; Sawada, 1976). This can be achieved by evaporation or chemical
bonding. The removal of the small molecule from a viscous molten polycondensate
is difficult as the surface tension and the viscosity of the liquid hinder the
nucleation of vapor bubbles. Even when bubbles are formed, they have such a small
volume that their rising velocity is extremely low.
To obtain high conversions, the Residence Time Distribution (RTD) has to be
very narrow (Thoenes, 1994; Westerterp et al., 1984; Biesenberger et al., 1983).
Therefore, most often a batch reactor is used. A disadvantage of a batch reactor is
that the viscosity of the reactor content increases dramatically during the process.
This usually induces flow conditions to change from turbulent to laminar. A second
disadvantage of a batch reactor is that at the beginning of the polycondensation
process the reaction rate is high as the concentration of end groups is highest.
Consequently, most heat of reaction has to be removed at the beginning of the
process. It is difficult to remove all the heat within a short period of time from a
large batch reactor because the surface to volume ratio is relatively low. Therefore,
external cooling has to be applied or the scale of production has to be reduced. A
number of reactors in series would sometimes be highly desirable, because different
types of impeller can be used during different stages of the process and the heat
production can be distributed over several reactors. However, a series of reactors
can only be used when the RTD shows exact plug flow behavior or can be kept
extremely narrow.
For a polycondensation process it is difficult to find a suitable reactor, because
the various requirements are often contradictory. The route to a successful
polycondensation reactor is a time-consuming process of trial and error. From the
resulting reactor, frequently little is known about the physical and chemical
conditions. Consequently, these conditions are difficult to reproduce at a different
scale, making the product quality sensitive to the scale of operation. However,
successful operation upon scale-up can sometimes be maintained when several
reactor-engineering aspects are known. Such aspects, for which generally the
collective term ‘hydrodynamics’ is used, comprise the RTD of a reactor system, the
flow patterns, the mixing under various flow conditions, and the mixing power. In
this thesis, hydrodynamic aspects are studied for the design and scale-up of two
types of reactor for the production of condensation polymers. The first type of
reactor is a horizontal stirred tank reactor for which it is not straightforward to keep
the mixing conditions constant upon scale-up. In the second type of reactor
pervaporation membranes are used for the removal of the byproduct water.
The need for controlling hydrodynamics in polycondensation reactors 5

1.3 Tools for studying hydrodynamics


For the optimization of polymerization reactors, characterization of the
occurring hydrodynamics is of prime importance. Two characterization methods are
available: Experimental and Computational Fluid Dynamics (known as EFD and
CFD, respectively). EFD is a collective term for several experimental techniques
used to characterize the hydrodynamics. In this thesis the following experimental
techniques have been used:

• Planar Laser Induced Fluorescence (PLIF) – With PLIF a fluorescent


dye can be traced in a 2-dimensional laser sheet. Using a high-speed
camera, a digital film can be made of the mixing pattern of the tracer in the
reactor (Chapter 2; Miller, 1981; Owen, 1976).
• Pulse Response Measurements – With Pulse Response Measurements an
inert tracer is injected in a batch reactor while at a different position the
concentration of the tracer is monitored as a function of time. From these
measurements macro-mixing times can be deduced (Chapter 3; Westerterp
et al., 1984).
• Laser Doppler Anemometry (LDA) – LDA is an optical method for fluid
flow research based on a combination of interference and Doppler effects.
LDA allows the measurement of the local, instantaneous velocities of
particles suspended in the flow. LDA has a high resolution power in time
and is non-invasive (Chapter 4; Durst et al., 1981).
• Ultrasonic Computer Tomography (U-CT) – U-CT is based on the
dependence of the propagation velocity of ultrasound on the temperature of
a medium. Over a line between a speaker and a microphone the velocity of
sound is measured and converted into an average line temperature. By
measuring a large number of lines in a plane and using Computer
Tomography, a 2-dimensional temperature distribution can be constructed.
The technique is non-invasive (Chapter 6; Norton et al., 1984; Peyrin et al.,
1983).
• Magnetic Resonance Imaging (MRI) – With MRI the proton density in a
medium can be measured over a line. By measuring a large number of lines
in a plane and using advanced computational techniques, a 2-dimensional
proton density distribution can be constructed. This distribution can
provide information on the location of different chemicals. The technique
is non-invasive. Additionally, MRI can be used to measure local velocities
by applying pulse field gradients (Chapter 7; Hornak).

CFD is a numerical tool for studying hydrodynamics. Fluid flow, mixing, heat
and mass transfer in a prescribed geometry can be calculated on a computer. For
this, several commercial packages are available, including CFX (AEA Technology,
Harwell, UK), Fluent (Fluent Inc., Lebanon, USA), Star-CD (Computational
Dynamics Ltd., London, UK). Only CFX has been used, in which both laminar and
6 Chapter 1

turbulent conditions can be handled. For turbulent flows, the Reynolds-Averaged


Navier-Stokes equations can be solved with the use of appropriate turbulence
models.

TDC
H 2O
PPD

Figure 1.2: Flow state and initial location of the


reactants in the Twaron® polymerization process.
The need for controlling hydrodynamics in polycondensation reactors 7

1.4 Scale-up of a Horizontal Stirred Tank Reactor.


The horizontal stirred tank reactor studied in this thesis is of the Drais type
and is further referred to as the Drais reactor. The aromatic polyamide Twaron is
produced in this type of reactor (Vollbracht, 1989; Banneberg-Wiggers et al., 1998).
During the reaction, the flow regime changes from turbulent to laminar, due to a
tremendous increase in viscosity. In both regimes mixing has to be sufficient as it
has a large influence on the final product quality. The Twaron polyamide PPTA
(Para-Phenylene TerphthalAmide) is produced via the condensation reaction of
PPD (Para-Phenylene Diamide) with TDC (Terephthaloyl DiChloride). The fast
propagation step is given as reaction (R.2):

O O H H O O H
(n+1) H2N NH2 + n Cl C C Cl H N N C C N NH2 + n HCl (R.2)
n

(PPD) (TDC) (PPTA or Twaron® polymer)

The degree of polymerization is controlled by adding a small amount of water


that can terminate a reactive acylchloride group via the slow reaction (R.3):

O O
R3 C Cl + H 2O R3 C O H + H Cl (R.3)

Water is added at the start of the process because at the end of the process the
viscosity is too high for mixing the water sufficiently with the reactor content. The
initial presence of water, however, complicates the process, as it is able to terminate
growing chains too early when mixing is insufficient, leading to short chains and a
broad molecular weight distribution (MWD). The set of reactions (R.2 and R.3) is
usually described as competitive-parallel.
In production, the reactor is partially filled with the diamide component
(containing the required amount of water), to which the diacylchloride component
is added semi-batchwise. To allow for an exact stoichiometric ratio of the two
monomers, only one injection point is used, as schematically depicted in Figure 1.2.
The use of a single injection point implies that a good overall circulation is needed
to allow the acylchloride molecules to react with all the amide-containing molecules
throughout the reactor before termination with water occurs. Thus, a proper insight
into the flow pattern, the mixing, and the mixing times are mandatory to guarantee
constant product quality upon scale-up and to allow for quality improvements in
existing equipment. These hydrodynamic aspects will be treated in the chapters 2 to
4 of this thesis.
8 Chapter 1

1.5 Design of the membrane reactor


Alkyd resins are low molecular weight polyesters, formed by the equilibrium
reaction between a di-alcolhol and a di-acid, as given in (R.4):

n R1 OH + n R2 COOH Resin + H2O (R.4)

Reaction conditions require elevated temperatures, typically above 200 °C. At these
temperatures, the mixtures are moderately viscous (µ = 5-25 mPa·s). To obtain high
yields, the reaction equilibrium has to be shifted to the right. This can be achieved
by using a large excess of the alcohol (usually as solvent) and by removing the
water. In current processes, water is removed afterwards or during the reaction by
distillation (Figure 1.3A), which is not very efficient in the case of azeotrope
formation. Additionally, due to the large reflux ratios required, energy consumption
can be significant (Keurentjes et al., 1994). In this operation the energy input can be
10 times larger than the energy required for the removal of water.

Pervaporation
membrane
H2O module
Distillation unit Reactants
Resin
H2O

B Reactor

Reactants
Resin A C Pervaporation
membrane reactor H22O
Reactants
Reactor

Resin

Figure 1.3A-C: A) Current batch process for the production of resins with a distillation unit
for the removal of water. B) Improved batch process in which the distillation unit has been
replaced with a pervaporation membrane module. C) Continuous process in which the
reactor and the distillation unit have been replaced with one pervaporation membrane
reactor, running in a once-through continuous mode.
The need for controlling hydrodynamics in polycondensation reactors 9

Membrane

Water concentration

Temperature

Arbitrary units Water

Viscosity

Resin concentration
δT δC

Figure 1.4: Concentrations, temperature and viscosity in


arbitrary units near the membrane surface.

With ceramic pervaporation membranes, water can be removed selectively


through the membrane by evaporation (Ho et al., 1992; Nunes et al., 2001;
Rautenbach et al., 1984, 1985; Bakker et al., 1998; Koukou et al., 1999; Verkerk et
al., 2001; Jafar et al., 2002). These membranes are selective to water as they are
extremely hydrophilic and have pores that are about the kinetic diameter of water.
Additionally, they are resistant to high temperatures. Use of these membranes
allows for two alternative process schemes. In Figure 1.3B, a pervaporation
membrane module is used instead of the evaporator. This bath-wise concept in
Figure 1.3B is relatively easy to implement in current production processes, because
only one unit operation has to be replaced. In Figure 1.3C, both the reactor and
evaporator have been replaced by a membrane reactor. This concept is less easy to
implement but has advantages, as it can be operated in continuous mode, which
makes it easier to control the product quality. In both concepts, the reduction in
energy consumption should be significant as, in principle, the only energy required
is used for the evaporation of water.
It can be anticipated that concentration and temperature polarization near the
membrane represent a major problem. The polarization effects are schematically
depicted in Figure 1.4. As water is transported through the membrane, the water
concentration near the membrane surface will be low. As a result the resin
concentration will be high, as the reaction will locally be in equilibrium. Also, the
temperature will be low as water evaporates at the membrane surface adiabatically.
Consequently, viscosity near the membrane will be relatively high. As a result, a
relatively thick stagnant layer can be formed, severely reducing water transport
from the bulk to the membrane surface.
Optimizing the flow conditions near the membrane can reduce the occurring
polarization effects. For the two process options given, different flow conditions are
possible. For the process in Figure 1.3B, polarization can be reduced by applying
10 Chapter 1

high turbulence levels. For example, a powerful pump can recycle the liquid with
high velocities over the membrane. However, this will reduce the energy savings.
Using a powerful pump is not appropriate in the process depicted in Figure 1.3C,
because velocities have to be low in order to limit equipment dimensions.
Additionally, in Figure 1.3C nearly plug flow characteristics are required to obtain a
narrow molecular weight distribution (MWD).
It will be difficult to design a membrane reactor according to the process in
Figure 1.3C with a flow pattern that reduces the polarization effects effectively,
while fulfilling the requirements. In this thesis, the application of buoyancy forces
is considered for obtaining a flow pattern that will effectively reduce polarization
effects. Also, a relatively new experimental technique has been implemented for
studying accompanying temperature effects.
The need for controlling hydrodynamics in polycondensation reactors 11

1.6 Outline of this thesis


This thesis describes the hydrodynamic aspects occurring in the two types of
polycondensation reactor using both EFD and CFD. For the horizontal stirred tank
reactor, fluid flow and mixing will be described both quantitatively and
qualitatively. Insight into these hydrodynamic aspects will lead to efficient scale-up
of the reactor in such a way that the product quality can be kept constant upon
scale-up. For the membrane reactor the characterization of the flow and polarization
effects should lead to the design of a once-through continuous reactor in which flow
conditions are laminar.

In Chapter 2, flow patterns of inert tracers under turbulent and laminar


conditions are described. PLIF is used as the experimental technique. Experiments
have been performed in three small-scale models of the Drais reactor. Mixing times
have been determined, leading to guidelines for scale-up.

In Chapter 3, the power input upon agitation has been measured in order to
classify the configuration relative to other standard configurations. The energy input
also provides the energy dissipation rate ε, which is crucial for defining well-mixed
and poorly mixed regions. Mixing times have been measured using Pulse Response
Measurements. Combined with the power measurements this leads to insight into
the mixing efficiency as a function of scale.

Chapter 4 is the final chapter on the Drais reactor describing LDA


measurements under turbulent conditions. Mean and fluctuating velocities provide a
general insight into the velocity scales and mixing processes occurring in the Drais
reactor. LDA and PLIF measurements are compared with CFD calculations. To
incorporate the turbulent properties, several turbulence models are available in
CFD. Comparison shows that for a correct prediction of the PLIF experiments the
selection of the appropriate turbulence models is the crucial parameter in CFD.

Chapters 5 and 6 concern the pervaporation reactor. Chapter 5 contains a


CFD study in which the effect of the orientation of a single membrane tube with
respect to the gravity field has been studied. For these cases the relevance of
buoyancy effects at laminar conditions has been examined. As an additional
parameter the superficial velocity has been varied. Heat and mass transfer rates
provide the essential parameters to distinguish the most optimal configuration.

Chapter 6 describes the successful implementation of U-CT (Ultrasonic


Computer Tomography) in a pervaporation membrane module. With this relatively
new experimental technique, 2-dimensional temperature distributions can be
measured. The measured temperature distributions are compared with CFD
calculations.
12 Chapter 1

A review will be given in Chapter 7. Additionally, the implications for scale-


up and future designs will be discussed.

The set-up of this thesis is such that each chapter can be read separately.
Consequently, some information will be repeated at more than one location. This
approach has been chosen deliberately in order to enable the reader to go only
through specific chapters of interest and to avoid a long list of references to other
parts of this thesis.
The need for controlling hydrodynamics in polycondensation reactors 13

References
Bakker, W.J.W.; Bos, I.A.A.C.M.; Rutten, W.L.P.; Keurentjes, J.T.F.; Wessling, M.;
“Application of ceramic pervaporation membranes in polycondensation reactions”, Int.
Conf. Inorganic Membranes, Nagano, Japan, 1998, 448-451.
Bannenberg-Wiggers, A.E.M.; Van Omme, J.A.; Surquin, J.M.; “Process for the batchwise
preparation of poly-p-terephtalamide”, U.S. Pat., 5,726,275, 1998.
Biesenberger, J.A.; Sebastian, D.H.; “Principles of Polymerization engineering”, John Wiley
and Sons, New York, 1983.
Carothers, W.H.; Trans. Faraday. Soc., 1936, 32, 39.
Durst, F.; Melling, A.; Whitelaw, J.H.; “Principles and Practice of Laser Doppler Anemo-
metry”, Academic Press, London, UK, 1981.
Ho, W.S.W.; Sirkar, K.K. (Eds.); “Membrane Handbook”, Chapman & Hall, New York,
1992.
Hornak, J.P.; “The basics of MRI”, http://www.cis.rit.edu/class/schp730/bmri/bmri.htm.
Jafar, J.J.; Budd, P.M.; Hughes, R.; “Enhancement of esterification reaction yield using
zeolite A vapour permeation membrane”, J. Membr. Sci., 2002, 199 (1-2), 117-123.
Keurentjes, J.T.F.; Janssen, G.H.R.; Gorissen, J.J.; “The esterification of tartaric acid with
ethanol: kinetics and shifting the equilibrium by means of pervaporation”, Chem. Eng.
Sci., 1994, 49, 4681-4689.
Koukou, M.K.; Papayannakos, H.; Markatos, N.C.; Bracht, M.; Van Veen, H.M.; Roskam,
A.; “Performance of ceramic membranes at elevated pressure and temperature: effect of
non-ideal flow conditions in a pilot scale membrane separator”, J. Membr. Sci., 1999,
155, 241-259.
Manaresi, P.; Munari, A.; “Factors affecting rate of polymerization”, Comprehensive
Polymer Science, Step Polymerization, 1989, 5, 35.
Miller, J.N.; “Standard in fluorescence spectrometry”, Chapman & Hall, London, UK, 1981.
Norton, S.J.; Testardi, L.R.; Wadley, H.N.G.; “Reconstruction internal temperature distri-
butions from ultrasonic time-of-flight tomography and dimensional resonance
measurements”, J. Res. Natl. Bur. Stand., 1984, 89(1), 65-74.
Nunes, S.P.; Peinemann, K.-V. (Eds.); “Membrane Technology in the Chemical Industry”,
Wiley-VCH, Weinheim, 2001.
Owen, F.R.; “Simultaneously laser measurements of instantaneous velocity and concentra-
tion in turbulent mixing flows”, AGARD-CP193, Paper No. 27, 1976.
Parodi, F.; Russo, S.; “Polycondensation and Related Reactions”, Comprehensive Polymer
Science, Step Polymerization, 1989, 5, 1.
Peyrin, F.; Odet, C.; Fleischmann, P.; Perdrix, M.; “Mapping of internal material tempera-
ture with ultrasonic computed tomography”, Ultrason. Imag., Conference Proceeding,
July 1983, Halifax, 31-36.
Rautenbach, R.; Albrecht, R.; “On the behavior of asymmetric membranes in
pervaporation”, J. Membr. Sci., 1984, 19, 1-22.
Rautenbach, R.; Albrecht, R.; “The separation potential of pervaporation. Part 2. Process
design and economics”, J. Membr. Sci., 1985, 25, 25-54.
Sawada, H.; “Thermodynamics in polymerization”, Chapter 6, Marcel Dekker, New York,
1976.
Thoenes, D.; “Chemical Reactor Development”, Kluwer Academic Publishers, Dordrecht,
The Netherlands, 1994.
Verkerk, A.W.; Van Male, P.; Vorstman, M.A.G.; Keurentjes, J.T.F.; “Description of
dehydration performance of amorphous silica pervaporation membranes”, J. Membr.
Sci., 2001, 193(2), 227-238.
14 Chapter 1

Vollbracht, L.; “Aromatic Polyamides”, Comprehensive Polymer Science, Step Polymeriza-


tion, 1989, 5, 374.
Westerterp, K.R.; Van Swaaij, W.P.M.; Beenackers, A.A.C.M.; “Chemical Reactor Design
and Operation”, John Wiley and Sons, New York, 1984.
2
HYDRODYNAMICS IN A HORIZONTAL
STIRRED TANK REACTOR

Abstract

In this study, the hydrodynamics in a horizontal stirred tank


reactor is investigated. This type of reactor is used in industry for fast
polycondensation processes. Overall circulation, poorly mixed zones
and macro-mixing times are determined in scale models under
turbulent (Re > 105) and laminar (Re < 300) conditions using planar
laser induced fluorescence. At both sets of conditions, the observed
overall circulation is complex and changes when the length-to-
diameter ratio is varied. Under laminar conditions, the flow appears to
be chaotic. The poorly mixed zones change in location, number, and
life span at different length-to-diameter ratios. Dimensionless macro-
mixing times under turbulent conditions are correlated with
parameters variations and show nonlinear relationships in fill ratio,
length-to-diameter ratio, and Reynolds number. Under laminar
conditions, macro-mixing times could not be determined
unambiguously, but they are only 2.5 times larger than under turbulent
conditions.

This chapter is a slightly modified version of the publication:


Van der Gulik, G.J.S.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodynamics in a Horizontal
Stirred Tank Reactor”, Ind. Eng. Chem. Res., 2001, 40(3), 785-794.
16 Chapter 2

2.1 Introduction
The design of a reactor for fast polycondensations is a major challenge for
chemical engineers, as often several conflicting needs have to be fulfilled.
Generally, two types of agitation are needed as the flow regime changes from
turbulent to laminar, because of a tremendous increase in viscosity. In both regimes
the mixing has to be sufficient as it has a large influence on the final product quality
(Thoenes, 1994; Manaresi et al., 1989). This influence can easily be understood by
considering the production process of Twaron, an aromatic polyamide produced
via the polycondensation reaction of a diamide with a diacyl chloride (Gaymans et
al., 1989; Vollbracht, 1989).The fast propagation step is given as reaction R.5:
O O
(n+1) H2N R1 NH2 + n Cl C R2 C Cl
(R.5)
H H O O H H
H N R1 N C R2 C N R1 N H + n H Cl
n

Theoretically, a small excess of the diamide determines the degree of


polymerization. When the relative excess equals 1/n, the average chain will contain
n+1 units. In practice, the degree of polymerization is controlled by adding water
that can terminate a reactive acyl chloride group via the slow reaction R.6:

O O
R3 C Cl + H 2O R3 C O H + H Cl (R.6)

Water is added at the start of the process because, at the end of the process, the
viscosity is too high to mix water sufficiently to molecular scale. The initial
presence of water complicates the process, as water is able to terminate chains too
early when mixing is insufficient, leading to short chains and a broad molecular
weight distribution (MWD).
Table 2.1: Order of magnitude of reaction time for reactions R.5 and R.6 (Jeurissen et al.;
Borkent, 1976).
Viscosity [Pa·s] Reaction time for propagation R.5 [s] Reaction time for termination R.6 [s]
10-3 10-4 102
10 10 102

In Table 2.1, the orders of magnitude of the reaction half-life times for both
reactions are given for low and high viscosity levels. Going from low to high
viscosity, the propagation rate slows by 5 orders of magnitude, while the
termination rate remains unchanged. The propagation reaction is slowed down for
two reasons. Firstly, RNH3+Cl--groups are formed, which are less reactive than
Hydrodynamics in a Horizontal Stirred Tank Reactor 17

Front Injection
L Side
C

D
β

h
Hatch

Top

w
α

Figure 2.1: The Drais reactor, given in front, top and side views with length L, diameter
D, blade angles α and β, blade width w, blade height h, and a hatch. The arrows point
out the direction of rotation during operation.

NH2-groups (Gaymans et al., 1989). Secondly, the reactive end groups are more
sterically hindered upon an increase in molecular weight. The chain stopper (water)
will not be hindered as much, because it is a relatively small molecule.
In production, the reactor is filled with the diamide-component (containing the
required amount of water), to which the diacyl-component is added semi-batchwise.
To allow for an exact stoichiometric ratio of the two monomers, only one single
injection point is used. This implies that good overall circulation is needed to allow
the acyl molecules to react with all the amide-containing molecules throughout the
reactor before termination occurs. This implies that a fundamental insight into the
hydrodynamic behavior of this type of reactor is mandatory to guarantee constant
product quality upon scale-up and to allow for quality improvements in existing
equipment.
The polymerization described above is performed in a horizontal stirred tank
reactor of the Drais type (Vollbracht, 1989). This multifunctional reactor, as
depicted in Figures 2.1 and 2.2, can be used for powder mixing, turbulent fluid
mixing and kneading at high viscosities with an energy dissipation up to 200 W/kg.
Literature on the hydrodynamics in horizontal mixing vessels is very limited
compared to the literature on vertical vessels. There is some literature on turbulent
mixing, but literature on laminar mixing is absent. Ando et al. (1971a) studied
power consumption and flow behavior under turbulent conditions in an unbaffled
horizontal vessel with Rushton turbine impellers with Di/D = 0.9. They distinguish
two flow states A and B. State A is obtained at a relatively low stirrer speed. The
18 Chapter 2

liquid is then pushed up by the impellers and sprayed, leading to the formation of
fine liquid droplets and fine air bubbles. State B, the so-called hollow state, is
obtained at a higher stirrer speed, providing a ring of fluid. As their research mainly
focused on applications to gas-liquid absorption, mainly state A in baffled vessels
was investigated (Ando et al., 1971b, 1974, 1981; Fukuda, 1990). This is due to a
larger gas-liquid interface in the A state compared to state B. Ando et al. (1990)
also studied turbulent mixing in an idealized horizontal vessel with baffles and
multiple impellers. Macro-mixing times were measured, and a model was proposed
for predicting them. It was established that the dimensionless macro-mixing time
N⋅t is proportional to L/D. The information available in the literature on turbulent
mixing in horizontal stirred tank reactors is not directly applicable to
polycondensations in the Drais reactor for two reasons. First, the impeller geometry
is completely different. Second, the fluid in the polycondensation process is in the
hollow state (or the B state according to Ando et al., 1971a) because of a high stirrer
speed.
Therefore, we conducted an experimental study on the hydrodynamics in this
type of reactor. For this purpose, mixing patterns, the life span of poorly mixed
zones, and the macro-mixing time have been established experimentally. This is
done for turbulent as well as laminar conditions for different reactor fill ratios.
Subsequently, scaling rules will be defined based on these macro-mixing times.
Hydrodynamics in a Horizontal Stirred Tank Reactor 19

Figure 2.2: Stirrer geometry for reactor-15 (top) and reactor-20 (bottom). The white
blades transport fluid to the right during rotation, the black stirrers transport fluid to
the left. Thus, both have a pumping action toward the center of the reactor. The
blades are evenly distributed over the shaft and are not drawn in perspective.
20 Chapter 2

2.2 Experimental Section


The reactor used in this study was a horizontal stirred tank of the Drais type
(Turbulent Schnellmischer, Drais Ltd, Mannheim, Germany), as depicted in Figures
2.1 and 2.2. Typical for the unbaffled cylindrical reactor is its horizontal position
and the heavily designed impeller. This impeller can provide a high mixing power,
needed to achieve sufficient mixing under highly viscous conditions. The reactor is
characterized by length L and diameter D. Typical for the reactor is the clearance C,
the distance between the blades and the reactor wall. For a small clearance, the
blades perform a scraping action that keeps the reactor walls free from polymer
material and also provides good heat exchange with the cooled walls. The blades
have a pumping action towards the reactor center, providing an easy way to empty
the reactor through the opened hatch.
To determine macro-mixing times and the life span of poorly mixed zones,
planar laser induced fluorescence (PLIF) was used. With PLIF, it is possible to
make a digital film of the mixing of a tracer in a 2-dimensional plane in which the
poorly mixed zones can easily be located. In contrast with other studies (Distelhoff
et al., 1997; Kersting et al., 1995; Mayr et al., 1994; Moo-Young et al., 1972;
Perona et al., 1998). PLIF is unobtrusive, has a small measurement volume, and is
flexible in changing the monitoring point, as only the position of the laser sheet has
to be changed. PLIF experiments were performed in three small-scale models of the
Drais reactor. These reactors, as depicted in Figures 2.1 and 2.2, were all 0.18 m in
diameter but differed in length, being 0.20, 0.27, and 0.36 m, providing L/D ratios
of 1.1, 1.5, and 2.0, respectively. From this point on, these scale-models are referred
to as reactor-11, reactor-15, and reactor-20, respectively. The blade width w was
equal to 0.1 m, the blade height h was equal to 0.015 m, and the shaft had a
diameter of 0.03 m. The blades were evenly distributed over the shaft for each
reactor. The mutual angle of the blades was 180, 120, and 135° for reactor-11,
reactor-15, and reactor-20, respectively. These angles correspond to industrial
configurations and proved to provide the most stable fluid ring in partially filled
reactors. The sidewalls, the shaft, and the impeller blades were made of stainless
steel, and the cylindrical wall of glass. Under turbulent conditions, tap water was
used as the reactor content for the three fill ratios 40, 60, and 100%. The rotational
speed varied between 3.8 and 11.6 Hz, resulting in turbulent flow with Reynolds
numbers ranging from 123,000 to 375,000, as defined by ρNDi2/µ. When the
reactor was partially filled, the rotational speeds always resulted in the hollow state
or the B state according to Ando et al. (1971a). For laminar conditions, glycerin
(Heybroek, Amsterdam; purity >99.9%) was used at Reynolds numbers ranging
from 90 to 270. Glycerin limits the use of PLIF to the examination of completely
filled reactors because air bubbles lead to an untransparent fluid in partially filled
reactors.
Figure 2.3 shows the experimental arrangement for the PLIF experiments,
which is comparable to the set-up used by Schoenmakers et al. (1997). The laser
beam was generated by a 2-W Ar/Kr-laser (model Stabilite 2017-005, Spectra
Hydrodynamics in a Horizontal Stirred Tank Reactor 21

Lens
δ(t)

x
Region of
Laser .
interest

Lasersheet

Camera
PC

Figure 2.3: Sketch of the experimental set-up.

Physics) and had a wavelength of 488 nm. The beam was converted to a laser sheet
with a thickness of 0.5 mm by a cylindrical lens (Dantec 9080XO.21). The position
of the laser sheet in the vessel geometry was always parallel to the shaft. Therefore,
the observed mixing process was always the mixing in the axial and radial
directions. This is secondary mixing, superimposed on the mixing in the tangential
direction.
Disodium fluorescein (C20H10O5Na2) was used as the fluorescent dye (Merck,
Darmstadt; purity >98wt%). This dye emits light with an intensity depending on the
power of the laser light, the concentration, and the pH of the solvent. The power of
the laser light was kept constant at 0.4 W. During every experiment, the final dye
concentration was around 10-7 M. Therefore, the amount of injected solution, with
dye concentration of 2·10-3 M, varied between 0.3 and 0.5 mL, depending on the
reactor volume and fill ratio. The pH of both the injected solution and the reactor
content was kept constant at a value of 10, as the intensity of the emitted light is
independent of the pH at pH > 8.
A high-speed camera (JAI CV-M30), connected to a PC, with an EISA
compliant frame grabber (Magic) recorded the light that was emitted by the
fluorescein molecules in the laser light plane. The commercial software package
DMA-MAGIC was used for data acquisition. The value of the recorded gray scales
ranged from 0 to 255, providing a resolution of 256 values. For complete mixing, a
gray scale of around 150 was obtained. In the range from 0 to 255 the gray scale
corresponds linearly with concentration. In Figure 2.3, the evaluated region is
depicted as a dotted rectangle. This region was always set to the left part of the
reactor. It was sufficient to monitor mixing in one half of the reactor because we
observed that mixing was symmetrical with respect to the reactor center. The
number of recorded images per second ranged between 30 and 120 and depended
on impeller speed and expected mixing time. The number of pixels per image in the
axial direction, i.e., from injection point to sidewall, ranged from 90 to 150 pixels
22 Chapter 2

for reactor-11 and reactor-20, respectively. In the radial direction, i.e., from shaft to
cylindrical wall, the number was 68 for every image. This results in a spatial
resolution of about 1×1×0.5 mm per pixel.
Hydrodynamics in a Horizontal Stirred Tank Reactor 23

2.3 Results and Discussion


This section starts with a brief description of the mixing pattern and the
establishment of poorly mixed zones or islands, as observed in the PLIF images.
Then, the technique for measuring concentrations and mixing times is shown.
Finally, the mixing times are correlated with process parameters in order to
formulate empirical correlations for scale-up.

2.3.1 Mixing Patterns and Chaotic Mixing


The mixing of the injected dye in the axial and radial direction in reactor-11 is
shown in Figures 2.4 and 2.5 for turbulent and laminar conditions, respectively.
Although not the same, these mixing patterns show a resemblance. In Figures 2.4b
and 2.5a, it can be seen that the dye is mainly transported in the axial direction as it
flows from the central injection point towards the sidewall. Figures 2.4d and 2.5b
show that the dye is subsequently transported in the radial direction along the
sidewall, followed by transport to the bulk along the shaft in the axial direction. The
overall circulation in Figures 2.4 and 2.5 is therefore turning counterclockwise. This
is the opposite of what was expected, based on the center-oriented pumping action
of the stirrer blades.

2.4a: 0.35 s. 2.4b: 0.55 s. 2.4c: 0.93 s.

2.4d: 1.65 s. 2.4e: 2.75 s. 2.4f: 5.6 s.

Figure 2.4: Mixing pattern in water in reactor-11 at N = 6 Hz.


24 Chapter 2

2.5a: 0.67 s. 2.5b: 1.0 s. 2.5c: 1.5 s.

2.5d: 2.37 s. 2.5e: 4.0 s. 2.5f: 7.4 s.


Figure 2.5: Mixing pattern in glycerin in reactor-11 at N = 8 Hz.

2.6a: 0.48 s. 2.6b: 0.79 s. 2.6c: 1.50 s.

2.6d: 2.50 s. 2.6e: 4.01 s. 2.6f: 9 s.


Figure 2.6: Mixing pattern in water in reactor-20 at N = 6 Hz.

2.7a: 0.7 s. 2.7b: 1.5 s. 2.7c: 2.2 s.

2.7d: 3.1 s. 2.7e: 7.7 s. 2.7f: 8.0 s.


Figure 2.7: Mixing pattern in glycerin in reactor-20 at N = 6 Hz.
Hydrodynamics in a Horizontal Stirred Tank Reactor 25

Figures 2.6 and 2.7 show how the dye is mixed in reactor-20. The PLIF images
show that two regions are present that differ in overall circulation: one at the left-
hand side of the PLIF image and one at the right-hand side. The circulation at the
left-hand side is counterclockwise and thus shows resemblance with the circulation
in reactor-11. The circulation at the right-hand side is clockwise. This circulation is
visible in Panels a and b of Figures 2.7 in which the injected dye is transported first
in the radial direction towards the shaft and subsequently in the axial direction
along the shaft. By way of illustration, the generalized overall circulation in the
radial and axial directions is depicted in Figure 2.8 for reactor-11, reactor-15 and
reactor-20.
The flow in Figures 2.5 and 2.7 can be considered as a 3-dimensional
discontinuous periodic flow in which the impeller blades provide the discontinuous
movement with a period equal to 1/N. This allows a comparison with the two-
dimensional flow in a cavity as studied by Leong and Ottino (1990) and Ottino
(1991). Clearly, the laminar flow has a chaotic nature as it is capable of stretching
and folding a region of fluid and returning it - stretched and folded - to its initial
location after one period, i.e., one impeller revolution. Furthermore, the formed
striations are reoriented when the impeller blades cross them. This event is visible
in Figure 2.5d, in which the vertical impeller arm, as present in the right-hand side
of the laser sheet, plows through the horizontal striations. These reorientations
further enhance chaotic mixing.

Figure 2.8: Observed overall circulation in reactor-11, reactor-15, and reactor-20.


26 Chapter 2

3
100

80 Reactor-20
Perimeter [cm]

2.5 60 N = 11 Hz
40
Liapunov-exponent σ [s ]
-1

20
2 0
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
t [s]

1.5

1 Reactor-11
σ = 0.22N Reactor-15
0.5 Reactor-20

0
0 2 4 6 8 10 12
N [Hz]
Figure 2.9: Liapunov exponents σ as a function of impeller frequency N. The inset provides
the perimeter against time for reactor-20 at 11 Hz. The curve represents PE=PE0·exp(σ·t).

A quantitative indication for chaotic behavior is the Liapunov exponent σ in


the formula PE=PE0·exp(σ·t). This formula (Leong and Ottino, 1989; Ottino, 1991)
represents the stretching rate of the flow as it describes the perimeter of
intermaterial area between the dye and the clear fluid as a function of time. A
positive exponent σ implies exponential generation of intermaterial interface and,
hence, implies chaotic flow. The inset in Figure 2.9 gives the perimeter against for
reactor-20 at N=11 Hz. The initial exponential increase yields the Liapunov
exponent by a fit procedure for which all the perimeter values after the maximum
has been reached are ignored. The perimeter decreases after the maximum has been
reached because the striations are lost when they become smaller than the pixel
size. For all experiments under laminar conditions, the Liapunov exponent is given
in Figure 2.9 as a function of impeller speed. The positive exponents increase
linearly with impeller speed and appear to be independent of L/D. This indicates
that chaotic mixing in all three reactors occurs in a similar fashion.

2.3.2 Poorly Mixed Zones and Islands


Under turbulent conditions, poorly mixed zones are visible in Figures 2.4 and
2.6 as the areas in the reactors that remain dark the longest. These areas are visible
in Figures 2.4e and 2.6e (reactor-11 and reactor-20), near the shaft between the two
outer impeller blades (the same holds for reactor-15, although not shown here).
Apparently, under turbulent conditions the location does not depend on L/D.
Nevertheless, the poorly mixed zones are not very stagnant, as they disappear
within seconds through turbulent dispersion.
Hydrodynamics in a Horizontal Stirred Tank Reactor 27

Under laminar conditions, the islands (as a poorly mixed zone is usually
referred to under laminar conditions) in reactor-15 and reactor-11 (Figure 2.5e) are
located below the injection point. These islands are unstable and consequently
disappear within seconds (compare e.g., panels e and f of Figure 2.5). This
observation leads to the conclusion that reactor-11 and reactor-15 are globally
chaotic. Leong and Ottino (1989) indicate that the existence of multiple folds along
the island boundary is indicative for the instability of islands. The presence of a
‘rough’ island boundary in Figure 2.5e can be regarded as an indication for this.
In reactor-20 (Figure 2.7e and f) four islands are visible throughout the reactor.
These islands are very stable in the range of impeller frequencies applied. This
leads to the conclusion that reactor-20 is not globally chaotic. The four islands, in
fact, form four segregated torii, fluid elements that have often been observed in
mixing vessels (Dong et al., 1994; Hoogendoorn et al., 1967; Lamberto et al., 1996;
Lamberto et al., 1999; Nomura et al., 1997). These fluid elements act as barriers to
mixing and are therefore highly undesirable in the polycondensation process as their
existence allows early termination, hence leading to an undesired broad MWD.
Probably, the segregated torii can be terminated by changing the mutual angle
between the blades or by periodically changing the impeller speed (Harvey III et al.,
1997; Lamberto et al., 1996; Leong et al., 1989; Unger et al., 1999).

2.3.3 Macro-Mixing Times under Turbulent Conditions


For turbulent conditions, mixing is quantified by determining macro-mixing

1.1
1
0.9
0.8
0.7
G/Gn [-]

0.6
0.5
0.4
0.3 6 12
3.5 24
0.2 18

0.1
0
0 1 2 3 4 5 6 7 8
Time (in s)
Figure 2.10: Normalized gray scale, plotted versus time, in reactor-11 at 6 Hz with a fill ratio
of 40%. The dots represent the raw data, and the line represents the data after FFT filtering.
The frequency plot from the Fourier transformation is depicted in the inset.
28 Chapter 2

times from response curves. These curves are created by plotting gray scales at a
certain position in the PLIF images against the elapsed time. The chosen position is
located as depicted in Figure 2.3 by an ‘x’ and is exactly two pixels from the
sidewall and two pixels from the cylindrical wall. This position is selected to be
representative for the mixing in the whole reactor as it is covered with fluid at every
fill ratio.
An example of a response curve is depicted in Figure 2.10 for the mixing in
reactor-11, 40% filled with water. The dots represent the raw data after
normalization on final gray scale. The maximum in gray scale is often obtained
because of the appearance of the blades and air bubbles in the measurement point as
a consequence of the presence of 60% air in the reactor. Because the stirrer had a
constant speed, the disturbance of the blades could be removed by using fast
Fourier transformation filtering (FFT filtering) from the commercial software
package TablecurveTM by Jandel Scientific. A standard 40% smoothing level is
used to zero 80% of the higher frequency components and all stirrer-related
frequencies, resulting in the line in Figure 2.10. The accompanying frequency plot
is also given in Figure 2.10 and is discussed in the Appendix.

1.6
1.4
1.2
1
G/Gn [-]

0.8
0.6 1.1
0.4 1.5
0.2 2
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
Time [s]
Figure 2.11: Normalized gray scale plotted versus time at 6 Hz in reactor-11,
reactor-15, and reactor-20 at a fill ratio of 100%.
Figure 2.11 shows normalized response curves after FFT filtering at 6 Hz in all
three reactors that are completely filled. The profiles for reactor-11 and reactor-15
exceed unity, meaning that the dye is preferentially transported in the axial
direction, toward the position where the gray scales are recorded. For reactor-20 the
profile gradually rises to unity, without exceeding this limit. From Figure 2.11 it is
concluded that when L/D is increased, it will take longer to mix the dye to the final
concentration.
Macro-mixing times are determined from the response curves after FFT
filtering, like in Figure 2.11. As a representative value, the time that the normalized
concentration differed by less than 10% from the final concentration was chosen. In
Figure 2.12, these mixing times, hereafter referred to as t10, are represented as a
function of the stirrer speed. The mixing times decrease with increasing impeller
Hydrodynamics in a Horizontal Stirred Tank Reactor 29

Reactor-11, x=0.4
8 Reactor-11, x=0.6
Reactor-11, x=1.0
7 Reactor-15, x=1.0
Reactor-20, x=0.4
6
Reactor-20, x=0.6
Reactor-20, x=1.0
5
T10 [s]

0
3 4 5 6 7 8 9 10 11 12

N [Hz]

Figure 2.12: Mixing time versus impeller frequency for the three reactors at three different
fill ratios for turbulent conditions.

speed, whereas the macro-mixing times increase with increasing L/D. This is due to
the fact that the distance between injection point and measuring position is larger
with higher L/D. A fill ratio of 60% results in shorter mixing times than those for
100% or 40%. Mixing at a fill ratio of 60% can be shorter than mixing at 100%,
because the slowest mixed zone near the shaft is absent at 60%. The difference from
the mixing times at 40% can be a result of a lower overall circulation at 40%. From
this it is suggested that, for a good circulation, a minimum amount of fluid is
needed. These results also reveal that the mixing time is not linearly dependent on
the fill ratio.
It is supposed that the macro-mixing time will depend on the process and
reactor variables as follows:

t m = t m ( ρ , µ , N , D, g , geometrical dimensions of the system) (2.1)

Then, using dimensional analysis, the functional relationship can be arranged as

N ⋅ t m = tm ρ ⋅ ⋅ ,
N D2 N 2 ⋅ D
( µ g
, geometrical dimensions as ratios ) (2.2)

which applies to mixing vessels in general. The polycondensation process is


operated in the hollow state in which the Froude-number N2·D/g is irrelevant. This
leaves L/D as the only relevant geometrical dimension. When the obtained mixing
times are correlated with the parameters varied by applying the relevant
30 Chapter 2

dimensionless numbers as given in equation 2.2, the following empirical correlation


holds for D = 0.18m:

1.21

N ⋅ t10 = 16 ⋅ f ( x ) ⋅ Re 0.11
()

L
D
(2.3)

with

f ( x ) = 0.22 + ( x − 0.70) 2 (2.4)

A power series was chosen to describe the dependence of the mixing time on
the fill ratio. The dimensionless mixing time N·t10 increases with the Reynolds
number, although the contribution is less than 15%. The influence is small, as the
mixing is turbulent over the entire range of applied impeller speeds. However, the
positive power indicates that an increase in impeller frequency results in an increase
in dimensionless mixing time. This suggests that mixing is less efficient at high
stirrer speeds as the fluid tends to more solid-body rotation. The power of 1.2 in
L/D indicates that the mixing mechanism is a combination of convection and
dispersion, as 1 would indicate full convective flow and 2 full dispersive flow.
As mentioned in the Introduction, Ando et al. (1971a) studied mixing in
horizontal vessels with baffles in order to prevent the tendency to solid-body
rotation. Their dimensionless mixing time correlated with L/D, indicating a larger
convective contribution due to the baffles. Mixing in the reactor investigated here is
more dispersive because of the absence of baffles, resulting in a dependency of
(L/D)1.2 and, therefore, larger mixing times. Applying baffles can decrease mixing
times under turbulent conditions. For high viscosity levels, however, baffles are not
required, as viscous shear will damp out behind these baffles.

2.3.4 Macro-Mixing Times for Laminar Conditions


Using the same black-box approach as used for turbulent conditions, macro-
mixing times could be determined for laminar conditions. In the range of the
parameters varied and with D = 0.18m, the empirical correlation N·t10 = 60·(L/D)
can be obtained. This result is in good agreement with the correlations provided by
Hoogendoorn and Hartog (1967) and Novak and Rieger (1975) for helical ribbon
impellers in vertical vessels. According to this equation, mixing under laminar
conditions is only 2.5 times slower than under turbulent conditions. However,
establishing the mixing time in this manner is of course troublesome as the
segregated torii in reactor-20 do not disappear within the determined macro-mixing
times.
A second method for quantifying mixing times is setting the macro-mixing
time equal to the life span of the islands in reactor-11 and reactor-15. In the
polycondensation process it is important to minimize this life span as the presence
of islands implies a concentration ratio deviating from unity. This deviation results
in an increased possibility of early termination, thereby leading to an undesired
Hydrodynamics in a Horizontal Stirred Tank Reactor 31

40

35
L/D=1.5
30 1
tm ∝ L/D=1.1
N 2 .6
Mixing time [s]
25

20

15
1
tm ∝
10 N 1.3
5

0
4 5 6 7 8 9 10 11 12
N [Hz]

Figure 2.13: Life span of islands versus N, in reactor-11


and reactor-15 for laminar conditions.
broad MWD. In Figure 2.13, the life spans are plotted against impeller frequency.
From Figure 2.13, it follows that the life span decreases with increasing impeller
speed. It also follows that, at low impeller frequency, the life span in reactor-15 is
larger than in reactor-11, whereas the two are comparable at high frequencies.

2.3.5 Mixing at Intermediate Viscosities


In this study, we use only water and glycerin, both with Newtonian behavior.
For the polycondensation process, one can imagine that viscosity and rheologic
behavior will change continuously, and mixing behavior will go through a wide
range of scenarios. Leong and Ottino (1990) investigated chaotic mixing in
viscoelastic 2-dimensional flows at various viscosities. Upon an increase in
viscosity, islands grow, and chaotic regions shrink. From this observation, it
appears plausible to conclude, that, between our extreme cases (water and glycerin),
nothing dramatic will occur. However, additional experiments by Leong and Ottino
in which the shear rate was increased show that, in viscoelastic fluids, the number
of islands increases, whereas in Newtonian fluids, this number is constant.
Translating this observation to our practical situation indicates that, at higher
impeller speeds (higher shear rate) more islands are formed. These observations
show that studying hydrodynamics in the Drais reactor with viscoelastic fluids is
mandatory for obtaining a complete picture of the mixing process.
32 Chapter 2

2.4 Conclusions
This study provides information about flow patterns, the presence of poorly
mixed zones, and macro-mixing times in three industrial configurations of the Drais
reactor. Under turbulent conditions, the flow pattern shows flow circulation which
is opposite to the pumping action of the impeller blades. The location of poorly
mixed zones is the same in all three reactors. The dimensionless macro-mixing time
N·t10 is correlated with L/D, stirrer frequency N, and fill ratio x. The obtained
empirical correlation shows the following three relationships:

• N·t10 is at a minimum value at fill ratios around 60%.


• N·t10 increases more than linearly with increasing L/D ratio.
• The incorporated Reynolds number has a positive power, indicating that
mixing at high stirrer speeds becomes less efficient.

Under laminar conditions, the flow patterns found indicate that the mixing is
chaotic. Reactor-11 and reactor-15 are globally chaotic, whereas reactor-20 appears
to have elements of order. This behavior shows that mixing in the Drais reactor is
complex and complicates effective scale-up. The location, number, and life span of
the islands, as well as the overall flow pattern in the Drais reactor, change when
L/D is enlarged. Macro-mixing times could not be determined unambiguously, as
the islands do not disappear within the measurement time. However, the macro
mixing time seems to be only 2.5 times larger than under turbulent conditions.
Hydrodynamics in a Horizontal Stirred Tank Reactor 33

Nomenclature
C clearance
m
Di impeller diameter m
D vessel diameter m
G gravitational constant m/s2
h blade height m
L vessel length m
N number of revolutions Hz
PE perimeter m
PE0 perimeter at t = 0 m
Re Reynolds number -
t time s
tm mixing time s
t10 time at which concentration only differs 10% from final concentration s
V reactor volume m3
w blade width m
x fill ratio -

Greek

α angle between impeller blade and shaft °


β angle between impeller blades in the tangential direction °
µ dynamic viscosity kg/(m·s)
ρ liquid density kg/m3
σ Liapunov exponent 1/s

References
Ando, K.; Hara, H.; Endoh, K.; “Flow behavior and power consumption in horizontal stirred
vessels”, Int. Chem. Eng., 1971a, 11, 735.
Ando, K.; Hara, H.; Endoh, K.; “On mixing time in horizontal stirred vessel”, Kagaku
Kogaku, 1971b, 35, 806.
Ando, K.; Fukuda, T.; Endoh, K.; “On mixing characteristics of horizontal stirred vessel
with baffle plates”, Kagaku Kogaku, 1974, 38, 460.
Ando, K.; Shirahige, M.; Fukuda, T.; Endoh, K.; “Effects of perforated partition plate on
mixing characteristics of horizontal stirred vessel”, AIChE J., 1981, 27(4), 599.
Ando, K.; Obata, E.; Ikeda, K.; Fukuda, T.; “Mixing time of liquid in horizontal stirred
vessels with multiple impellers”, Can. J. Chem. Eng., 1990, 68, 278.
Borkent, G.; Tijssen, P.A.T.; Roos, J.P.; Van Aartsen, J.J.; “Kinetics of the reactions of
aromatic amines and acid chlorides in hexamethylphosphoric triamide”, Recl. Trav.
Chim. Pays-Bas, 1976, 95, 84.
Distelhoff, M.F.W.; Marquis, A.J.; Nouri, J.M.; Whitelaw, J.H.; “Scalar mixing measure-
ments in batch operated stirred tanks”, Can. J. Chem. Eng., 1997, 75, 641.
Dong, L.; Johanson, S.T.; Engh, T.H.; “Flow induced by an impeller in an unbaffled tank”,
Chem. Eng. Sci., 1994, 42, 549.
34 Chapter 2

Fukuda, T.; Idogawa, K.; Ikeda, K.; Ando, K.; Endoh, K.; “Volumetric Gas-phase mass
transfer coefficient in baffled horizontal stirred vessel”, J. Chem. Eng. Jpn., 1990, 13(4),
298.
Gaymans, R.J.; Sikkema, D.J.; “Aliphatic polyamides”, Comprehensive Polymer Science,
Step Polymerization, 1989, 5, 357.
Harvey III, A.D.; Wood, S.P.; Leng, D.E.; “Experimental and computational study of
multiple impeller flows”, Chem. Eng. Sci., 1997, 52(9), 1479.
Holmes, D.B.; Voncken, R.M.; Dekker, J.A.; “Fluid flow in turbine-stirred, baffled tanks - I
-Circulation time”, Chem. Eng. Sci., 1964, 19, 201.
Hoogendoorn, C.J.; den Hartog, A.P.; “Model studies on mixers in the viscous flow region”,
Chem. Eng. Sci., 1967, 22, 1689.
Jeurissen, F.T.H.; Surquin, J.; Private communication.
Kersting, Ch.; Prüss, J.; Warnecke, H.J.; “Residence time distribution of a screw-loop
reactor: experiments and modelling”, Chem. Eng. Sci., 1995, 50, 299.
Lamberto, D.J.; Muzzio, F.J.; Swanson, P.D.; Tonkovich, A.l.; “Using time-dependent RPM
to enhance mixing in stirred vessels”, Chem. Eng. Sci., 1996, 51(5), 733.
Lamberto, D.J.; Alvarez, M.M.; Muzzio, F.J.; “Experimental and computational investiga-
tion of the laminar flow structure in a stirred tank”, Chem. Eng. Sci., 1999, 54, 919.
Leong, C.W.; Ottino, J.M.; “Experiments on Mixing due to Chaotic Advection in a Cavity”,
J. Fluid Mech., 1989, 209, 463-499.
Leong, C.W.; Ottino, J.M.; “Increase in regularity by polymer addition during chaotic
mixing in two dimensional flows”, Phys. Rev. Lett., 1990, 64(8), 874.
Manaresi, P.; Munari, A.; “Factors affecting rate of polymerization”, Comprehensive
Polymer Science, Step Polymerization, 1989, 5, 35.
Mayr, B.; Nagy, E.; Horvat, P.; Moser, A.; “Scale-up on basis of structured mixing models:
A new concept”, Biotechnol. Bioeng., 1994, 43, 195.
Moo-young, M.; Tichar, K.; Takahashi, A.L.; “The blending efficiencies of some impellers
in batch mixing”, AIChE. J., 1972, 18, 178.
Nomura, T.; Uchida, T.; Takahashi, K.; “Enhancement of mixing by unsteady agitation of an
impeller in an agitated vessel”, J. Chem. Eng. Jpn., 1997, 30(5), 875.
Novák, V.; Rieger, F.; “Homogenization efficiency of helical ribbon and anchor agitators”,
Chem. Eng. J., 1975, 9, 63.
Ottino, J.M.; “Unity and diversity in mixing: Stretching, diffusion, breakup and aggregation
in chaotic flows”, Physics of Fluids A3, 1991, 3(5), 1417.
Perona, J.J.; Hylton, T.D.; Youngblood, E.L.; Cummins, R.L.; “Jet mixing of liquids in long
horizontal cylindrical tanks”, Ind. Eng. Chem. Res., 1998, 37, 1478.
Schoenmakers, J.H.A.; Wijers, J.G.; Thoenes, D.; “Determination of feed stream mixing
rates in agitated vessels”, Proc. 8th European Mix. Conf. (Paris): Récents Progrès en
Génie des Procédés, ed. Lavoisier, 1997, 11(52), 185.
Thoenes, D.; “Chemical Reactor Development”, Kluwer Academic Publishers: Dordrecht,
1994.
Unger, D.R.; Muzzio, F.J.; “Laser-induced fluorescence technique for the quantification of
mixing in impinging jets”, AIChE J., 1999, 45(12), 2477.
Vollbracht, L.; Comprehensive Polymer Science, Step Polymerization, 1989, 5, 374.
Hydrodynamics in a Horizontal Stirred Tank Reactor 35

Appendix
In this appendix some remarks are made on the frequency plot in Figure 2.10.
The plot shows spikes at frequencies that are characteristic for raw data like the
impeller speed, i.e., 6 Hz and accompanying higher frequencies such as 12, 18 and
24 Hz. The peak around 3.5 Hz originates from differences in dye concentration in
the tangential direction and is responsible for the large fluctuations in the response
curve. In the response curve of reactor-11 in Figure 2.11, the fluctuations reappear
with a time period of 1/3.5 s, the so-called circulation time tc (Holmes et al., 1964).
As 10 periods can be distinguished, one can estimate a mixing time in the tangential
direction of 3 s. Because the macro-mixing time t10 for this experiment was found to
be 4.0 s, distributive mixing in the tangential direction is faster than in the axial and
radial direction.
The mixing in the tangential direction was also faster than that in the axial and
radial direction in reactor-15 and reactor-20 because no large fluctuations were
observed in the accompanying response curves in Figure 2.11. In these reactors, the
dye is homogeneously distributed in the tangential direction before it has reached
the position where gray scales are read. It is suggested that the clockwise circulation
in the larger reactors, as depicted in Figure 2.8, enhances the tangential mixing.
According to the frequency plot in Figure 2.10, the concentration in the
tangential direction in reactor-11 fluctuates with a specific frequency. However, this
frequency is not constant
throughout the reactor. Figure Injection
2.14 shows the ratio of the 0.18

frequency and impeller


frequency in one-fourth of 0.15
reactor-11. The Figure shows
that, near the wall, the fluid is
Radius [m]

0.12
retained more than in the bulk
as the ratio is smaller. The
0.09
largest ratio is 0.9, positioned
between the two impeller
blades. The fluid in that area 0.06

rotates almost as a solid-body


and will therefore not be well- 0.03
-0.18 -0.16 -0.14 -0.12 -0.10 -0.08 -0.06 -0.04 -0.02 -0.00
mixed, as is confirmed by the
Axial distance [m]
presence of the poorly mixed
Figure 2.14: Ratio of fluid velocity to stirrer
zone in Figure 2.4e at the
speed in reactor-11.
same position.
3
HYDRODYNAMICS AND SCALE-UP OF
HORIZONTAL STIRRED REACTORS

Abstract

In this chapter, the hydrodynamics in horizontal stirred-tank


reactors are investigated. The flow state, agitation power, and macro-
mixing times have been determined experimentally. Two flow states,
i.e., 'slosh' and 'ring', can be distinguished, with transition between the
two states that shows hysteresis. The agitation power was determined
by measuring the temperature increase upon mixing. The power
number appears to be comparable to power numbers in unbaffled
vertical vessels. A variation in fill ratio indicates that agitation energy
dissipates uniformly throughout the reactor under laminar conditions.
Under turbulent conditions, however, most energy is dissipated at the
vessel wall. Using pulse-response measurements, macro-mixing times
have been determined. The mixing times correlate with momentum
input and liquid volume, thus indicating different hydrodynamics at
large and small scales. A combination of mixing times and agitation
power shows that, at small scale and intermediate fill ratios, the
mixing is most energy-efficient.

This chapter is a slightly modified version of the publication:


Van der Gulik, G.J.S.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodynamics and Scale-Up of
Horizontal Stirred Reactors”, Ind. Eng. Chem. Res., 2001, 40(22), 4731-4740.
38 Chapter 3

3.1 Introduction
Horizontal stirred-tank reactors are widely used in industry. A commercial
example is the unbaffled Drais reactor that can be used for multiple purposes such
as powder mixing in catalyst preparation, liquid mixing and kneading during CMC-
production, and polycondensation processes (Vollbracht, 1989; Bannenberg-
Wiggers et al., 1998). In all applications, the reactor is only partially filled. A
schematic representation of the Drais reactor is given in Figure 2.1. Typical for the
unbaffled cylindrical reactor is its horizontal position and the heavily designed
impeller. The impeller makes possible the application of high mixing power, which
is needed to achieve sufficient mixing in viscous fluid processes. The reactor is
characterized by its length L, diameter D, and clearance c, which is the distance
between the blades and the reactor wall. Because the clearance is small, the blades
perform a scraping action that keeps the reactor wall free from sticking material.
The clearance also provides a region with high shear rates and good heat exchange
with the cooled walls. The blades have a pumping action towards the reactor center,
providing an easy way to discharge the reactor through the open hatch as
represented in Figure 2.1.
Despite the wide application of Drais reactors in liquid mixing, little is known
about the hydrodynamics, mixing performance, and scale-up of such reactors. A
literature survey shows that the literature on hydrodynamics in horizontal vessels is
rather limited, in contrast with that on vertical stirred vessels. Some literature exists
on the hydrodynamics under turbulent conditions. Ganz (1957) describes power
measurements in horizontal stirred gas absorbers. Ando et al. (1971a) measured
power input upon stirring in partially filled horizontal vessels in relation to flow
behavior. In the vessel, two flow states could be distinguished. The 'slosh'-state is
obtained at low stirring speed. The liquid is then pushed upward by the impeller and
sprayed, which is ideal for use in gas absorption processes (Ganz, 1957; Ando et al.,
1971b; 1974; 1981; Fukuda et al., 1990). The 'ring'-state, which is obtained at high
stirrer speeds, results in a cylindrical liquid layer on the inside wall. Ando et al.
(1990) also studied turbulent mixing in a horizontal vessel with baffles and multiple
impellers. Macro-mixing times were measured, and a model was proposed for
scale-up purposes. It has been established that the dimensionless macro-mixing time
Ntm is proportional to L/D.
Because literature on mixing in horizontal vessels under laminar conditions is
virtually absent, we have to rely on mixing studies in vertical unbaffled tanks, for
which many studies are available. The usual purpose of these studies has been to
find geometries that provide good mixing performance. Data on mixing vary from
author to author as a result of differences in geometry, definitions, experimental
techniques, and fluid properties. However, a general consensus emerges concerning
impeller designs. Judging from power input and mixing time experiments (Ando et
al., 1990; Hoogendoorn et al., 1967; Novák et al., 1967), it can be concluded that
the flow pattern of a good laminar mixer should include 1) axial flow, 2) all
streamlines passing through the impeller region, 3) no closed streamlines occurring
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 39

outside the impeller region, and 4) frequent disruption of fluid along the wall.
Figure 2.1 shows that conditions 2 and 4 will probably be achieved in the Drais
reactor, because the impeller blades pass through the entire reactor volume. In
Chapter 2, the flow patterns in a Drais reactor have been described, as investigated
using planar laser induced fluorescence. It was found that axial flow in the reactor is
rather effective. Also, closed streamlines can be present, acting as toroidal vortices.
We have studied the macro-mixing in a Drais reactor filled with low- and
high-viscosity liquids with the aim of obtaining a better understanding of the
mixing during a polycondensation process in which the viscosity strongly increases
as a result of the formation of polyaramid molecules (Bannenberg-Wiggers et al.,
1998). To obtain a polymer product with the required quality in terms of MW
(molecular weight) and MWD (molecular weight distribution), it is important to
expose the polymerizing liquid to high shear rates (Agarwal et al., 1992). The high
shear rates increase reaction rates through molecular orientation and rotational
diffusion of the rods. This has been shown in an experimental study by Agarwal and
Khakhar (1992; 1993), using two reactors in series in which the first reactor is a
vertical reactor with a high-speed stirrer ensuring good overall mixing. The second
reactor provides Couette-flow hydrodynamics, with nearly homogeneous high shear
flow over the entire reactor but with little overall mixing. The authors were able to
improve product quality considerably by increasing the shear rate in the second
reactor. The Drais reactor investigated here, combines the important features of
both reactors (Vollbracht, 1989; Bannenberg-Wiggers et al., 1998): a large impeller
provides overall mixing while at the same time high shear rates occur in the small
clearance between impeller and vessel wall.
As the clearance of the Drais reactor is small, its volume is small compared to
the total liquid volume: in completely filled reactors, this volume ratio is 2⋅10-4.
Therefore, macro-mixing in the reactor has to be optimized so that all liquid in the
bulk will pass the high-shear region in the clearance frequently. A second reason
that emphasizes the importance of short macro-mixing times is that the
polycondensation is performed in a semi-batch manner: one reactant is fed to the
other and has to be mixed quickly throughout the reactor to prevent the occurrence
of a premature termination reaction (Bannenberg-Wiggers et al., 1998; Chapter 2).
Judging from the available literature, it is unclear what the macro-mixing time will
be and how it will evolve in scale-up. Therefore, we have conducted an
experimental study at different scales in which we established macro-mixing times
by means of pulse-response measurements. The applied agitation power was also
measured to link the mixing performance with power consumption. All
measurements were performed in the 'ring'-state as the polycondensation process is
performed at high stirrer frequencies, thus forcing the liquid into the 'ring'-state. As
it is known that application of the 'ring'-state is required for a high MW to be
obtained (Bannenberg-Wiggers et al., 1998), we determined the impeller frequency
at which the 'ring'-state forms or disappears during operation. Also, flow behavior
was studied as a function of vessel and stirrer geometry, impeller frequency, and
fluid viscosity.
40 Chapter 3

3.2 Scale-up theory


For scale-up of the polycondensation process, it is important to know how the
mixing time evolves upon reaction. From the literature it is well-known that the
mixing time will be a function of the process conditions and reactor configuration:

tm = ƒ(ρ,µ,N,DI,g,geometrical dimensions of the system) [s] (3.1)

Using dimensional analysis and omitting the Froude number (Fr), the functional
relationship can be rearranged to:

Ntm = ƒ(Re, geometrical dimensions as ratios) [-] (3.2)

Ntm is the dimensionless mixing time and is expected to be independent of the


Reynolds number (Re) under both turbulent and laminar conditions (Harnby et al.,
1992). Under intermediate conditions, Ntm will be a power law in Re. In our
specific case, Fr can be omitted as Fr is only important under conditions at which
transition occurs between 'slosh'- and 'ring'-state. We are only interested in macro-
mixing times in the 'ring'-state as the polycondensation process is performed in this
state.
To maintain tm constant during scale-up, a constant impeller frequency N is
required according to equation 3.2. This will also result in a constant average shear
rate γ& a , which is related to N by

dv
γ& a = − = k1 N [1/s] (3.3)
dy

with k1 close to unity (Metzner et al., 1957; Thoenes, 1994). The highest shear rate,
γ&max occurs in the clearance and can be estimated using

dv πND − 0 πNDi
γ& max = − ≈ = ∝ k2 N [1/s] (3.4)
dy c
c D / 2 − Di / 2

Thus, γ&max is constant when N is kept constant, provided that the clearance is kept
in a constant ratio with the vessel diameter. Keeping N constant in scale-up,
however, is strongly reflected in the power requirement. Under turbulent conditions,
the required power is given by

P = ρN P N 3 Di5 [W] (3.5)


Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 41

and the average energy dissipation rate per unit of mass is

P N P f g N 3 Di2
ε= = [m2/s3] (3.6)
ρVl x

where ρ is the liquid density [kg/m3], NP is the power number [-], N is the agitation
rate [1/s], Di is the impeller diameter [m], x is the fill ratio [-], Vl is the liquid
volume [m3], and fg is the geometrical factor, which is 1.05 [-].
For geometrically similar systems, the power number NP can be rewritten as a
functional relationship of dimensionless groups. Under turbulent conditions, only
Re is relevant for the Drais reactor; thus

N p = k 3 Re a [-] (3.7)

Under laminar conditions, the following relation applies

P = k 4 µN 2 Di3 [J/s] (3.8)

and consequently

P
ε= = k 5νN 2 [J/(kg⋅ s)] (3.9)
ρVl

with k4 and k5 as constants and ν the kinematic viscosity. For these equations, it is
assumed that Np depends on Re-1 only, which is a good approximation when Re <
100.
From equations 3.5 and 3.8, it can be seen that, for constant N, P increases
with Di5 and Di3, respectively, which are both highly impracticable. Therefore, on
larger scales, lower specific power input has to be applied which is usually obtained
by reducing the impeller speed. This results in an increase in tm, which is
undesirable in the polycondensation process. With this experimental study, we
identify the limitations that can be faced during scale-up.
42 Chapter 3

3.3 Experimental Section

3.3.1 The Drais reactor


To investigate the mixing process at different scales, four scale models of the
Drais reactor were available. Typical geometrical data are given in Table 3.1. The
small models are named reactor11, reactor15 and reactor20, where the numbers 11,
15, and 20 refer to the L/D-ratio. The large-scale model is reactor11-60, which
refers to the L/D-ratio and the diameter. Reactor11 and reactor11-60 are
geometrically similar. Reactor15 and reactor20 differ from reactor11 in angle β
between blades (as shown in Figure 2.1) and in length. A detailed description has
been given in Chapter 2.

Table 3.1: Geometric data regarding the four reactors.


Small-scale Large-scale
Parameter Reactor11 Reactor15 Reactor20 Reactor11-60
Reactor length L [m] 0.198 0.27 0.36 0.66
Reactor diameter D [m] 0.18 0.6
Reactor volume Vr [L] 5.0 6.7 8.9 185
Impeller diameter Di [-] D⋅(29/30) D⋅(29/30)
Blade width w [-] D/2 D/2
Blade height h [-] D/12 D/12
Blade thickness [-] D/90 D/54
Shaft diameter [-] D/6 D/10
β (°) 180 135 120 180
Materials
Shaft Stainless steel Stainless steel
Cylindrical wall Glass Perspex
Side walls Stainless steel Perspex

3.3.2 Transition in flow state


The flow state was determined for different reactor geometries and fluid
viscosities by observing the impeller frequency at which a fluid ring was formed or
collapsed. The formation was complete when the inner gas/liquid surface was flat.
Both transitions were clearly visible.

3.3.3 Power measurements


Power measurements were performed in reactor11, reactor20, and reactor11-
60. Depending on the desired value of Re, tap water, glycerin (purity > 99.9%
Heybroek, Amsterdam), or a mixture of the two was used as the working fluid. The
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 43

Figure 3.1: Schematic representation of the experimental


set-up for the temperature measurements.

power input to the liquid by stirring was measured by determining the temperature
increase of the liquid with time. The experimental set-up is depicted in Figure 3.1.
Two Pt100-elements, denoted T1 and T2, measured the temperature in the reactors
during agitation. There was no difference observed between T1 and T2 under
turbulent conditions. Under laminar conditions, the difference never exceeded 0.2
°C. Pt100-element T3 measured the ambient temperature, Ta, during the
experiments. All temperatures were measured with an accuracy of 0.1 °C. The
power input followed from the energy balance over the mixing vessel, as given in
equation 3.10.

dT
ρC pVl = P − h(T − Ta ) [J/s] (3.10)
dt

where ρ is the liquid density [kg/m3]; Cp is the heat capacity [J/(kg⋅K)]; Vl is the
liquid volume [m3]; T is the liquid temperature [K]; Ta is the ambient temperature
[K]; P is the dissipated stirring energy [J/s]; and h is overall heat transfer coefficient
[J/(K⋅s)].
The term on the left-hand side in equation 3.10 represents the accumulation
term, P represents the dissipated stirring energy, and the last term represents the
lumped losses to the environment. Assuming density, heat capacity, overall heat
transfer coefficient and ambient temperature to be constant and temperature-
independent, differential equation 3.10 can be solved by considering the initial
condition T(0) = T0, resulting in

 − ht  P  
T (t ) = Ta + (T0 − Ta ) exp  + 1 − exp − ht  [K] (3.11)
 ρC V  h  ρC V 
 p l    p l 
44 Chapter 3

3.3.4 Pulse-response experiments


The experimental set-up for the pulse-response experiments is depicted in
Figure 3.2. It was decided to perform measurements outside the vessel, because the
clearance was too small for a probe to be placed inside. Therefore, the reactor
contents were circulated through a spectrophotometer (A) placed in an external
loop. The liquid was withdrawn from an outlet, placed at the same height as the
clearance and fed back at the top in the reactor center. The flow rate in the loop was
monitored by a flow meter (F). The measurement started when a small amount of a
concentrated aqueous methylene blue solution was injected as a pulse, δ(t), in the
reactor center. The concentration was measured by the spectrophotometer, which
was connected to a PC for automatic monitoring.

δ
A
F

PC
A

Figure 3.2: Schematic representation of the experimental set-up


for the pulse-response measurements.

For reactor11-60, the volume of the piping was 450 mL, and the flow rate was
50 mL/s. Because the liquid volume was between 72 and 180 L, higher order
mixing effects can be ignored. The time needed for the tracer to leave the reactor
and reach the spectrophotometer was 0.9 s. This delay time was measured by
injecting a small amount of tracer at the reactor outlet. Macro-mixing times were
corrected for this delay. For the smaller reactors, the volume of the piping was 250
mL, the flow rate was 25 mL/s, the total liquid volume varied between 2 and 10 L,
and the delay was 2.1 s. The total amount of injected solution was 3 mL. Using a
high-speed camera, as described in chapter 2, the injection time was determined to
be 0.20 ± 0.03 s. This time was negligible compared to the mixing time.
An example of a tracer concentration response curve with time is given in
Figure 3.3. The plotted response on the left y-axis was normalized using:

C (t ) − C (0)
C n (t ) = [-] (3.12)
C∞ − C (0)
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 45

1.2 10

σ [-]
1 1

2
0.8 0.1

Cn [-] 0.6 0.01

0.4 0.001

0.2 0.0001
tm = 12.3
0 0.00001
0 5 10 15 20 25
Time [s]

Figure 3.3: Typical response curve with the corresponding variance


for D = 0.6 m, L = 0.66 m, x = 100 %, and N = 4.8 Hz.

On the right y-axis of Figure 3.3 is plotted the variance of concentration, σ2, around
the equilibrium value, as defined by

σ 2 (t ) = (1 − C n (t )) 2 [-] (3.13)

The macro-mixing time was defined to be the time at which the variance was below
10-3. This resulted in tm = 12.3 s for the experiment in Figure 3.3. Every presented
macro-mixing time (shown in Figures 3.11-14) is the average of at least three
measurements.
Table 3.2 provides the impeller frequencies used in the pulse-response
experiments. Under turbulent conditions, the chosen frequency ensured that the
liquid was in the 'ring'-state. Consequently, the values of Fr at large and small scale
were similar. The values of Re were not similar, which was less important as it
followed from literature that Ntm was independent of Re at the high applied values
of Re (Coulson and Richardson, 1996). Using glycerin in reactor11-60 provides Re
values up to 1730, which does not really justify the laminar classification. However,
for convenience we have grouped all measurements using glycerin.

Table 3.2: Range of impeller frequency and the corresponding shear rate γ& , tip speed, Re and
Fr values as applied in the pulse-response measurements.
Kinematic Dynamic
Reactor id. D [m] Regime N [Hz] γ& [1/s]
Vtip[m/s] Re [-] Fr [-]
Laminar 3.0-11.2 560-2100 1.7-6.3 100-360 0.17-2.3
Reactor11, 5
0.18 Turbu- 1.77⋅10 -
-15, and -20 5.5-9.1 1040-1715 3.1-5.1 5 0.55-1.5
lent 2.95⋅10
Laminar 0.6-4.8 110-900 1.1-9.0 210-1730 0.02-1.4
Reactor11-60 0.6 5
Turbu- 6.48⋅10 -
1.8-4.8 340-900 3.4-9.0 5 0.20-1.4
lent 1.73⋅10
46 Chapter 3

3.4 Results and discussion

3.4.1 Flow state in water


For the reactor filled with water, the Fr values at which fluid rings form or
collapse, are plotted in Figure 3.4 against the fill ratio for reactor11, reactor20, and
reactor11-60. The solid symbols mark the Fr value at which the fluid ring forms
with increasing impeller speed, and the open symbols mark the Fr value at which
the ring collapses with decreasing impeller speed. From Figure 3.4, it follows that,
at all fill ratios hysteresis occurs between the values of Fr for ring formation and
collapse. The values of Fr at which the transition occurs coincide for the ring
collapse, however, they do not coincide for ring formation. At low as well as at high
fill ratios, flow changes occur at higher values of Fr than is the case at intermediate
fill ratios. This implies that, at intermediate fill ratios, weaker inertial forces, i.e.,
less power, are required to form a fluid ring.
From Figure 3.4, the value of Fr can be determined at which a fluid ring is
present. For the large-scale reactor, ring formation occurs around Fr = 1, which can
practically be used as a criterion for ring formation. This information is important,
as the fluid in the pulse-response measurements and power measurements has to be
in the 'ring'-state for the polycondensation process (Bannenberg-Wiggers et al.,
1998). Because of the hysteresis effect, however, it is possible to maintain the 'ring'-
state at stirrer frequencies lower than the frequency needed for ring formation.

Ring formation: Reactor11-60 Reactor11 Reactor20


Ring collapse: Reactor11-60 Reactor11 Reactor20
2.5

1.5
Fr [-]

0.5

0
0 0.2 0.4 0.6 0.8 1
Fill ratio [-]

Figure 3.4: Flow state transition, expressed in Fr as function of the fill ratio
for reactor11-60, reactor11, and reactor20.
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 47

3.4.2 Flow state in glycerin


In reactors partially filled with glycerin, no fluid ring is formed at high stirrer
speeds. For illustration, an image of the flow state is recorded with a high speed
camera and is presented in Figure 3.5. Viscous forces are too high to obtain a fluid
ring as in water. The gas phase is finely dispersed in the liquid, providing a milky
fluid with small gas bubbles. Figure 3.5 also shows the presence of large holes in
the fluid in the wake behind the blades. The holes do not completely extent to the
cylindrical wall, indicating that the wall is entirely covered by a liquid film. The
presence of the liquid film was proven to exist as small gas bubbles were found,
using a stroboscope, that are transported tangentially by the impeller with a lower
speed than the impeller speed. The presence of the liquid film probably is
important, as this is the region with the highest shear rate.

Figure 3.5: Photograph of glycerin in reactor20 at a fill ratio of 40 % at 8 Hz.

3.4.3 Power measurements


In Figure 3.6, the temperature rise is plotted as a function of time for five
different situations. The time for measurement ranged from 5 minutes in reactor11
with glycerin to over 2 hours in reactor11-60 with water. By fitting the temperature
profiles with equation 3.11, the applied agitation power P was obtained. The overall
heat transfer coefficient h also follows from the fitting procedure. However, as the
interpretation of these results appeared not to be very straightforward, these data
have not been included.
In Figure 3.7, the power number NP, as defined by P/ρN3D5, is given as a
function of Re. The solid line represents measurements in a completely filled
reactor11 and reactor11-60, which are geometrically similar but differ only in size.
The measurements are in close agreement with data for a propeller mixer in an
unbaffled vertical vessel as reported by Rushton et al. (1950). NP becomes
independent of Re at high values of Re, i.e., under turbulent conditions.
Accordingly, the power 'a' of Re in equation 3.7 approaches 0. The applied power is
plotted against Re in Figure 3.8. From the slope of the lines at high values of Re, it
follows that the applied power is proportional to the cubic root in impeller speed,
which is in accordance with equation 3.5. Under laminar conditions, NP becomes
inversely proportional to Re (the exponent 'a' in equation 3.7 approaches -1). From
Figure 3.8, it follows that
48 Chapter 3
6.0 N = 12 Hz
x = 40 %
P = 154 W
5.0 Glycerin, small scale model
N = 8 Hz
4.0 x = 80 %
P = 87.4 W
T-T0 [°C]

N = 12 Hz
Water, small scale model
3.0 x = 100 %
P = 48.6 W
N = 8 Hz Water, large scale model
2.0
x = 100 %
P = 17.5 W
N = 2.4 Hz
1.0
x = 100 %
P = 202 W
0.0
0 1000 2000 3000 4000 5000 6000 7000 8000

Time [s]

Figure 3.6: Temperature rise versus elapsed time for the small- (D = 0.18 m) and large-scale
(D = 0.6 m) models, filled with glycerin or water.

Np(0.4,1.1) Np(0.6,1.1) Np(0.8,1.1) Np(1.0,1.1)


Np(0.4,2.0) Np(0.6,2.0) Np(0.8,2.0) Np(1.0,2.0)

Np [-]

0.1

100 1000 10000 100000 1000000


Re [-]
Figure 3.7: Power number as a function of Re with the fill ratio and L/D ratio as parameters.
1000
D=0.18
P(0.001)
P(0.003)

100 P(0.014)

3.0
P(0.099)
P~N
P [W]

P(0.297)
2.6 P(0.527)
P~N
10 P(1.4)
2.1 D=0.60
P~N
P(0.001)

1
10 100 1000 10000 100000 1E+06 1E+07
Re [-]

Figure 3.8: Power as a function of Re with viscosity and diameter as parameters.


Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 49

10

Water
Np (0.4)
Np (0.6)
1
Np (0.8)
Np [-]

Np (1.0)
Glycerin
0.1 Np (0.4)
Np (0.6)
Np (0.8)
Np (1.0)
0.01
0 0.5 1 1.5 2 2.5 3 3.5 4
Fr [-]
Figure 3.9: Power number as a function of Fr with fill ratio and fluid type as parameters.

the applied power at low values of Re is proportional to the square root of the
impeller speed, which is in accordance with equation 3.8.
The power number NP is given as a function of Fr in Figure 3.9. Figure 3.9
shows that NP is virtually independent of Fr, apart from the case in which the
reactor is completely filled with glycerin. For a horizontal reactor, Ando et al.
(1971a) also found that NP is independent of Fr.
In Figures 3.10A-D, the power P and dissipated energy per unit mass ε are
given as functions of fill ratio for water and glycerin. With glycerin, the applied
power was around 5 times higher than it was with water. From Figures 3.10A and
3.10B, it follows that, for water and glycerin, the applied power increases with
increasing fill ratio. For water, the increase is less than proportional, as follows
from Figure 3.10C in which ε decreases with increasing fill ratio. In water, most
energy is dissipated at the impeller tips and at the cylindrical wall. At low fill ratios,
this region represents a relatively larger volume than at high fill ratios. Therefore,
ε will be higher at low fill ratios.
Figure 3.10D shows that, in glycerin, ε is independent of the fill ratio,
suggesting that energy is dissipated more uniformly throughout the fluid than in
water. This also follows from Figure 3.10F, in which ε is plotted against Re. The
values for ε at all fill ratios coincide and follow a power law of 2.1. As ε is
homogeneous, the shear rate is also homogeneous. Consequently, the contribution
of the high shear rate in the clearance to ε is limited, which can be explained by the
small clearance volume.
50 Chapter 3

80 600
A) Water N L/D
B) Glycerin N L/D
70 8 , 1.1 8 , 1.1
500
10 , 1.1 10 , 1.1
60
12 , 1.1 400 12 , 1.1
50 14 , 1.1

P [W]
14 , 1.1
P [W]

40 8 , 2.0 300 8 , 2.0


12 , 2.0
30
200
20
100
10

0 0
0.4 0.5 0.6 0.7 0.8 0.9 1 0.4 0.5 0.6 0.7 0.8 0.9 1

25 90
C) Water 80
20 70
60
D) Glycerin
15
ε [m /s ]

50
2 3
ε [m /s ]
2 3

40
10
30

5 20
10
0 0
0.4 0.5 0.6 0.7 0.8 0.9 1 0.4 0.5 0.6 0.7 0.8 0.9 1
Fill ratio [-] Fill ratio [-]

25 90
x L/D
F) Glycerin
E) Water 0.4 , 1.1
80
20 70
0.6 , 1.1 x L/D
0.8 , 1.1 60 0.4 , 1.1
15 1.0 , 1.1
ε [m /s ]
ε [m /s ]

50 0.6 , 1.1
2 3
3

0.4 , 2.0
2

40 0.8 , 1.1
10 0.6 , 2.0 1.0 , 1.1
0.8 , 2.0 30
1.0 , 2.0 20
5
10
0 0
100000 200000 300000 400000 500000 0 100 200 300 400 500
Re [-] Re [-]

Figure 3.10: Results of the power measurements: (A) power in water against fill ratio,(B)
power in glycerin against fill ratio, (C) ε in water against fill ratio, (D) ε in glycerin against
fill ratio, (E) ε in water against Re, (F) ε in glycerin against Re. Parts (A)-(D) have N and
L/D as parameters, and Parts (E) and (F) have fill ratio and L/D as parameters.

Using multivariable analysis, the relationship in equation 3.14 between NP and


the varied parameters can be obtained for laminar conditions as

0.89
x 0.92  L 
N P = 93 0.85   [-] (3.14)
Re  D 

and for turbulent conditions as


Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 51

0.44
L
N P = 0.15 x 0.36
  [-] (3.15)
D

The standard errors of the exponents are given in Table 3.3. For laminar
conditions, NP proves to be nearly linear with liquid volume as the exponent for the
fill ratio comes close to unity. Exponent 'a' approaches –1, which is in agreement
with equation 3.7.
Summarizing, it follows that, the value of NP in this case is similar to that of a
propeller in an unbaffled tank. In water, ε is a function of the fill ratio and is
highest near the vessel wall. In glycerin, ε is independent of the fill ratio and is
homogeneous over the reactor.

Table 3.3: Standard errors of the constants in the correlations for NP.
Conditions Pre-exp. factor Exponent in x Exponent in (L/D) Exponent in Re
Lam. (Re < 400) 93 ± 9.6 0.92 ± 0.066 0.89 ± 0.064 -0.85 ± 0.02
5
Turb. (Re > 10 ) 0.15 ± 0.0055 0.36 ± 0.064 0.44 ± 0.072 -

3.4.4 Macro-mixing times in water under turbulent conditions


The dimensionless mixing time Ntm in the large reactor11-60 is depicted in
Figure 3.11A as a function of the fill ratio. It shows that, at every fill ratio, Ntm is
independent of Re. Also, Ntm increases with increasing fill ratio, indicating that
more impeller revolutions are required for a given degree of mixing with increasing
liquid volume.
From Figure 3.11B-D, in which Ntm is given for reactor11, reactor15, and
reactor20, respectively, it follows that the lowest Ntm is found at a fill ratio of 0.7,
which is in agreement with the results in chapter 2. This suggests that at low fill
ratios (x < 0.4), the total amount of fluid is too low to provide good overall
circulation in the fluid ring, whereas at high fill ratios (x ≈ 1), the poorly mixed
zone near the shaft reduces the advantages of the better overall circulation.
A comparison of Ntm for reactor11, reactor15, and reactor20 shows that Ntm
increases with increasing L. Also, Ntm shows to be independent of Re, although Ntm
increases slightly with increasing Re in reactor20. This increase indicates that the
fluid tends toward solid-body rotation at higher stirrer speeds. Solid-body rotation
can occur more easily in reactor20 than in reactor11 and reactor15, because of the
relatively small effect of the sidewalls in reactor20.
The above observations show that reactor11 and reactor11-60 differ
significantly in hydrodynamic behavior despite their geometric similarity. In
reactor11-60, the shortest mixing time is found at a fill ratio of 0.4, which is the
lowest fill ratio applied, whereas in reactor11, it is found at a fill ratio of 0.7. The
Ntm values at the two scales are also different. Apparently, turbulent and convective
52 Chapter 3

vtip [m/s] vtip [m/s]


2.6 5.2 7.8 10.4 2.6 3.0 3.4 3.8 4.2 4.6 5.0
90 90
A) Reactor11-60 B) Reactor11 x [-]
80 80
1
70 70
0.7
Ntm [-]

60 60 0.6
50 50 0.5
40 No stable 40 0.4
30 fluid ring 30
20 20
10 10
0 0
500000 1000000 1500000 2000000 150000 200000 250000 300000

2.6 3.0 3.4 3.8 4.2 4.6 5.0 2.6 3.0 3.4 3.8 4.2 4.6 5.0
90 90
C) Reactor15 D) Reactor20
80 80
70 70
60 60
50 50
Ntm [-]

40 40
30 30
20 20
10 10
0 0
150000 200000 250000 300000 150000 200000 250000 300000
Re [-] Re [-]

Figure 3.11: Dimensionless macro-mixing time against Re and impeller speed using water for
(A) reactor11-60, (B) reactor11, (C) reactor15, and (D) reactor20.

mixing on both scales are different. An important difference between reactor11 and
reactor11-60 is the wall-surface/liquid-volume ratio. This ratio is higher at small
scale in reactor11 (22.2 m2/m3) than at large scale in reactor11-60 (6.66 m2/m3). As
a result, the fluid tends more toward solid-body rotation at large scale and,
therefore, shows increased mixing times.
The observed trends indicate that the mixing time depends on scale and liquid
volume in contrast to the conclusion of Harnby et al. (1992), who state that, for
geometrically similar systems, Ntm is constant. Fox and Gex (1956), Middleton
(1979), and Mersmann et al. (1976) have observed that, in vertical vessels, mixing
times depend on liquid volume. Fox and Gex (1956) have correlated mixing times
with liquid volume and momentum input according to:

Vl0.5
tm ∝ [s] (3.16)
(N 2
D4 )
0.42

Application of this approach has the advantage that the fill ratio can easily be
implemented. When we use all presented mixing data, the following relationship is
obtained:
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 53

tm = 66 ⋅
(x π D L)
4
2 0.68

[s] (3.17)
(N D ) 2 4 0.42

Obviously, this is in good agreement with equation 3.16. This empirical


relationship can be used for scale-up of the Drais reactor for processes under
turbulent conditions (Re > 105).

3.4.5 Macro-mixing times in glycerin under laminar conditions


When the reactor is filled with glycerin at fill ratios below 1, the gas phase
is finely dispersed in the liquid. In the milky liquid thus obtained, as shown in the
photograph in Figure 3.6, it is impossible to perform spectrophotometric pulse-
response measurements with the current set-up. Therefore, experiments have only
been performed using completely filled reactors.
In Figure 3.12A, Ntm is plotted against Re. It follows that the mixing time
in glycerin is approximately 2.5-5 times longer than in water for the small and large
reactors. Ntm is hardly dependent on Re on the different scales. Thus, in accordance
with the observation of Harnby et al. (1992), Ntm is independent of Re for both
laminar and turbulent conditions. However, this is only valid for a given scale, as
the Ntm values for reactor11 and reactor11-60 differ despite their similar
geometries. This emphasizes our previous observation that the hydrodynamics in
the two geometrically similar systems are significantly different.
Fox and Gex (1956) also provided a correlation for mixing times under
laminar conditions:

Vl 0.5
tm ∝ [s] (3.18)
(N 2
D4 )
1.25

500
400

300

100
200
Nt m [-]

Reactor11-60
Reactor11-60
tm [s]

Reactor20
100 Reactor20
90 Reactor15
80 Reactor15
70
Reactor11
A 10 Reactor11 B
60
50
100 1000 0.1 1 10
εε [m
2 3
Re [-] /s ]

Figure 3.12: (A) Dimensionless macro-mixing time against Re and impeller speed, using
glycerin. (B) tm as a function of ε using glycerin.
54 Chapter 3

Using the same approach, we obtain

tm = 530 ×
( D L) p
4
2 0.98

[s] (3.19)
(N D ) 2 4 0.44

The standard errors are given in Table 3.4. The differences between equations 3.18
and 3.19 can be a result of different geometries. Also, the results of Fox and Gex
(1956) might be biased, as their experiments were performed in one single vessel in
which only the liquid height was varied.
Table 3.4: Standard errors of the constants in the correlations for t m
Conditions Pre-exp. factor Exponent in p/4 D2L Exponent. in N2 D4
Laminar (Re < 1730) 519 ± 36.4 0.98 ± 0.038 -0.44 ± 0.014
Turbulent (Re > 105) 66 ± 3.1 0.68 ± 0.022 -0.45 ± 0.022

3.4.6 Macro-mixing efficiencies


The term Ntm is an efficiency parameter that represents the number of
revolutions required to obtain the desired mixing at time tm. According to this term,
the highest mixing efficiency is found in reactor11 at a fill ratio of 0.7. Less than 10
impeller revolutions are required to obtain the required mixing which is 5 times
more efficient than in reactor11-60 or reactor20.

35 14
A) Reactor11-60 B) Reactor11
30 12
25 10
x [-]
20 8 1
15 6 0.7
10 4 0.6
0.5
tm [s]

5 2
0.4
0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
14 14
C) Reactor15 D) Reactor20
12 12
10 10
tm [s]

8 8
6 6
4 4
2 2
0 0
0 2 4 6 8 10 12 0 2 4 6 8 10 12
2 3 2 3
e [m /s ] e [m /s ]

Figure 3.13: tm as a function of e for (A) reactor11-60, (B)


reactor11, (C) reactor15, and (D) reactor20.
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 55

In Figure 3.12B and 3.13A-D, the mixing times are plotted against the average
energy dissipation ε for laminar and turbulent conditions, respectively. ε is
calculated from equation 3.6, from which NP is calculated using the exponents in
Table 3.3. From these figures, it can be concluded that, even though ε is the same,
on average, tm in reactor11-60 is 5 times longer than in reactor11, although the
length and diameter are only 3 times greater. In reactor11 at a fill ratio of 0.7, 1
m2/s3 is required to obtain mixing times of 1 s, whereas in reactor20, the same ε
value provides a mixing time of 9 s. Therefore, an important conclusion is that
scale-up of the Drais reactor based on energy consumption only, will lead to longer
mixing times.

3.5 Concluding remarks


In this study, we have examined the hydrodynamics in the Drais reactor at
different scales. It has been concluded that the specific power consumption as a
function of Reynolds number is the same on small and large scales. Over a wide
range of Re values, the power number Np is comparable to Np of vertical unbaffled
reactors with propellers. However, it has been shown that scale-up with constant
power consumption will lead to longer mixing times at larger scale. This is reflected
in Ntm, which depends linearly on the fill ratio at large scale and with the square
root at small scale (Chapter 2). Thus, an important conclusion is that the
hydrodynamics at different scale are different, making scale-up of the Drais reactor
a difficult task. For scale-up purposes, however, the macro-mixing times have been
correlated, using an approach as suggested by Fox and Gex (1956) that includes
liquid volume and flux of momentum. Especially for turbulent conditions, the
agreement in the obtained correlation is remarkable.
Two flow states occur in this reactor, i.e., a 'slosh'-state at low vales of Fr and
a 'ring'-state at high values of Fr. The transition between the two states as a result of
changing impeller speed shows hysteresis. Using the criterion Fr=1 in scale up, one
can determine the number of revolutions required to obtain the 'ring'-state. The
'ring'-state only applies for low-viscosity liquids. Using high-viscosity liquids, we
observe a liquid phase with a dispersed gas phase. Nevertheless, still a liquid film
exists on the cylindrical wall. The existence of this thin liquid film is important,
because the highest shear rates are obtained in this film, which is crucial for
obtaining products with high molecular weights in polymerization reactions
(Agarwal et al., 1992).
Mixing times under laminar conditions are approximately only 2.5 times
higher than under turbulent conditions. In a previous paper, we showed that mixing
under laminar conditions is very efficient, because of the chaotic nature. Striations
were made visible using planar laser-induced fluorescence and shown to stretch,
fold, and reorient throughout the whole reactor. The deformation process of the
striations appears to occur homogeneously, which corresponds to the observed
homogeneous energy dissipation as described in this paper. Because of this
homogeneous dissipation, it can be concluded that the Drais reactor is a very
efficient mixer, especially under laminar conditions.
56 Chapter 3

Nomenclature
a constant -
b constant -
c clearance m
C arbitrary concentration -
Cn normalized concentration -
C∞ final concentration -
Cp heat capacity J/(kg⋅K)
D vessel diameter m
Di impeller diameter m
fg geometrical factor -
Fr Froude number -
g acceleration of gravity m/s
h overall heat transfer coefficient J/(K⋅s)
h blade height m
k1-k5 constants -
L length of the vessel m
NP power number -
N agitation rate 1/s
P dissipated stirring energy J/s
Re Reynolds number -
t time s
tm macro-mixing time s
T liquid temperature K
Ta ambient temperature K
T0 temperature at t = 0 K
Vl liquid volume m3
Vr reactor volume m3
vtip impeller speed m/s
w blade width m
x fill ratio -
y distance m

Greek

α blade angle °,degrees


β mutual blade angle °,degrees
γ& shear rate 1/s
ε average power input per unit of mass m2/m3
µ dynamic liquid viscosity kg/(m⋅s)
ν kinematic viscosity m2/s
ρ liquid density kg/m3
σ2 variance -
Hydrodynamics and Scale-Up of Horizontal Stirred Reactors 57

References
Agarwal, U.S.; Khakhar, D.V.; “Enhancement of polymerization rates for rigid rod-like
molecules by shearing”, Nature, 1992, 360, 53.
Agarwal, U.S.; Khakhar, D.V.; “Shear flow induced orientation development during
homogeneous solution polymerization of rigid rodlike molecules”, Macromolecules,
1993, 26, 3960.
Ando, K.; Hara, H.; Endoh, K.; “Flow behavior and power consumption in horizontal stirred
vessels”, Int. J. Chem. Eng., 1971a, 11, 735.
Ando, K.; Hara, H.; Endoh, K.; “On mixing time in horizontal stirred vessel”, Kagaku
Kogaku, 1971b, 35, 806.
Ando, K.; Fukuda, T.; Endoh, K.; “On mixing characteristics of horizontal stirred vessel
with baffle plates”, Kagaku Kogaku, 1974, 38, 460.
Ando, K.; Shirahige, M.; Fukuda, T.; Endoh, K.; “Effects of perforated partition plate on
mixing characteristics of horizontal stirred vessel”, AIChE J., 1981, 27(4), 599.
Ando, K.; Obata, E.; Ikeda, K.; Fukuda, T.; “Mixing time of liquid in horizontal stirred
vessels with multiple impellers”, Can. J. Chem. Eng., 1990, 68, 278.
Bannenberg-Wiggers, A.E.M.; Van Omme, J.A.; Surquin, J.M.; “Process for the batchwise
preparation of poly-p-terephtalamide”, U.S. Pat., 5,726,275, 1998.
Coulson, J.M.; Richardson, J.F.; Backhurst, J.R.; Harker, J.H.; “Coulson and Richardson’s
Chemical Engineering”, Vol. 1, 5th ed., Fluid flow, heat transfer and mass transfer;
Butterworth-Heinemann Ltd: Oxford, England, 1996.
Fox, E.A.; Gex, V.E.; “Single-phase blending of liquids”, AIChE J., 1956, 2(4), 539.
Fukuda, T.; Idogawa, K.; Ikeda, K.; Ando, K.; Endoh, K.; “Volumetric Gas-phase mass
transfer coefficient in baffled horizontal stirred vessel”, J. Chem. Eng. Jpn., 1990, 13(4),
298.
Ganz, S.N.; Zh. Prikl.Khin., 1957, 30, 1311.
Harnby, N.; Edwards, M.F.; Nienow, A.W. (Eds); “Mixing in the Process Industries”, 2nd
ed., Butterworth-Heinemann Ltd: London, England, 1992.
Hoogendoorn, C.J.; den Hartog, A. P.; “Model studies on mixers in the viscous flow region”,
Chem. Eng. Sci., 1967, 22, 1689.
Mersmann, A.; Einenkel, W.D.; Kappel, M.; “Design and scale up of mixing equipment”,
Int. Chem. Eng., 1976, 16, 590.
Metzner, A.B., Otto, R.E.; “Agitation of Non-Newtonian Fluids”, AIChE J., 1957, 3(1), 3.
Middleton, J.C; “Measurement of circulation within large mixing vessels”, Proc. 3rd Eur.
Conf. On Mixing, University of York, BRHA Fluid Eng. Cranfield, England, 1979, A2,
15.
Novák, V.; Rieger, F.; “Homogenization efficiency of helical ribbon and anchor agitators”,
Chem. Eng. J., 1975, 9, 63.
Rushton, J.H.; Costisch, E.W.; Everett, H.J.; “Power characteristics of mixing impellers,
Parts I and II”, Chem. Eng. Prog., 1950, 46, 395 & 467.
Tatterson, G.B.; “Fluid mixing and gas dispersion in agitated tanks”, McGraw-Hill Inc: New
York, United States of America, 1991.
Thoenes, D.; “Chemical Reactor Development”, Kluwer Academic Publishers: Dordrecht,
The Netherlands, 1994.
Vollbracht, L.; “Aromatic Polyamides”, Compr. Polym. Sci., Step Polym., 1989, 5, 374.
FLUID FLOW AND MIXING IN
4
AN UNBAFFLED HORIZONTAL
STIRRED TANK

Abstract

The turbulent flow field in a horizontal stirred tank reactor has


been examined using Laser Doppler Anemometry and Computational
Fluid Dynamics. The LDA experiments show that in the unbaffled
reactor the mean velocities in tangential direction are at least one
order of magnitude higher than in axial and radial direction. In these
directions the turbulent and periodic fluctuations dominate over the
mean velocities. Therefore, macro mixing in these directions is
determined by turbulent dispersion.
In the CFD calculations the isotropic k-ε model and the
anisotropic Differential Stress Model have been applied to incorporate
the turbulent properties of the flow. The mean flow properties are
reasonably well predicted with both models. Both models
underestimate the fluctuating properties. Scalar mixing experiments
are simulated properly only when the anisotropic Differential Stress
Model is used. This implies that when the choice for a turbulence
model (isotropic k-ε model or anisotropic Differential Stress Model),
is based on flow properties only, mixing simulations with passive and
reacting scalars might fail completely for those cases where
anisotropic turbulence is expected.

This chapter is a slightly modified version of:


Van der Gulik, G.J.S.; Wijers, J.G.; Keurentjes, J.T.F.; “Fluid Flow and Mixing in an
Unbaffled Horizontal Stirred Tank”, Submitted for publication in AIChE. J.
60 Chapter 4

4.1 Introduction
Horizontal stirred tank reactors are widely used in industry. A commercial
example is the unbaffled horizontal reactor of the Drais type (Turbulent
Schnellmischer, Drais Ltd, Mannheim, Germany) as depicted in Figure 2.1. This
reactor can be used for multiple purposes like powder mixing in catalyst
preparation, liquid mixing and kneading during CMC production or
polycondensation processes (Vollbracht, 1989; Bannenberg-Wiggers et al., 1998).
In all applications the reactor is only partially filled (40% < x < 75%). The reactor
is characterized by length L, diameter D and clearance B, which is the distance
between the blades and the reactor wall. Since the clearance is small, the blades
perform a scraping action that keeps the reactor wall free from sticking material.
The small clearance also implies the presence of a region with high shear rates and
good heat exchange. The blades have a pumping action towards the reactor center
plane at xL = 0, providing an easy way to discharge the content of the reactor
through the opened hatch as depicted in Figure 2.1.
We have previously studied the mixing in the horizontal reactor (Van der
Gulik et al., 2001a and 2001b) within the framework of the application in
polycondensation processes in which viscosity increases several orders of
magnitude during reaction. The choice for the Drais reactor has been made based on
the good mixing performance when the reactor content is highly viscous. However,
at the start of the polycondensation process, when the components are added and
mixed, the reactor content has a low viscosity, resulting in turbulent conditions.
Consequently, the mixing performance has also to be sufficient at turbulent
conditions as this determines the local monomer ratio. We have characterized the
mixing in the horizontal reactor at both low and high viscous conditions by
determining power consumption (Van der Gulik et al., 2001a), mixing times (Van
der Gulik et al., 2001a and 2001b) and the flow pattern (Van der Gulik et al.,
2001b). Because the reactor is unbaffled and the impeller speed is usually high, it
has been found that the fluid mainly flows in tangential direction. The flow in axial
and radial direction can be interpreted as secondary flow superpositioned on the
tangential flow. The secondary flow pattern has been studied by using Planar Laser
Induced Fluorescence from which it follows that flow in axial direction dominates
over the flow in radial direction (Van der Gulik et al., 2001b). Pulse response
measurements have confirmed this observation (Van der Gulik et al., 2001a). For a
quantitative description of the internal flow field computational fluid dynamics
(CFD) calculations are becoming increasingly popular. CFD requires models that
describe the turbulent properties of the flow. The k-ε model, as proposed by
Launder and Spalding (1974), is based on an eddy viscosity hypothesis. One of the
main assumptions is that the turbulence is isotropic. Although turbulence is in
general anisotropic in mixing vessels (Kresta, 1998), the k-ε model is often used
because of simplicity and computational convenience (Schoenmakers, 1998;
Montante et al., 2001; Brucato et al., 2000; Togatorop et al., 1994; Read et al.,
1997; Rousseaux et al., 2001; Brucato et al., 1998). Reynolds Stress Models (RSM)
Fluid flow and mixing in an unbaffled horizontal stirred tank 61

are computationally much more demanding (5 additional equations compared to the


k-ε model) but are able to incorporate the anisotropic nature of turbulence. CFD
studies on unbaffled vessels that compare the viability of both eddy viscosity
models and RSM show different results. For example, Armenante et al. (1997) and
Ciofalo et al. (1996) could only reproduce their Laser Doppler Anenometry data
(LDA data) properly using RSM. Montante et al. (2001) used the k-ε model and a
RSM to study the effect of the impeller clearance on the flow pattern in a vertical
vessel. The k-ε model sufficed for all cases, except for the smallest clearance.
In this paper we have examined CFD for describing the hydrodynamics in the
Drais reactor. As a start we have checked the viability of the k-ε model and the
Differential Stress Model (DSM), which is a RSM in differential form (Launder et
al., 1975). For validation purposes, we have performed LDA measurements
providing local velocities and turbulent quantities. Insight into these parameters
should allow for effective scaling up and should allow definition of routes for
improvement on the current reactor. Using CFD we have also performed scalar
mixing simulations in the obtained flow fields for examining the mixing
performance. The results of these simulations have been compared with previously
reported PLIF data (Planar Laser Induced Fluorescence, Van der Gulik et al.,
2001b) providing an additional validation on the turbulence modeling.

0.0 xL 0.17 0.33


0.05

0.17 = w⋅sin(30) +
Ur
-

- Ua +
Figure 4.1: Quarter of the reactor in which the axial coordinate xL has been
defined: xL = 0.0 in the reactor center and xL = 0.33 at the side wall. The
direction of the axial velocity Ūa and the radial velocity Ūr have been
exemplified.
62 Chapter 4

4.2 Experimental and numerical set-up

4.2.1 The horizontal vessel


The vessel as depicted in Figure 2.1 was used in this study. Detailed
geometrical data were given previously (Van der Gulik et al., 2001b). The diameter
and the length were equal to 0.60 and 0.66 m, respectively, providing an L/D-ratio
of 1.1. The clearance B was equal to 0.01 m. The impeller blades make an angle of
30° with the shaft. More detailed dimensions of the blades are given in Figure 4.1.
The total blade length ‘w’, as defined in Figure 2.1, was 0.3 m. In Figure 4.1, the
length, projected on the x-y-plane, was equal to 0.17 m. The blade height ‘h’ had a
maximum of 0.05 m in the middle of the blade and was equal to 0 at the far ends of
the blade.

4.2.2 LDA measurements


LDA is an optical method for fluid flow research, based on a combination of
interference and Doppler effects. LDA allows the measurement of the local,
instantaneous velocities of particles suspended in the flow. LDA has a high
resolution power in time and is non-invasive. A detailed description of the method
has been given by Durst et al. (1981). Theoretical aspects of the turbulent flow
occurring in the Drais reactor and how the turbulent flow has been investigated
using LDA, are given in Appendix 4A.
The LDA system used here has been described by Schoenmakers (1998) and
Schoenmakers et al. (1997). A 2W Argon laser (Spectra Physics, Model Stabilite
2017-055) was used to provide two colored beams with wavelengths of 514.5 and
488 nm, respectively. Using a lens with a focal length of 400 mm both laser beams
converged in a elliptical control volume of approximately 0.15 × 0.15 × 3.3 mm. An
automated traverse system with the laser probe and the receiving optics was utilized
in Backscatter acquisition mode. Two photo multipliers and two Burst Spectrum
Analysers (Dantec, Denmark) were used for measuring two velocity components
simultaneously.
The LDA measurements were performed in a perspex model of the Drais
reactor that was placed in a square perspex container. The impeller frequency was
set to 3 Hz. All measurements were performed using water. The square perspex
container was also filled with water to minimize lensing effects. Positioning of the
measurement volume was realized using the procedures of Kehoe and Prateen
(1987). Nevertheless, it was not possible to measure the tangential and radial
components close to the cylindrical wall and the side wall. The measurement points
are summarized in Table 1. To obtain the three velocity components, two separate
measurements were taken for the same (r, Z) point: one in the plane at 0°, yielding
the tangential and axial components, and one in the plane at -90°, yielding the axial
and radial components. In axial direction steps were made of 0.01 m. Only half the
reactor was examined since flow proved to be symmetrical over an axial plane at x
Fluid flow and mixing in an unbaffled horizontal stirred tank 63

= 0. In each location 30,000 measurements were performed with an average data


rate of 0.3 kHz. For seeding we used an aqueous dispersion of polystyrene particles
with an average diameter of 4 µm.

Table 4.1: Radial locations for performed measurements between axial positions 0 < xL< 0.29.
r r/R Tangential (0°) Axial (0° and –90°) Radial (-90°)
0.295 0.983 √
0.29 0.967 √
0.285 0.95 √ √
0.28 0.933 √ √
0.27 0.9 √ √ √
0.25 0.833 √ √ √
0.22 0.733 √ √ √
0.29 0.633 √ √ √
0.16 0.533 √ √ √
0.13 0.433 √ √ √

The experimental error in mean velocity, periodic and turbulent fluctuations


was less than 10%. The mean and fluctuating axial velocities in the 0° and -90°
plane appeared to be the same within 10%, indicating that the average flow
characteristics are rotation symmetric. Therefore, all experimental data will be
presented as if measured in a single plane as presented in Figure 4.1. The direction
of the axial and radial velocities has also been exemplified in Figure 4.1. A negative
axial velocity is directed towards the reactor center whereas a positive velocity is
directed towards the reactor side wall. A negative radial velocity is directed towards
the shaft and a positive radial velocity is directed towards the cylindrical wall.

4.2.3 Computational Fluid Dynamics


For the CFD calculations the Finite Volume Package CFX-4.2 was used (AEA
Technology), installed on a Silicon Graphics Origin 200 workstation. The grid used
is depicted in Figure 4.2. The grid was built up in Cartesian coordinates and
comprised of 402 blocks and 374,820 cells. Because the angle between the impeller
blades and the shaft is 30°, the grid strongly deviates from a computationally ideal
orthogonal grid. The Block-Stone solver was required to solve the differential
equations.
64 Chapter 4

Figure 4.2: The grid as used in the CFD Calculations, in front and side view.

The SIMPLE algorithm was adopted to couple the continuity and Navier–Stokes
equations. Two turbulence models were used: the DSM and the standard k-ε model.
Mathematical details are given in Appendix 4B. Because of the distorted grid, it
was necessary to adopt the fully deferred correction for the ε equation in both
turbulence models. On walls the conventional linear logarithmic “wall functions”
were used (Launder and Spalding, 1974). The y+-values ranged from an average
value of approximately of 300 on both the cylindrical wall and the impeller blade to
over 1000 at the shaft. The latter region was thought to be less important. The 1st-
order accurate Hybrid differencing scheme was used for all advection terms.
Additional calculations were performed using the 2nd-order CCCT-scheme
(Curvature Compensated Convective Transport) for all equations. This scheme is a
modification of the 3rd-order Quick-scheme in that it is bounded.
Simulations were only carried out at the impeller speed at which the LDA
measurements were carried out (i.e., 3 Hz). The rotating action of the impeller was
implemented using the “rotating coordinates” approach. In this approach the
coordinates of the complete grid rotated with 3 Hz while the vessel wall rotated
back with 3 Hz. Consequently, the net velocity of the vessel wall equals 0 m/s. The
Coriolis forces are implemented automatically in the “rotating coordinates”
approach.
The simulations were conducted as a transient event. The simulations with the
k-ε model were started from still fluid conditions. Then 80 steps with a time step
corresponding to 90° impeller rotation were applied using 25 iterations per time
step. This corresponds to 20 complete impeller rotations of 6.66 seconds real time at
3 Hz. These large time steps were applied to obtain fluid motion in the tank. The
attained solution was subsequently refined by allowing 12 rotations of the impeller
with time steps covering an angular extent of 360/56 = 6.43°. The refined
simulation thus consisted of 672 time steps and 25 iterations per time step.
Fluid flow and mixing in an unbaffled horizontal stirred tank 65

For the simulations with DSM, the final solution obtained with the k-ε model
was used as an initial guess. Subsequently, the geometry was allowed to rotate with
time steps covering an angular extent of again 6.43° with a total of 672 time steps,
i.e., 4 seconds of real time. Here also 25 iterations were used per time step.
It was found that using the above-described time-marching procedure, a
pseudo steady state was obtained in the vessel. This implies that at the end of the
transient calculation the solution only differed from the solution in the previous
time step in that the complete field was rotated with the required angular extent,
indicating that sufficient convergence was obtained by the procedure adopted.

4.2.4 Scalar Mixing


In the obtained velocity field, the mixing of an inert tracer was studied. In a
cell next to the inlet (which is depicted in Figure 2.1) an initial concentration equal
to 1 was defined. The progress of mixing was monitored at the position that is
marked in Figure 2.3 with an ‘×’ (xL = 0.32, r = 0.295). The tracer was mixed in the
previously mentioned pseudo steady state. Time steps in the tracer mixing
simulations again covered an angular extent of 6.43°.
66 Chapter 4

4.3 Results and discussion LDA

4.3.1 Mean flow characteristics


The time average tangential velocity Ūt proves to be almost independent of the
axial position. Therefore, all the measurements performed at one radial coordinate
have been averaged. These averages are presented as a function of the radius r in
Figure 4.3 with a solid line. The location of the shaft, impeller arm, impeller blade
and clearance are depicted on the x-axis. The impeller speed Uim as a function of the
radius r is also depicted with a dotted line. The tip speed is equal to Uim(r = 0.29) =
2⋅π⋅r⋅N = 2⋅π⋅0.29⋅3 = 5.47 m/s.
According to Figure 4.3, there appears to be a bulk region (0 < r < 0.22) where
the tangential velocity is almost equal to the local impeller speed. The flow
behavior in this bulk region can be qualified as ‘solid body rotation’. A second
region can be distinguished near the cylindrical wall (0.22 < r < 0.3) where the
tangential velocity is fairly constant (Ūt ≈ 3.8 m/s). In this region viscous forces
reduce the tangential velocity. As the velocity of the cylindrical wall is equal to zero
there will be a steep fall in tangential velocity very close to the wall. This could not
be quantified as no measurements could be performed sufficiently close to the wall.

1 5.47
Uim(0.29) = 5.47 [m/s]
0.9 4.92

0.8 4.37

0.7 3.83
Ūt (r)/Uim(0.29) [-]

0.6 3.28
Ūt(r) [m/s]

0.5 2.73

0.4 2.19

0.3 1.64
Ūim(r)
0.2 1.09
Clearance .
0.1 0.55
Shaft Blade arm Blade .
0 0.00
0 0.05 0.1 0.15 0.2 0.25 0.3
r [m]
Figure 4.3: Tangential velocity Ūt as a function of the radial coordinate. On the left y-
axis Ūt has been made dimensionless through division by the impeller tip speed. On the
right y-axis the absolute velocity is given. On the x-axis, the location of the shaft, the
blade arm, the blade and the clearance have been exemplified.
Fluid flow and mixing in an unbaffled horizontal stirred tank 67

0.2 r = R = 0.3 m
0.3 0.30
0.15
0.275 0.28
Clearance
0.1 r [m] 0.25 0.25

0.05 0.29 0.225 0.23


Impeller blade
0.28 0.2 0.20
0 0.27 0.175 0.18
Ua [m/s]

r [m]
-0.05 0.25 0.15 Impeller arm 0.15

| 0.22 0.125 0.13


-0.1
0.19 0.1 0.10
-0.15
A B
0.075 0.08

-0.2 0.05 0.05


Shaft
0.025 0.03
-0.25
0 0.00
0 0.05 0.1 0.15 0.2 0.25 0.3 -0.08 -0.06 -0.04 -0.02 _0 0.02 0.04 0.06 0.08 0.1
xL [m] Ua [m/s]

Figure 4.4: A) Time averaged axial velocity Ūa at several radii r, as a function of the
axial coordinate xL. B) Averages of the velocities as presented in Figure 4.4A.

The time averaged axial velocities Ūa are depicted in Figure 4.4A as a function
of the axial coordinate. The maximum absolute velocity is 0.2 m/s at (xL,r) =
(0.1,0.29). At this position Ūt is equal to 3.8 m/s. Based on these measurements it
can be concluded that the axial velocities are much smaller than the local tangential
velocities. The axial velocity appears to be a function of the radius. There are
regions with positive velocities (e.g., at r = 0.22) and region with negative velocities
(e.g., at r = 0.28). As these regions cannot be determined easily from Figure 4.4A,
the xL-averaged axial velocities are given as a function of the radius in Figure 4.4B.
Near the cylindrical wall and near the impeller blade (0.3 < r < 0.27) the velocities
are negative, meaning that these velocities are directed towards the reactor center.
For 0.26 < r < 0.2 the axial velocities are a positive, thus directed towards the side
wall.
In Figure 4.5, the radial velocities are given. On average the velocity is
negative, which violates with continuity. However, their order of magnitude is
similar to the axial velocities. The highest velocities are measured near the side wall
at high xL-values.
0.05

-0.05 r [m]
0.27
Ur [m/s]

-0.1 0.25
0.22
| -0.15 0.19
0.16
-0.2

-0.25

-0.3
0 0.05 0.1 0.15 0.2 0.25 0.3

xL [m]
Figure 4.5: Time averaged radial velocity Ūr at several radii r,
as a function of the axial coordinate xL.
68 Chapter 4

0.35 0.35
Average turbulent fluctuations Average periodic fluctuations
0.3 0.3

0.25 0.25
uT [m/s]

uP [m/s]
0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05
A B
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
r [m] r [m]

0.4 0.4
Turbulent fluctuations on r = 0.27 m Periodic fluctuations on r = 0.27 m
0.35 0.35
0.3 0.3
0.25 0.25
uT [m/s]

uP [m/s]

0.2 0.2
0.15 0.15
Axial
0.1 0.1
Tangential
0.05 Radial 0.05
0
C 0
D
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
xL [m] xL [m]

Figure 4.6: A) Time averaged turbulent fluctuation on r = 0.27 for axial, radial and
tangential direction. B) Time averaged periodic fluctuation on r = 0.27 for axial, radial and
tangential direction. C) xL- averages of the fluctuations presented in Figure 4.6A, but for all
radii. D) xL-averages of the fluctuations presented in Figure 4.6B, but for all radii.

4.3.2 Turbulent quantities


In Figure 4.6A, the time averaged turbulent fluctuations are given at a radial
coordinate r = 0.27 m. The turbulent fluctuations for the three directions are of the
same size, indicating that the turbulence shows isotropic behavior. The fluctuations
in the vicinity of the side wall are higher than in the reactor center. Similar isotropic
behavior in unbaffled vessels has also been observed by Montante et al. (2001).
The periodic fluctuations at a radial coordinate r = 0.27 are presented in Figure
4.6B and show to be of the same order of magnitude as the turbulent fluctuations.
Anisotropic behavior is observed near the side wall where the tangential
fluctuations show a strong increase. In this region, the side wall scraper probably is
responsible for the stronger periodic fluctuations.
In Figure 4.6C and 4.6D, the xL-averaged values for the turbulent and periodic
fluctuations, respectively, are plotted against the radial coordinate. The turbulent
fluctuating velocities in Figure 4.6C indicate isotropy as the fluctuations in all three
directions are of the same order of magnitude over the complete radius. The
periodic fluctuations in Figure 4.6D show differences for the three directions,
indicating the presence of an anisotropic turbulent mesostructure.
Fluid flow and mixing in an unbaffled horizontal stirred tank 69

0.3

0.25

0.2
r [m]

0.15

0.1
ε [m2/s3]
0.05

0
0 0.05 0.1 0.15 0.2 0.25 0.3
xL [m]
Figure 4.7: Distribution of the energy dissipation rate ε for the measurable region.

In Figure 4.7, the energy dissipation rate ε is given in a 2-dimensional plot.


This area represents 80% of the total reactor volume. Only the turbulent fluctuating
part is included as the periodic fluctuations are supposed to represent pseudo
turbulence. In the measured region, ε ranges from 2.5 m2/s3 near the cylindrical wall
to 0.01 near the shaft. The maximum of 2.5 m2/s3 is located in the region where the
impeller tips overlap. Integration of ε and multiplication by 2 to take the other half
of the reactor into account, provides a dissipated energy of 45 W. On average, ε
equals 0.31 m2/s3. Previously reported results (Van der Gulik et al., 2001a) provided
a value of 375 W for the complete reactor so that almost all energy dissipates very
close to the wall.

4.3.3 Overall remarks


The detailed LDA information provides guidelines that can be useful in the use
and scale-up of Drais reactors. For example, Figure 4.6B shows that the periodic
fluctuations in the reactor center are lower than near the side wall. Consequently,
mixing will be more efficient near the side wall. From this observation one can
expect that the L/D-ratio should not be made too large, as the fluctuations in the
vicinity of the reactor center lower upon an increase in L/D, thus decreasing the
mixing performance. This is in agreement with results presented in a previous paper
in which we have shown that mixing times increase more than linearly upon an
increase in L/D (Van der Gulik et al., 2001a and 2001b).
In Table 2 the average velocity and average turbulent components are given
for r = 0.27 which are representative for the reactor. Note that the turbulent and
periodic fluctuations in axial and radial direction are much higher than the averaged
70 Chapter 4

velocities. In axial direction, the difference is one order of magnitude. Therefore,


the flow in axial and radial direction is exclusively determined by turbulent and
periodic fluctuations. Consequently, macro mixing in these directions will be
determined by turbulent dispersion. In tangential direction the convective flow is
determining.
The reactor can be considered as a moderate turbulent mixer as the liquid is
mainly rotated in tangential direction and overall mixing is slow. Moreover, at radii
below r = 0.22 the liquid rotates almost as a solid body. Figures 4.6C and 4.6D
show that the fluctuations reduce at decreasing radii, meaning that the mixing
performance reduces. Therefore, the reactor is usually used at fill ratios below 50%,
so that at high impeller speeds a ring of fluid exists above r = 0.2 where at least
some mixing occurs. This limitation in fill ratio up to 50% can be seen as a
disadvantage of the Drais reactor.

Table 4.2: RMS values in m/s of velocity and fluctuations scales at r = 0.27.
r = 0.27 Tangential Axial Radial
Mean velocity 3.73 0.035 0.14
Turbulent fluctuations 0.32 0.32 0.31
Periodic fluctuations 0.2 0.18 0.16
Fluid flow and mixing in an unbaffled horizontal stirred tank 71

4.4 Results and Discussion CFD

4.4.1 Mean flow characteristics


The calculated tangential velocities Ūt only depend on the radius which has
also been observed in the LDA experiments. Therefore, all tangential velocities are
xL-averaged analogous to the LDA velocities as represented in Figure 4.8. Overall,
DSM combined with the Hybrid differencing scheme provides the best results as
the data are the closest to the LDA data. Especially around the impeller blade at r =
0.27, the comparison is relatively good. The k-ε model with the Hybrid differencing
scheme underestimates the tangential velocity around r = 0.27 whereas the DSM
with the CCCT-differencing scheme underestimates the tangential velocity. The
dissipative nature of the Hybrid differencing scheme seems to be effective in
predicting the measured velocity. In the LDA data the highest tangential velocity is
measured at r = 0.22m. All CFD solutions underestimate the velocity at this radius.
As will be shown, axial velocities at this radius are also underestimated.

Uim(0.29) = 5.47 m/s


1 5.47

0.9 4.92

0.8 4.37
LDA-data
0.7 k-e/hybrid 3.83
dsm/hybrid
Ūt (r)/Uim(0.29) [-]

Ūt (r) [m/s]

0.6 3.28
dsm/ccct
0.5 2.73

0.4 2.19

0.3 1.64

0.2 1.09
Uim(r)
0.1 0.55
Tangential velocities .
0 0.00
0 0.05 0.1 0.15 0.2 0.25 0.3
r [m]

Figure 4.8: xL-averaged tangential velocity Ūt as a function of the radial coordinate


for the LDA measurements and the CFD calculations.

In Figure 4.9 simulated mean axial and radial velocities at r = 0.27 are
compared with LDA data. The axial velocities in Figures 4.9A are all of the same
order of magnitude and follow the same trend. In detail the velocities differ but
regarding the quality of the numerical grid and the large periodic fluctuations that
have been observed in the LDA measurements, the comparison is as good as
expected. The simulated radial velocities in Figure 4.9B differ substantially from
72 Chapter 4

0.15 0.1

0.10 0

0.05 -0.1
Ūa [m/s]

Ūr [m/s]
0.00 -0.2
-0.05 -0.3
lda
-0.10 -0.4 r = 0.27 dsm
k-e
-0.15 -0.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
xL [m] xL [m]

Figure 4.9: A) Comparison of the axial velocity Ūa, and B) the radial velocity Ūr. Both
time averaged measured with LDA and calculated using the k-ε model and the DSM.

the measured velocities. Near the side wall a relatively strong radial flow (0.3 m/s)
towards the shaft has been obtained. The effect of this flow phenomenon will be
discussed in section 4.5. The strong radial velocity could not be confirmed using
LDA data as no measurements could be performed near the side wall.
In Figure 4.10A the xL-averaged axial velocities are compared with LDA
measurements. The k-ε model and the DSM are well matched despite some minor
differences near the shaft. The profile of the simulated velocities has the same
nature as the LDA data, meaning that alternating negative (towards reactor center)
and positive (towards side wall) velocities occur as a function of the radius r. The
origin of the alternating direction can be explained from the vector plot in Figure
4.10B which has been obtained using the DSM with the Hybrid scheme. In this
vector plot, the axial and radial components of the total velocity vector are
0.3

0.275
0.25

0.225

0.2

0.175
r [m]

0.15 DSM/Hybrid
0.125

0.1
A LDA-data
B
0.075 k-e/hybrid
dsm/hybrid Plane with axial and
0.05
dsm/ccct C radial velocity vectors
0.025
Axial velocities
0
-0.08 -0.06 -0.04 -0.02 _0 0.02 0.04 0.06 0.08 0.1
Ua [m/s]

Figure 4.10: A) xL-averaged velocities Ūa as a function


of the radius r. B) Vector plot representing the axial
and the radial components. C) Exemplification of the
point of view in the reactor for obtaining Figure 4.10B.
Fluid flow and mixing in an unbaffled horizontal stirred tank 73

projected on a plane that crosses the shaft and the impeller blade near the side wall.
The point of view is along the impeller blade as depicted in Figure 4.10C. The
impeller blade (0.24 < r < 0.29) transports the fluid towards the reactor center while
the flow reverses between the impeller blade and the shaft. Between impeller tip
and cylindrical wall the flow appears to be directed towards the side wall which is
not observed in the LDA measurements.

4.4.2 Turbulent quantities


The periodic velocity fluctuations are not directly available from CFD data,
but can be calculated using equation 4.1:

∑ [(U ]
N0
− U 1 ) + (U 2 − U 2 ) + (U 3 − U 3 )
2 2 2
kP = 1 1
2 N0 1 [m2/s2] (4.1)
1

in which Ui are instantaneous local velocities and Ūi the time averaged velocities
(Montante et al., 2001). N0 is the number of cells in the tangential direction and kP
represents the energy content of the periodic fluctuations. The results are presented
in Figure 4.11A. The data presented are the averages of all kP in tangential and axial
direction. kP as calculated with both the k-ε model and DSM are of the same size as
the measured kP. However, they differ in trend: the simulated kP remains more or
less constant over the radius, the measured kP decreases with decreasing radius. The
overestimated kP at low radii is probably due to the poor mesh quality near the
shaft. Near the cylindrical wall, the differences in kP are approximately one order of
magnitude.
The turbulent fluctuations in all three directions are represented by the
turbulent kinetic energy kT, which has been calculated using equation 4A.5 in
Appendix 4A. Although the use of kT for DSM masks anisotropic properties, it
makes comparison with kT from the k-ε model straightforward. The presented kT
values in Figure 4.11B are xL- and time averaged. Clearly the k-ε model predicts the
kT values better than DSM. However, both simulated kT are lower than the
measured kT. Similar differences have been observed before and have been
recognized as one of the major of discrepancies between LDA measurements and
CFD calculations (Montante et al., 2001; Brucato et al., 1998; Armenante et al.,
1997; Ng et al., 1998).
Figure 4.11C shows the (xL-averaged and time-averaged) energy dissipation
rate for the LDA measurements and the simulations. The simulated ε is significantly
lower than the measured one. The simulated ε shows a strong increase near the
cylindrical wall. The total dissipated energy using both turbulence models is given
in Table 4.3. The predicted values appear to be about 30% of the measured value,
for both the complete reactor and the part of the reactor in which LDA
measurements could be performed (Van der Gulik et al., 2001a). The ratio between
the values for the complete and for part of the reactor are the same for LDA and
CFD data, meaning the distribution in dissipated energy appears to be similar.
74 Chapter 4

Table 4.3: Dissipated energy [W] in the complete reactor and in a part of the reactor using
different experimental techniques and numerical models.
Experimental CFD
Region Temperature k-ε DSM
LDA
measurements Hybrid Hybrid
Complete reactor 375 (100%) - 101 (100%) 93 (100%)
Part of the reactor
- 45 (12%) 13.9 (14.8%) 7.7 (8.2%)
-0.29<xL<0.29 0<r<0.285

Concluding, based on the main flow characteristics the DSM is preferred


above the k-ε model. According to the turbulent properties, the k-ε model is
preferred although the differences with measured quantities are significant for both
turbulence models.

1 1

A B
0.1
0.1
kP [m2/s2]

kt [m2/s2]

0.01

0.01
0.001

0.001 0.0001
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
r [m] r [m]

10
dsm C
1 k-e
LDA

0.1
ε [m /s ]
2 3

0.01

0.001

0.0001
0 0.05 0.1 0.15 0.2 0.25 0.3
r [m]

Figure 4.11: A) Periodic fluctuations, represented by the periodic kinetic energy kP,
A) Turbulent fluctuations, represented by the turbulent kinetic energy kT, and C) the
energy dissipation rate ε. All are time and xL-averaged.
Fluid flow and mixing in an unbaffled horizontal stirred tank 75

4.5 Scalar mixing


In the flow fields obtained using the k-ε model and the DSM, mixing of an
added scalar has been monitored. This is especially relevant for simulations with
chemical reactions in which the components are represented by reactive scalars.
In Figure 4.12, the scalar concentrations for the DSM and the k-ε model are
given against real time. The concentrations are monitored in the position that is
marked with an ‘x’ in Figure 2.3. Scalar concentrations are normalized to their final
concentration. It can be seen that is takes about 2 seconds for the scalar to reach the
monitoring point. The concentration initially fluctuates strongly for both
simulations because concentration differs in tangential direction. The main
difference between the k-ε model and DSM is that the curve for the k-ε model
gradually rises towards the final value of 1.0 while the curve for the DSM exceeds
the final concentration and then gradually lowers to the final value.
In Figure 4.12 also a curve has been given that results from a pulse-response
experiment with a colored tracer as described before (Van der Gulik et al., 2001a).
The curve does not show the strong fluctuations because the monitoring probe was
placed outside the reactor. The shape of the response curves shows strong
resemblance with the curve for the DSM. Clearly the DSM represents the mixing
experiment closer than the k-ε model does.
In Figures 4.13A and B the calculated scalar concentration distribution is given
for the DSM and the k-ε model, respectively, after a mixing time of 4 seconds. The
red color represents the region with a higher concentration than the final
concentration. Figure 4.13B emphasizes that using the k-ε model, the mixing in
radial direction is faster than in axial direction as the red region extends until the
shaft. This can be explained by the isotropic turbulence making the mixing speed in
all directions equal, but in radial direction apparently the fastest because of the
reduced volume towards the shaft. Using the DSM, the mixing in radial and axial

1.4
Mixing time measurement [3]
1.2
Normalized scalar conc. [-]

1.0
dsm
0.8
k-ε
0.6

0.4

0.2

0.0
0 5 10 15 20 25 30 35 40
Time [s]

Figure 4.12: Passive scalar concentration plotted against time for a pulse response experiment
(Van der Gulik et al., 2001a) and for CFD calculations using the k-ε model and the DSM.
76 Chapter 4

direction apparently proceeds at the same speed. This difference in mixing behavior
also clarifies the shape of the response curves in Figure 4.12. The monitoring
position differs only in axial coordinate from the inlet position, so that using the
DSM the scalar will be traced before it will be traced using the k-ε model.
Previously we have reported on PLIF experiments (Planar Laser Induced
Fluorescence, Van der Gulik et al., 2001b) in a Drais reactor with a diameter of 0.18
m at an impeller speed of 6 Hz. This reactor is geometrically similar to the reactor
currently investigated. The mixing in the complete reactor using PLIF
measurements is presented in Figure 4.13C. The measured mixing pattern in this
figure resembles the mixing pattern in Figure 4.13A as calculated using the DSM.
This resemblance and the resemblance of the response curve shows that for
simulations of scalar mixing in this reactor the DSM performs much better than the
k-ε model. An important conclusion from this work is that the difference in scalar
mixing behavior should make CFD users cautious, as based on mean flow and
fluctuating properties no clear preference could be made between the k-ε model and
DSM.

A B

Figure 4.13: Distribution of passive scalars, modeled using A) the DSM, B) the k-ε model,
and C) determined experimentally using PLIF (Van der Gulik et al., 2001b).
Fluid flow and mixing in an unbaffled horizontal stirred tank 77

Nomenclature
A constant in eq. A.7. -
B clearance m
Cµ,C1,C2 parameters in k-ε model
D vessel diameter m
E one dimensional power spectrum s/m2
G production of turbulent kinetic energy kg/(m⋅s3)
h blade height m
k turbulent kinetic energy m2/s2
L vessel length m
LI integral length scale m
p pressure Pa
Ri time autocorrelation function -
TI integral time scale s
u` velocity fluctuations m/s
U instantaneous velocity m/s
Ū time averaged velocity m/s
w blade length m
xL axial coordinate m

Greek

α angle between blade and shaft 60°


β angle between blades 180°
δij Kronecker delta -
ε energy dissipation rate m2/s3
µ dynamic viscosity kg/(m⋅s)
ρ density kg/m3
σk, σz parameters in k-ε model

Subscripts

a axial
i one of three directions
im impeller
P periodic
r radial
rms root mean square
t tangential
T turbulent
78 Chapter 4

References
Adrian, R.J.; Yao, C.S.; “Power spectra of fluid velocities measured by laser Doppler
velocimetry”, Exp. in Fluids, 1987, 5, 17.
Armenante, P.M.; Luo, C.; Chou, C.; Fort, I.; Medek, J.; “Velocity Profiles in a Closed,
Unbaffled Vessel: Comparison Between Experimental LDV Data and Numerical CFD
Predictions”, Chem. Eng. Sci., 1997, 52(20), 3483.
Bannenberg-Wiggers, A.E.M.; Van Omme, J.A.; Surquin, J.M.; “Process for the Batchwise
Preparation of Poly-p-terephtalamide”, U.S. Pat., 5,726,275, 1998.
Bird, R.B.; Stewart W.E.; Lightfooth, E.N.; “Transport phenomena”, John Wiley & Sons, US
1960.
Brucato, A.; Ciofalo, M.; Grisafi, F.; Micale, G.; “Numerical Prediction of Flow Fields in
Baffled Stirred Vessels: A Comparison of Alternative Modeling Approaches”, Chem.
Eng. Sci., 1998, 53(21), 3653.
Brucato, A.; Ciafalo, M.; Grisafi, F.; Tocco, R.; “On the Simulation of Stirred Tank Reactors
via Computational Fluid Dynamics”, Chem. Eng. Sci., 2000, 55, 291.
Ciofalo, M.; Brucato, A.; Grisafi, F.; Torraca, N.; “Turbulent Flow in Closed and Free-
Surface Unbaffled Tanks Stirred by Radial Impellers”, Chem. Eng. Sci., 1996, 51(14),
3557.
Clarke, D.S.; Wilkes, N.S.; “The calculation of turbulent flows in complex geometries using a
differential stress model”, UKAEA Report AERE-R 13428, Harwell, UK.
Cutter, L.A.; “Flow and turbulence in a stirred tank”, AIChE J., 1966, 12, 35.
Durst, F.; Melling, A.; Whitelaw, J.H.; “Principles and Practice of Laser Doppler
Anemometry”, Academic Press, London, UK, 1981.
Kehoe, A.B.; Prateen, V.D.; “Compensation for Refractive-Index Variations in Laser
Doppler Anemometry”, Applied Optics, 1987, 26(13), 2582.
Kresta, S.; “Turbulence in Stirred Tanks: Anisotropic, Approximate, and Applied”, Can. J.
Chem. Eng., 1998, 75, 563.
Kusters, K.A.; “The influence of turbulence on aggregation of small particles in agitated
vessels”, Ph.D. Thesis, Eindhoven University of Technology, The Netherlands, 1991.
Launder, B.E.; Reece, G.J.; Rodi, W.; “Progress in the Development of a Reynolds Stress
Turbulence Closure”, J. of Fluid Mechanics, 1975, 68, 537.
Launder, B.E.; Spalding, D.B.; “The Numerical Computation of Turbulent Flows”, Comp.
Meth. Appl. Mech. Eng., 1974, 3, 269.
Montante G.; Lee, K.C.; Brucato, A.; Yianneskis, M.; “Numerical Simulation of the
Dependency of Flow Pattern on Impeller Clearance in Stirred Vessels”, Chem. Eng. Sci.,
2001, 56, 3751.
Mujumdar, A.R.; Huang, B.; Wolf D.; Weber, M.E.; Douglas, W.J.M.; “Turbulence
parameters in a strirred tank”, Can. J. Chem. Eng., 1970, 48, 475.
Ng, K.; Fentiman, N.J.; Lee, K.C.; Yianneskis, M.; “Assessment of Sliding Mesh CFD
Predictions and LDA Measurements of the Flow in a Stirred by a Rushton Impeller”,
Trans. Inst. Chem. Eng. Res. Des., 1998, 76, 737.
Nieuwstad, F.T.M.; “Turbulentie, inleiding in de theorie en toepassingen van turbulente
stromingen (In Dutch)”, Epsilon uitgaven, The Netherlands, 1998.
Read, N.K.; Zhang, S.X.; Ray, W.H.; “Simulations of a LDPE Reactor using Computational
Fluid Dynamics”, AIChE J., 1997, 43(1), 104.
Rogers, M.M.; Mansour, N.N.; Reynolds, W.C.; “An algebraic model for the turbulent flux
of a passive scalar”, J. Fluid Mech., 1989, 203, 77.
Rousseaux, J.M.; Vial, C.; Muhr, H.; Plasari, E.; “CFD Simulation of Precipitation in the
Sliding-Surface Mixing Device”, Chem. Eng. Sci., 2001, 56, 1677.
Fluid flow and mixing in an unbaffled horizontal stirred tank 79

Schoenmakers, J.H.A.; “Turbulent Feed Stream Mixing in Agitated Vessels”, Ph.D. Thesis,
Eindhoven University of Technology, The Netherlands, 1998.
Schoenmakers, J.H.A.; Wijers, J.G.; Thoenes, D.; “Non Steady-State Behavior of the Flow
in Agitated Vessels”, Proc. 4th World Conf. on Experimental Heat Transfer, Fluid
Mechanics and Thermodynamics, ExHFT 4, Brussels, Ed. Giot, Mayinger & Celata,
1997, Vol. 1, 477-481.
Taylor, G.I.; “Statistical Theory of Turbulence Parts I-IV”, Proc. Roy. Soc., 1935, A151,
421.
Tennekes H.; Lumley, J.L.A.; “First course in turbulence”, MIT Press, USA, 1972.
Togatorop, A.; Mann, R.; Schofield, D.F.; “An Application of CFD to Inert and Reactive
Tracer Mixing in a Batch Stirred Vessel”, AIChE Symp. Series, 1994, 299, 19.
Van der Gulik, G.J.S.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodynamics and Scale-Up of
Horizontal Stirred Reactors”, Ind. Eng. Chem. Res., 2001a, 40(22), 4731.
Van der Gulik, G.J.S.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodynamics in a Horizontal
Stirred Tank Reactor”, Ind. Eng. Chem. Res., 2001b, 40(3), 785.
Van der Molen, K.; Van Maanen, H.R.E.; “Laser-Doppler measurements of the turbulent
flow in stirred vessels to establish scaling rules”, Chem. Eng. Sci., 1978, 33, 1191.
Van ‘t Riet, K.; Smith, J.H.; “The trailing vortex system produced by a Rushton turbine
agitators”, Chem. Eng. Sci., 1975, 30, 1093.
Vollbracht, L.; “Aromatic Polyamides”, Compr. Polym. Sci., Step Polym., 1989, 5, 374.
80 Chapter 4

Appendix 4A: Theoretical Aspects of LDA Measurements

Turbulent velocities
At turbulent conditions, instantaneous fluid velocities (U1, U2, U3) can be
represented by a mean and a fluctuating component, ( U1 + u1' , U 2 + u '2 , U 3 + u 3' ) ,
which is the so called Reynolds decomposition for three dimensions. The mean
flow in direction i, U i , is time averaged and is represented by:

lim  1 
t + ∆t
Ui = 
∆t → ∞  ∆t

∫t
U i dt 


[m/s] (4A.1)

The absolute averaged velocity is obtained using a vector summation:

2 2 2
U = U1 + U 2 + U 3 [m/s] (4A.2)

Often this is referred to as convective velocity or speed in CFD. The root-mean-


square (rms) value of the turbulent fluctuations is given by:

2 2 2
u rms = 1
2 (u 1' + u '2 + u 3' ) [m/s] (4A.3)

The intensity of the turbulent fluctuations, I, is equal to:

u rms
I= [-] (4A.4)
U

Turbulent fluctuations
The velocity fluctuations represent a significant amount of turbulent kinetic
energy. The contribution per unit of mass is equal to:

2 2 2
k T = 12 (u 1,' T + u '2,T + u 3,
'
T ) [m2/s2] (4A.5)

in which u i' ,T is the turbulent part of the velocity fluctuations that can be
distinguished from periodic fluctuations as will be discussed later. The energy kT is
initially present in large eddies generated by the impeller. The large eddies break-up
into smaller eddies transferring the energy to smaller scales until the Kolmogorov
scale is reached where the kinetic energy is dissipated into heat by viscosity. The
rate of energy transfer from large to small scale is thought to be proportional to kT
Fluid flow and mixing in an unbaffled horizontal stirred tank 81

divided by the time scale of the largest eddies i.e., L/u. The rate of energy transfer is
usually referred to as the energy dissipation rate ε:

u 3rms
ε~ [m2/s3] (4A.6)
L

This can be interpreted as the amount of energy u2rms that a turbulent eddy
dissipates in a timescale T~urms/L. Using the ‘inviscid estimate’ as postulated by
Tennekes and Lumley (1972), ε can be rewritten as:
3
k T2
ε=A [m2/s3] (4A.7)
LI

‘A’ has been found to be equal to 1 ± 0.2 which is to be expected based on the
energy balance (Cutter, 1966; Kusters, 1998; Nieuwstadt, 1998).

Integral length and time scales


LI represents the integral length scale of the eddies. Using Taylor’s hypothesis
(1935), LI can be calculated from experimentally determined local time-averaged
velocity Ū and the integral time scale TI. This hypothesis assumes that temporal
velocity fluctuations measured in an Eulerian frame can be interpreted as spatial
turbulence along a line parallel to the average convective velocity Ū. Hence, it
follows that:

L I = U ⋅ TI [m] (4A.8)

Another formulation of Taylor’s hypothesis is that the average convective velocity


Ū transports the eddies with such a high speed through a measurement point that the
eddies do not change during the measurement. The turbulent fluctuations then
appear to be frozen.
The integral time scale TI is equal to the sum of the time scales for the three
directions:

3
TI = ∑T
i =1
i [s] (4A.9)

The time scale for each direction i can be obtained from zero frequency intercept of
the one-dimensional power spectrum Ei(f) as proposed by Mujumdar et al. (1970):

1 lim
Ti = E i (f ) [s] (4A.10)
4u i'
2 f →0
82 Chapter 4

Ei(f) is the Fourier cosine transform of the time autocorrelation function Ri(τ):


2
∫ R (τ) e
2 πifτ
E i (f ) = 4u i' i dτ [m2/s] (4A.11)
0

with

u i ( t ) ⋅ u i ( t + τ)
R i (τ) = [-] (4A.12)
u i2

τ is the time over which the correlation is being determined. The length scale of the
largest eddies are expected to be equal to the height of the impeller blades.
An equidistant time signal is required to calculate spectra or auto correlation
functions of a signal, which is not obtained from LDA measurements due to the
Poison-like shape of the inter arrival time distribution. The interpolation procedure
used to obtain an equidistant velocity signal, is the Sample-Hold technique
proposed by Adrian and Yao (1987). The first moment of the velocity signal proved
to deviate only a few percent from the variance of the velocity signal from spectral
analysis using the Sample-Hold as a interpolation procedure.

Periodic fluctuations
The flow in the Drais reactor is mechanically driven by the impeller. In this
type of flow the velocity fluctuations can be split up in a turbulent and periodic part.
The turbulent part is random. The periodic part results from the trailing vortices
coming from behind the rotating stirring device (Van ‘t Riet and Smith, 1975). The
periodic part has a frequency equal to the impeller frequency multiplied by the
number of revolutions. Also sub- and higher harmonics occur. Van der Molen and
Van Maanen (1978) showed that the turbulent and periodic parts are not correlated
and can be separated using:

2 2 2
u i' = u i' ,T + u i' ,P [m2/s2] (4A.13)

Periodic fluctuations are visible as strong peaks in the power spectrum which
can be constructed using equation 4A.11. The contribution of the periodic
fluctuations can be obtained by integrating the corresponding peaks. The remaining
frequencies are assigned to the turbulent fluctuations which are used to obtain the
turbulent kinetic energy in equation 4A.5.
Fluid flow and mixing in an unbaffled horizontal stirred tank 83

Appendix 4B: Computational Fluid Dynamics

The equations that are solved in Computational Fluid Dynamics for a turbulent
flow are the Reynolds-averaged continuity and momentum equations. With constant
density and viscosity, the continuity equation has the following form for a Cartesian
grid:

 ∂U x ∂U y ∂U z 
ρ div(U i ) = ρ  + + =0 [kg/(m3⋅s)] (4B.14)
 ∂x ∂y ∂z 

The momentum equation for the x-component has the following form:

 ∂ U x ∂(U x U x ) ∂(U x U y ) ∂(U x U z )  ∂p  ∂ u ' u ' ∂ u 'y u 'x ∂ u ' u ' 


ρ + + +  =− + µ∇ 2 U x − ρ x x + + z x 
 ∂t ∂x ∂y ∂z  ∂x  ∂x ∂y ∂z 
 
[kg/(m2⋅s2)] (4B.13)

or more condensed:

 ∂Ux  ∂p
ρ + div(U x U i ) = − + µ∇ 2 U x − div(τ x ) [kg/(m2⋅s2)] (4B.14)
 ∂t  ∂x

For the y and z-direction similar equations can be derived. τx, τy, and τz are the row
vectors comprising the traceless part of the turbulent stress tensor τij. These vectors
are the components that represent the Reynolds stresses acting on surfaces of
normal x, y, and z, respectively, with the normal components depleted by p. For the
closure of these equations, turbulence models are used.

The k-ε turbulence model


The first model used is the well-known k-ε model which is based on an eddy
viscosity hypothesis (Launder and Spalding, 1974). The Reynolds stresses are
modeled as stress terms with a viscosity µT that depends on flow properties. In most
general form, the stresses can be written as:

 ∂ Ui ∂ U j  2
− ρu i' u 'j = µ T  +  − ρ kδij [kg/(m⋅s2)] (4B.15)
 ∂x j ∂x i  3

in which i and j refer to either x, y or z. The term 2/3ρkδij, in which δij is the
Kronecker delta (Bird et al., 1960), is added so that continuity and the definition of
turbulent kinetic energy are satisfied. The turbulent viscosity is described as:
84 Chapter 4

k2
µ T = cµρ [kg/(m⋅s)] (4B.16)
ε

k and ε are the turbulent kinetic energy and energy dissipation rate, respectively, for
which the following equations apply:

∂k ∂k ∂  µ T ∂k 
ρ + ρU j =   + G − ρε [kg/(m⋅s3)] (4B.17)
∂t ∂x j ∂x j  σ k ∂x j 
 

∂ε ∂ε ∂  µ T ∂ε 
ρ + ρU j =   + (C1 G − C 2 ρε) ε [kg/(m⋅s4)] (4B.18)
∂t ∂x j ∂x j  σ ε ∂x j  k
 

In these equations G is the production of turbulent kinetic energy by deformation


and is equal to:

∂Ui
G = −ρu i' u 'j [kg/(m⋅s3)] (4B.19)
∂x j

The standard values of the constants in the model were used: Cµ=0.09, C1=1.44,
C2=1.92, σk=1.0, and σε =1.3.

The DSM turbulence model


The second model used, was the DSM (Differential Stress Model) by Launder
et al. (1975). In this model the turbulent stresses in the stress tensor τij were
computed by solving the six independent transport equations, that have the
following general form:

div(ρU i τij ) = G ij + div(d ij ) + φij + ε ij + Fij [kg/(m⋅s3)] (4B.20)

In this equation the terms on the right-hand side represent generation, diffusion,
pressure strain correlation, dissipation and generation by fluctuating body forces,
respectively. In the DSM a further transport equation for the dissipation ε is also
solved, while the turbulence energy kT can be computed as the half sum of the
diagonal terms τkk. The DSM is especially suitable for problems in which
turbulence is expected to be strongly anisotropic in time and space. The
disadvantage of the DSM is the numerical cost of solving 10 strongly coupled
partial differential equations. The details of the model formulation are cumbersome
and will not be given here. The implementation of the DSM in the CFX-4 code is
presented by Clarke and Wilkes.
Fluid flow and mixing in an unbaffled horizontal stirred tank 85

Scalar mixing
The transport of a passive scalar in a turbulent flow field can be described by
the following equation (after Reynolds decomposition and time averaging):

' '
∂C ∂C ∂ 2 C ∂u i c j
+ Ui =D 2 − [1/s] (4B.21)
∂t ∂x i ∂x i ∂x i

In this expression the turbulent transport of the scalar u i' c 'j is closed by invoking the
gradient diffusion assumption:

∂C
u i' c 'j = −ΓT [m/s] (4B.22)
∂x j

ΓT can be interpreted as an effective turbulent diffusion coefficient. Using the k-ε


model ΓT has the following form:

µT
ΓT = [m2/s] (4B.23)
ρσ c

in which µT is the previously mentioned turbulent viscosity and σc the turbulent


Schmidt number which is equal to 0.7. The turbulent diffusion coefficient in the k-ε
model implies isotropic behavior in that ΓT is the same in all directions. Using the
DSM, ΓT is treated as a tensor for which the details are given by Rogers et al.
(1989). This tensor invokes the non-isotropic mixing of the passive scalar.
HYDRODYNAMICS IN A CERAMIC
5
PERVAPORATION MEMBRANE
REACTOR FOR RESIN PRODUCTION

Abstract

The hydrodynamics in a pervaporation membrane reactor for resin


production has been investigated. In this type of reactor it is important
to reduce concentration and temperature polarization to obtain high
water fluxes during operation. The influence of secondary flow on
polarization, as induced by small density differences, is studied using
Computational Fluid Dynamics in a model system. This model is
operated in three parallel flow situations: horizontal, vertical opposed,
and vertical adding flow. Density-induced convection is found to be
most effective in the horizontal situation, increasing water fluxes up to
50%. Water fluxes were also determined experimentally using the
horizontal set-up. The influence of density-induced convection was
observed experimentally.

This chapter is a slightly modified version of the publication:


Van der Gulik, G.J.S.; Janssen, R.E.G.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodynamics in
a Ceramic Pervaporation Membrane Reactor for Resin Production”, Chem. Eng. Sci., 2001,
56, 371-379.
88 Chapter 5

5.1 Introduction
Resins are polycondensation products, formed by the following equilibrium
reaction:

n R1 OH + n R2 COOH Resin + H2O (R.7)

Reaction conditions require elevated temperatures, typically above 200 °C. The
mixtures are moderately viscous at these temperatures (µ = 5-25 mPa·s). To obtain
high yields, the reaction equilibrium has to be shifted to the right. This can be
achieved using a large excess of one of the starting reagents (usually the alcohol,
which is also used as the solvent). However, this results in a relatively inefficient
use of reactor space, and requires an efficient separation afterwards. The reaction
can also be forced to completion by removing the water. In current processes water
is removed by distillation, which is not very efficient in the case of azeotrope
formation. Additionally, due to the large reflux ratios required, energy consumption
can be significant (Keurentjes et al., 1994).
Alternatively, pervaporation membranes can be used for selective water
removal. This has shown to be effective for low molecular weight esterifications
(Keurentjes et al., 1994; David et al., 1991; Okamoto et al., 1993). For this purpose,
polymeric membranes have been used, mainly based on polyvinyl-alcohol.
However, in resin production only ceramic membranes are appropriate, as they are
resistant to the high reaction temperatures (Bakker et al., 1998; Koukou et al., 1999;
Verkerk et al., 2001). These membranes are selective to water, as they are
extremely hydrophilic, and have pores that are about the kinetic diameter of water
(Van Veen et al., 1999).
It can be anticipated that concentration and temperature polarization will be a
major problem in the application of ceramic membranes for the production of
resins. The polarization effects are schematically depicted in Figure 5.1. As water is
transported through the membrane, the water concentration near the membrane
surface will be low. Resin concentration will be high, as the reaction will locally be
in equilibrium. Also, the temperature will be low, as water evaporates at the
membrane surface adiabatically. Consequently, the viscosity near the membrane
will be relatively high. As a result, a relatively thick stagnant layer will be formed,
severely reducing water transport from bulk to the membrane surface.
To reduce these polarization effects, the hydrodynamics near the membrane
has to be optimized. Applying high turbulence levels in the main flow field can
effectively reduce polarization effects. In laminar flows, Dean vortices can be
induced by curved polymer (Moulin et al., 1999; Winzeler et al., 1993) or ceramic
membranes (Broussous et al., 1997). Also, vibrating devices have been proposed
(Vane et al., 1999).
Hydrodynamics in a pervaporation membrane module 89

Membrane

Water concentration

Temperature

Arbitrary units Water

Viscosity

Resin concentration
δT δC

Figure 5.1: Concentrations, temperature and viscosity in


arbitrary units near the membrane surface.

The application of high turbulence levels is not appropriate in the resin


production process, because we aim to develop a membrane reactor that is operated
in a once-through continuous mode. Therefore, flows have to be low in order to
limit equipment dimensions. Additionally, plug flow characteristics are required to
obtain a narrow molecular weight distribution (MWD).
The objective of this study is to design a membrane module, in which
hydrodynamics is optimized for maximum water removal. For this purpose, the
hydrodynamics in an annular membrane module has been investigated. Model
calculations have been performed using computational fluid dynamics (CFD).
Additionally, flux measurements have been performed in an experimental set-up.

Tank
T = Temperature sensor (10 l.) Heater

T T Recycle
T
v(x,r)

Mixture v(x,r)
P
80°C
1 bar
Inlet pipe Membrane Outlet pipe
Vacuum pump
Module

Figure 5.2: Schematic representation of the experimental set-up.


90 Chapter 5

5.2 Experimental

5.2.1 The model system


The model system, used in this study, is depicted in Figure 5.2. The horizontal
module consisted of a glass pipe with a length of 1650 mm, outer and inner
diameter of 50 and 40 mm, respectively. The hydrophilic membrane (supplied by
ECN, The Netherlands, (Van Veen et al., 1999) consists of a support tube, made of
α-Al2O3, on which on the outside an amorphous SiO2 top layer is applied with an
average pore diameter smaller than 0.5 nm and a thickness δ of 200 nm. Porosity is
estimated to be between 50 and 80%. The length of the membrane was 900 mm,
with an outer and inner diameter of 14 and 8 mm, respectively. An inlet pipe, with a
length of 350 mm, was positioned in front of the membrane to ensure a fully
developed flow. An outlet pipe (250 mm) was positioned after the membrane to
minimize disturbances due to outflow effects.
Aqueous solutions of glycerin or 1,4-butanediol (Heybroek, Amsterdam,
Purity > 99.7%) were fed under laminar conditions. The flow direction is parallel to
the membrane. Weight percentages of glycerin and 1,4-butanediol up to 75% were
used, in order to provide moderately viscous conditions. Reynolds numbers, as
defined by ρvsupdh/µ, were below 300.

5.2.2 Flux measurements


Flux measurements were performed at several superficial velocities and
temperatures. To evaluate changes in membrane properties, a pure water flux was
measured before and after changes in fluid composition. The permeate was
condensed in a cold trap with liquid N2. Permeate pressure ranged from <1 mbar to
50 mbar, depending on composition and temperature.
The water flux φw through the membrane was found to depend linearly on the
driving force over the membrane (Verkerk et al., 2001):

C
φw = ⋅ ( Pw* ,r − Pw* , p ) [kg/(m2⋅h)] (5.1)
δ

in which C is a constant, and δ the thickness of the top layer of the


membrane. Pw* ,r and Pw* , p represent the equilibrium vapor pressure of water in the
retentate and permeate (vapor system), respectively. Pw* ,r is equal to:

Pw* ,r = γ w ⋅ xw ⋅ Pw° [Pa] (5.2)


Hydrodynamics in a pervaporation membrane module 91

in which γw is the activity coefficient of water in the mixture, xw is the mole fraction
of water, and Pw° is the vapor pressure of the pure water. γw is calculated with the van
Laar correlation for activity coefficients. Pw* , p is equaled to the absolute pressure in
the vacuum system when the separation selectivity is high. This selectivity is
defined by:

ww , p ( 1 − ww , p )
S= [-] (5.3)
ww ,r ( 1 − ww ,r )

with ww,p and ww,r as water fraction in the permeate and retentate, respectively. In
order to determine separation selectivity, the permeate composition was measured
with HPLC for glycerin/water and refractometry for 1,4-butanediol/water mixtures.
Karl-Fisher titration was used to determine retentate compositions.
92 Chapter 5

5.3 CFD simulations


Flow simulations on the model system were performed using CFX 4.2, a finite
volume fluid dynamics package (AEA Technology, Harwell, UK) which was
installed on a Silicon Graphics Origin 200 workstation (IRIX 6.4). To study the
contribution of natural convection in the horizontal set-up, 3D simulations have
been undertaken on a half-tube. The grid used (90×30×25), is depicted in Figure
5.3. A symmetry plane is applied in the x-z-plane, as the problem is considered
symmetrical, thus significantly reducing computation time. In the vertical
configuration of the model system, 2D calculations were sufficient, as the system
becomes rotation symmetrical. Therefore, the z-component was omitted in these
cases. To demonstrate the independence of the solution from the grid, the
simulations were repeated with 1.5 times the number of cells without a noticeable
change in solution.
The CFD simulations were only performed on 60/40 w% glycerin/water
mixtures at 80°C. The water weight fraction (ww) is simulated as a scalar that does
not represent any volume. All fluid properties depend on this scalar and on
temperature as given in the appendix. The membrane is simulated as a wall patch,
on which a sink term is defined for the scalar. The water flux is equal to:

φw = C⋅P*w,m [kg/(m2⋅h)] (5.4)

with C as a constant, and P*w,m as the equilibrium feed-phase vapor pressure of


water at the membrane calculated using the Van Laar correlation for activity

14

z x
y
40 Outlet pipe (250)

Membrane (900)

Inlet pipe (350)

Figure 5.3: Representation of the grid (90×30×25) of the model system. Geometric
progression of the grid was 1.1 in all directions. Distances are given in mm.
Hydrodynamics in a pervaporation membrane module 93

coefficients. Required data were obtained from Gmehling et al. (1980). C is chosen
such that φw is equal to 1.82 kg/m2⋅h based on inlet conditions. The temperature at
the membrane decreases adiabatically and is accounted for in a sink term as
follows:

φh = φw ⋅∆Hv [W] (5.5)

with ∆Hv as the heat of vaporization of water.


Special attention is given to obtain a converged solution. Incorporating natural
convection results in rather unstable CFD simulations. Underrelaxation factors
down to 0.0001 for temperature and v-velocity (parallel to gravity vector) are
needed. In addition, 5 inner iterations were performed on temperature before
changes were implemented on the body force. Up to 40,000 iterations was needed
to obtain a converged solution, requiring over three day’s simulation time for the
full 3D simulation. The upwind scheme was used in all calculations.

Water (5 mm/s) Water (8.3 mm/s)


66% 1,4-Butanediol (5 mm/s) 66% 1,4-Butanediol (8.3 mm/s)
60% Glycerin (5 mm/s) 60% Glycerin (8.3 mm/s)
12
10
8
φ w [kg/m .h]
2

6
4
2
0
0 50 100 150 200 250 300
° °
P w,r-P w,p [mbar]

Figure 5.4: Water flux as a function of the difference in retentate and


permeate water vapor pressure.
94 Chapter 5

5.4 Results

5.4.1 Flux measurements


In Figure 5.4, the flux is given as a function of the difference between the
equilibrium vapor pressure of water in the retentate and permeate. In this figure, the
water flux is presented for two superficial velocities for pure water, glycerin, and
1,4-butanediol mixtures. The influence of the superficial velocity on the flux
appears to be small, indicating that hydrodynamics is mainly determined by natural
convection. The observed selectivity appeared to be close to 100%. In the cold trap
neither glycerin nor 1,4-butanediol could be detected. However, selectivity was
thought to be around 99.5% as small condensed droplets were formed in the outlet
pipe, containing glycerin or 1,4-butanediol.

5.4.2 CFD simulations


CFD simulation in which natural convection is omitted

Initially, CFD simulations have been performed, in which natural convection is


omitted, creating a 2-dimensional flow problem that is fully determined by forced
convection. The results in Figures 5.5a-f show the relevant variables in a slice
through the module at x = 0.2 m. The profile of the axial velocity over the annulus
in Figure 5.5a is in agreement with the profile according to laminar conditions. The
water fraction ww, temperature T, density ρ, and viscosity µ are given in Figures
5.5b-e, respectively. From Figures 5.5b and 5.5c it can be observed that the water
fraction (ww) profile over the annulus is less developed than the temperature profile.
Comparison of density and viscosity profiles with the water fraction profile shows
that density and viscosity are mainly determined by the water fraction, as the
profiles show a strong resemblance. Additionally, it can be seen that relatively
strong concentration and temperature polarization occurs.

Horizontal set-up with natural convection.

When natural convection is incorporated, gravity force induces complex


secondary flow as a result of density variations. In Figure 5.6, the induced flow
pattern is depicted in a slice through the module at x = 0.2 m. Near the membrane
the flow direction is downwards, while flow is upwards in the bulk, creating a
circular flow next to the membrane. Two circular flow regions are present below
and above the membrane, resulting in a total of three circulation regions. Hattori et
al. (1979) also observed three circulation zones. The highest downward velocity
occurs below the membrane tube.
Hydrodynamics in a pervaporation membrane module 95

a: u [m/s] b: ww [-] c: T [K] d: ρ [kg/m3] e: µ [kg/(m⋅s)]

Figure 5.5a-e: Axial velocity u, fraction water ww, temperature T, density ρ, and viscosity µ
in the simulation without natural convection (vsup = 5⋅10-3 m/s, x = 0.2 m).

a: u [m/s] b: v [m/s] c: ww [-] d: T [K] e: ρ [kg/m3] f: µ [kg/(m⋅s)]

Figure 5.7a-f: Axial velocity u, z-velocity v, fraction water ww, temperature T, density ρ, and
viscosity µ in the simulation with natural convection (vsup = 5⋅10-3 m/s, x = 0.2m).
96 Chapter 5

vsup = 0.005 m/s vsup = 0.1 m/s


a: u [m/s] b: ww [-] c: T [K] d: u [m/s] e: ww [-] f: T [K]

Flow direction

Figure 5.8a-f: Axial velocity u, weight fraction water ww, and temperature T for the
simulations with opposing natural convection at Vsup = 0.005 and 0.1 m/s. The horizontal lines
indicate the location of the membrane.

v = 0.005 m/s v = 0.1 m/s


a: u [m/s] b: ww [-] c: T [K] d: u [m/s] e: ww [-] f: T [K]
Flow direction

Figure 5.9a-f: Axial velocity u, weight fraction water ww, and temperature T for the
simulations with adding natural convection for Vsup = 0.005 and 0.1 m/s. The horizontal lines
indicate the location of the membrane.
Hydrodynamics in a pervaporation membrane module 97

Figure 5.7a-f shows all relevant parameters in


the same slice as in Figure 5.5 (at x = 0.2 m).
Figures 5.7c-f show that the stagnant layers, as
observed in Figure 5.5, are largely reduced. As
a result, concentration and temperature
polarization are effectively reduced. Fluid with
low temperature and low water fraction is
concentrated at the bottom of the annulus.
Consequently, this region will have a higher
viscosity, slowing down the forced convection
locally. This behavior is observed in Figure
5.7a.

Vertical set-up with opposing natural


convection

In figures 5.8a-f, the v-velocity, weight


fraction water, and the temperature are given for
the vertical set-up with upward flow at
superficial velocities vsup = 0.005 (5.8a-c) and
0.1 m/s (5.8d-f), respectively. In this situation,
the natural convection acts opposed to the Figure 5.6: Secondary flow in a
forced convection. In figure 5.8a, the position of horizontal positioned annulus,
the membrane is depicted at the left. It can be induced by free convection with
seen that at low superficial velocities, a strong vsup = 5⋅10-3 m/s at x = 0.2 m.
downward flow is induced at the membrane
surface. Flow behavior is fully determined by natural convection. As a result, the
cooled mixture subsides towards the entrance. At higher superficial velocities, no
downward flow occurs. The cooled mixture is then pushed upwards to the outlet.

Vertical set-up with adding natural convection

In Figures 5.9a-f, the v-velocity, weight fraction water, and the temperature are
given for the vertical set-up with downward flow at superficial velocities vsup =
0.005 m/s (5.9a-c) and 0.1 m/s (5.9d-f). In these situations, the natural convection is
called adding. Figure 5.9a shows that at low superficial velocities, natural
convection accelerates the downward flow near the membrane. At higher velocities,
no acceleration is observed, resulting in flow conditions that are fully determined by
forced convection.

Nu and Sh numbers

For all simulations, the heat and mass transfer along the membrane can be
expressed in the dimensionless heat and mass transfer numbers Nu and Sh. These
were calculated using the following equations:
98 Chapter 5

φ w ( x ) ⋅ d h ⋅ 3600
Sh ( x ) = [-] (5.6)
D in ⋅ ρ in ⋅ ( w in m
w − w w ( x ))
φ w ( x ) ⋅ ∆ H v ⋅ d h ⋅ 3600
Nu ( x ) =
λ in ⋅ ( Tin − T m ( x )) [-] (5.7)

with φw as the water flux, dh the hydraulic diameter, Dw the diffusion coefficient of
water, ρin the inlet density, wwin the weight fraction water at the inlet, wwm the weight
fraction water at the membrane, ∆Hv the heat of vaporization, λin the inlet
conductivity, Tin the inlet temperature, and Tm the temperature at the membrane.
In Figures 5.10 and 5.11, the Nu and Sh numbers are plotted for four flow
situations described above: without natural convection, horizontal flow, vertical
opposing, and vertical adding flow. In Figure 5.10a and 5.11a, the situations with
and without natural convection are compared for the horizontal situation. Nu and Sh
decline very rapidly for both situations. However, with natural convection, Nu and
Sh stay on a level that is 4 times higher than without natural convection. The
enhanced heat and mass transfer for this situation is evident.
The Sh and Nu numbers for the vertical situation with adding natural
convection, in Figures 5.10b and 5.11b, resemble the situation without natural
convection. At high superficial velocities, the numbers are comparable as forced
convection dominates. At low superficial velocities, the numbers are considerably
higher than in the situation without natural convection.
Table 5.1: Dimensionless groups for the three different flow situations (Pr=19.4, Sc=735).
Horizontal Vertical, opposing flow Vertical, adding flow

vsup Gr Pr Grd h Gr Pr Grd h Gr Pr Grd h


Re Gr Gr Gr
[m/s] Re 2 L Re 2 L Re 2 L
(·10-5) (·10-5) (·10-6) (·10-5) (·10-6) (·10-5)
0.001 15.7 - - - 1.20 4854 6.72 1.26 5085 7.04
0.002 31.5 - - - 1.17 1177 6.52 1.10 1108 6.13
0.005 78.7 7.93 128 4.43 1.13 182 6.30 1.08 175 6.06
0.01 157 7.70 31.1 4.31 1.11 45.0 6.22 1.06 42.8 5.93
0.02 315 7.58 7.55 4.18 1.12 11.3 6.26 1.03 10.4 5.77
0.05 786 7.31 1.18 4.09 1.39 2.25 7.79 9.51 1.54 5.31
0.1 1571 7.25 0.29 4.05 1.05 0.42 5.86 8.46 0.34 4.73

For the vertical opposing situation, the Sh and Nu curves in Figures 5.10c and
5.11c are similar to the adding flow situation at high velocities. For both situations,
flow behavior is determined by forced convection. However, at low velocities, all
Sh and Nu numbers coincide as flow behavior is fully determined by natural
convection. At vsup = 0.05 m/s, a transition takes place from natural convection to
forced convection, determining the flow behavior.
The ratio Gr/Re2 is often used to estimate which flow behavior prevails. Gr is
the Grashof number representing the ratio of buoyancy and viscous forces:
Hydrodynamics in a pervaporation membrane module 99

Gr = g ⋅ d h3 ⋅ ρ ⋅ ∆ρ µ 2 [-] (5.8)

Re is defined as:

Re = v ⋅ d h ⋅ ρ ⋅ µ [-] (5.9)

In the vertical opposing flow situation, the transition takes place at vsup = 0.05
m/s which corresponds to Gr/Re2 = 2.25, as given in Table 5.1. In the other two
situations, this transition is not obvious from the results presented here.
Nevertheless, it can be expected that it will be at about the same number.

40 140
35 A: Horizontal 120 A: Horizontal
30
100
With natural convection 0.005
25 0.005
Sh(x) [-]

80
Nu(x) [-]

0.01 With natural convection


20 0.01
0.02 0.02
60
15 0.05 Without natural convection 0.05
Without natural convection
0.1 40 0.1
10

5 20

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
x [m] x [m]

50 140
45
40
B: Vertical, adding 120 B: Vertical, adding
0.001 100
35
0.002 0.001
30
Nu(x) [-]

0.005 80 0.002
Sh(x) [-]

25 0.01 0.005
20 60 0.01
0.02
15 0.05 0.02
40
0.1 0.05
10
20 0.1
5
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
x [m] x [m]

50 140
45
40
C: Vertical, opposing 120 C: Vertical, opposing
0.001 100
35 0.001
0.002
30 0.002
Nu(x) [-]

0.005 80
Sh(x) [-]

0.005
25 0.01
60 0.01
20 0.02
0.02
15 0.05 40 0.05
0.1
10 0.1
20
5
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
x [m] x [m]

Figure 5.10a-c: Nu-numbers along the Figure 5.11a-c: Sh-numbers along the
membrane for horizontal, vertical adding, membrane for horizontal, vertical adding,
and vertical opposing flow. and vertical opposing flow.
100 Chapter 5

1.6

Horizontal
1.4

Flux [kg/m2.h]
1.2 Vertical, adding

1.0 Vertical, opposing

0.8
Without natural convection

0.6
0.001 0.010 0.100
Superficial velocity [m/s]

Figure 5.12: Flux versus superficial velocity for simulations with and without natural

In a situation with only natural convection, the Rayleigh number Ra, as


defined by Pr·Gr·d/L, determines the transition from laminar to turbulent flow. Ra
exceeds 109 under turbulent conditions, which is not obtained in our configurations,
as can be seen in Table 5.1. In a situation with only forced convection, turbulence
can also be obtained by applying high superficial velocities. This transition occurs
at Re = 2300, which is also not reached in our situation. In mixed convection
situations, turbulent conditions can already be obtained at Re numbers around 900
at Rayleigh numbers of 105, for both horizontal and vertical configurations (Metais
and Eckert, 1964). The occurrence strongly depends on the presence of secondary
flow (Hallman, 1961). Some of our simulations have been carried out at Re values
around 900 and higher, while a strong secondary flow is observed. Therefore, our
presumed laminar flow may not be valid, which will be subject for further research.

Simulated fluxes

In Figure 5.12 the obtained water fluxes are plotted against superficial
velocity. It is observed that the horizontal set-up provides the highest fluxes.
Natural convection increases the fluxes up to 50% at low superficial velocities.
Calculations were performed down to 5 mm/s for the horizontal set-up. Below this
velocity, the flow problem is fully determined by natural convection. Numerics then
become unstable, indicating that the solution will be time dependent. Therefore,
these results are not included. At high superficial velocities all flow situations
become determined by forced convection. As a result, all fluxes tend to coincide.
When the velocity will be increased further to fully turbulent conditions, with no
polarization, the maximum flux of 1.82 kg/m2⋅h will be obtained (based on inlet
conditions). Note that the horizontal set-up is only 23 % below this maximum. This
emphasizes the significant contribution of the natural convection to the performance
of the membrane module.
Hydrodynamics in a pervaporation membrane module 101

5.5 Conclusions
This report presents the occurring hydrodynamics in a model system of a
membrane reactor for resin production. Flux measurements have been performed
experimentally at different superficial velocities, temperatures, and fluid
compositions, showing that hydrodynamics is determined by forced and natural
convection simultaneously.
CFD calculations have also been performed on the model system These
calculations show that the hydrodynamics is indeed determined by forced and
natural convection, simultaneously. It was shown that natural convection is
effective in reducing concentration polarization. Natural convection has been
studied in three different configurations: horizontal, vertical adding, and vertical
opposing flow. The horizontal configuration showed an increase in water flux up to
50% compared to a simulation in which natural convection is omitted. The
dimensionless heat and mass transfer coefficients Nu and Sh are up to 4 times
higher.

Nomenclature
A12,A21 binary parameters in Van Laar correlation for activity coefficients -
Cp heat capacity J/(kg⋅K)
dh hydraulic diameter m
D diffusivity m2/s
Gr Grashof number -
∆HV heat of vaporization J/kg
∆HV0 heat of vaporization at T0 J/kg
L membrane length m
M molecular weight g/mol
Nu Nusselt Number
Pr Prandtl number -
P° vapor pressure of pure component Pa
P* equilibrium vapor pressure Pa
Ra Rayleigh number -
Re Reynolds number -
S separation selectivity -
Sc Schmidt number -
Sh Sherwood number -
T temperature K
Tc critical Temperature K
u axial velocity m/s
v velocity in z-direction m/s
V molar volume m3/kmol
vsup superficial velocity m/s
w weight fraction -
x mole fraction -
102 Chapter 5

Greek

δ thickness of the top layer m


φw flux of water kg/(m2⋅h)
φh heat flux J/(m2⋅h)
γ activity coefficient -
λ heat conductivity J/(m⋅s⋅K)
µ dynamic viscosity kg/(m⋅s)
ρ density kg/m3

Super and subscripts

g glycerin
in inlet conditions
m at the membrane surface
mix mixture of water and glycerin
p permeate
r retentate
w water

References
Bakker, W.J.W.; Bos, I.A.A.C.M.; Rutten, W.L.P.; Keurentjes, J.T.F.; Wessling, M.;
“Application of ceramic pervaporation membranes in polycondensation reactions”, Int.
Conf. Inorganic Membranes, Nagano, Japan, 1998, 448-451.
Broussous, L.; Ruiz, J.C.; Larbot, A.; Cot, L.; “Stamped ceramic porous tubes for tangential
filtration”, Proc. Euromembrane '97, Twente, The Netherlands, 1997, 394-396.
David, M.-O.; Gref, R.; Nguyen, T.Q.; Néel, J.; “Pervaporation-esterification coupling: Part
I, Basic kinetic model”, Trans. Inst. Chem. Eng., 1991, 69A, 335-340.
Gmehling, J.; Oncken, U.; Rary-Nies, J.R.; “Vapor-liquid equilibrium data collection,
Aqueous systems”, Suppl. 2, Dechema Chemistry data series, I-1b, 1980, 182-188.
Hallman, T.M.; “Combined forced and free-laminar heat transfer in a vertical tube with
uniform internal heat generation”, Trans. ASME, 1956, 78, 1831-1841.
Hattori, N.; “Combined free and forced-convection heat-transfer for fully developed laminar
flow in horizontal concentric annuli (numerical analysis)”, Heat transfer Jpn. Res., 1979,
8, 27.
Keurentjes, J.T.F.; Janssen, G.H.R.; Gorissen, J.J.; “The esterification of tartaric acid with
ethanol: kinetics and shifting the equilibrium by means of pervaporation”, Chem. Eng.
Sci., 1994, 49, 4681-4689.
Koukou, M.K.; Papayannakos, H.; Markatos, N.C.; Bracht, M.; Van Veen, H.M.; Roskam,
A.; “Performance of ceramic membranes at elevated pressure and temperature: effect of
non-ideal flow conditions in a pilot scale membrane separator”, J. Membr. Sci., 1999,
155, 241-259.
Metais, B.; Eckert, E.R.G.; “Forced mixed and free convection regimes”, J. Heat Transfer,
1964, 86C, 295-296.
Moulin, P.; Manno, P.; Rouch, J.C.; Serra, C.; Clifton, M.J.; Aptel, P.; “Flux improvement
by Dean vortices: ultrafiltration of colloidal suspensions and macromolecular solutions”,
J. Membr. Sci., 1999, 156, 109-130.
Newman, A.A.; “Glycerol”, Morgan-Gampian, London, 1968.
Hydrodynamics in a pervaporation membrane module 103

Okamoto, K.; Yamamoto, M.; Otoshi, Y.; Semoto, T.; Yano, M.; Tanaka, K.; Kita, H.;
“Pervaporation-aided esterification of oleic acid”, J. Chem. Eng. Japan, 1993, 26, 475-
481.
Reid, R.C.; Prausnitz, J.M.; Poling, B.E.; “Properties of gases and liquids”, 4th ed., Mc
Graw-Hill Book Co, 1988.
Van Veen, H.M.; Van Delft, Y.C.; Engelen, C.W.R.; Pex, P.P.A.C.; “Dewatering of organics
by pervaporation with silica membranes”, Proc. Euromembrane '99, Leuven, Belgium,
1999, 209.
Vane, L.M.; Alvarez, F.R.; Giroux, E.L.; “Reduction of concentration polarization in
pervaporation using a vibrating membrane module”, J. Membr. Sci., 1999, 153, 233-241.
Vargaftik, N.B.; Vinogradov, Y.K.; Yargin, U.S.; “Handbook of physical properties of
liquids and gases: pure substances and mixtures”, 3rd augm. and rev. ed., Begell Mouse,
New York, 1996.
Verkerk, A.W.; Van Male, P.; Vorstman, M.A.G.; Keurentjes, J.T.F.; “Description of
dehydration performance of amorphous silica pervaporation membranes”, J. Membr.
Sci., 2001, 193(2), 227-238.
Vignes, A.; “Diffusion in binary solutions”, Ind. Eng. Chem. Fundam., 1966, 5, 189.
Weast, P.C.; Astle, M.J.A.; “Handbook of chemistry and physics”, 63rd edition, CRC Press,
1982.
Wilke, C.R.; Chang, P.; “Correlation of diffusion coefficients in dilute solutions”, AIChE J.,
1955, 1, 264.
Winzeler, H.B.; Belfort, G.; “Enhanced performance for pressure-driven membrane
processes: the argument for fluid instabilities”, J. Membr. Sci., 1993, 80, 35-47.
104 Chapter 5

Appendix

Heat of vaporization of water (Reid et al., 1988).


0.38
 1 − T / Tc 
∆H V = ∆H V 0 ⋅  

 1 − T0 / Tc 
with: ∆Hv0 = 2.26⋅106 J/kg; T0 = 298 K; Tc = 647.3 K.

Density and viscosity (Empirical correlations, based on experimental data from


Newman, 1968).

X = w w ⋅ (a + b ⋅ T + c ⋅ T 2 + d / T) + (1 − w w ) ⋅ (e + f ⋅ T + g ⋅ T 2 + h / T) +
(1 − w w ) ⋅ w w ⋅ (i + j ⋅ T + k ⋅ T 2 + l / T)

X = ρmix X = ln(µmix⋅103)
(273<T<373, 0<xw<1) (333<T<363, 0.09<xw<0.5)
a 2114 -1.534⋅104
b -2.288 44.34
c 6.75⋅10-4 -4.274⋅10-2
d -1.478⋅105 1.769⋅106
e 1767 -1504
f -1.262 4.284
g 2.654⋅10-4 -4.096⋅10-3
h -46056 1.784⋅105
i -581.7 3.274⋅104
j 0.2971 -94.51
k 9.638⋅10-4 9.093⋅10-2
l 1.166⋅105 -3.783⋅106

Heat capacity, 273<T<373 K (Weast et al., 1982; Vargaftik et al., 1996).

Cp w = −1.355 ⋅10 5 + 1.165 ⋅10 7 / T + 629.0 ⋅ T − 1.261 ⋅ T 2 + 9.504 ⋅10 −4 ⋅ T 3


Cp g = −38368 + 3.712 ⋅10 6 / T + 158.3 ⋅ T − 0.2628 ⋅ T 2 + 1.680 ⋅10 − 4 ⋅ T 3
Cp mix = w w ⋅ Cp w + (1 − w w ) ⋅ Cp g
Hydrodynamics in a pervaporation membrane module 105

Heat conductivity, 273<T<623 K (Reid et al., 1988).

λ w = (−0.384 + 5.25 ⋅10 −3 ⋅ T − 6.37 ⋅10 −6 ⋅ T 2


0.235 ⋅ (1 − T / 726) 0.38
λg =
(T / 726)1 6
λ mix = w w ⋅ λ w + (1 − w w ) ⋅ λ g − 0.72 ⋅ w w ⋅ (1 − w w ) ⋅ (λ w − λ g )

Diffusion coefficients of water in mixture (Wilke et al., 1955; Vignes, 1966).

1 .173 ⋅ 10 − 16 ⋅ T ⋅ 2 .26 ⋅ M w
D gw o =
µ w ⋅ V g0 .6
1 .173 ⋅ 10 −16 ⋅ T ⋅ 1 .5 ⋅ M g
D wg o =
µ g ⋅ V w0.6
xw xg
Dw = D ⋅D /( d ln γ w / d ln x w )
gw 0 wg 0
2
 A 21 ⋅ (1 − x w ) 
γw = exp  A 21 ⋅ 
 A12 ⋅ x w + A 21 ⋅ (1 − x w ) 

with: Mw = 18 g/mol, Mg = 92.1 g/mol, Vg = 0.096 m3/kmol; Vw = 0.0756


m3/kmol, A12 = 0.2305, A21 = -0.6978.
MEASUREMENT OF 2D-TEMPERATURE
6
DISTRIBUTIONS IN A PERVAPORATION
MEMBRANE MODULE USING
ULTRASONIC COMPUTER TOMOGRAPHY

Abstract

Temperature polarization effects can considerably limit the


overall performance of pervaporation modules. This study shows the
measurement of temperatures in a pervaporation membrane module
using Ultrasonic Computer Tomography. With this technique,
complete 2D temperature distributions that occur in the module can be
measured within seconds. Temperature distributions were also
determined using Computational Fluid Dynamics. Both temperature
distributions are comparable, and show that in the tubular
pervaporation system studied, temperature segregation occurs. This
segregation is a result of the occurring hydrodynamics in the system
that is determined by mixed convection. Also, the measured water
fluxes through the membrane show to be dependent of forced and
natural convection.

This chapter is a slightly modified version of the publication:


Van der Gulik, G.J.S.; Wijers, J.G.; Keurentjes, J.T.F.; “Measurement of 2D-Temperature
Distributions in a Pervaporation Membrane Module using Ultrasonic Computer
Tomography and Comparison with Computational Fluid Dynamics Calculations”, J. Membr.
Sci., 2002, 204, 111-124.
108 Chapter 6

6.1 Introduction
During the past decades, pervaporation has emerged as a promising separation
technique. Using hydrophilic membranes, water can be removed from organic
mixtures, whereas using hydrophobic membranes, organic molecules can be
removed from aqueous solutions (Ho et al., 1992; Nunes et al., 2001). In both cases,
concentration and temperature polarization occur, which can limit the feasibility of
the separation.
Temperature polarization occurs as a result of the energy required for
evaporation of the compound removed. As the phase change occurs in the
membrane, the temperature of the membrane will be lower than the bulk liquid
temperature. In a heat transfer study, Rautenbach and Albrecht (1984, 1985) have
reported on these temperature effects during pervaporation by measuring
temperatures of feed, retentate, and permeate. More recent, heat transfer has been
studied by Karlsson and Trägårdh (1996), indicating that in some cases temperature
polarization is even rate limiting for the pervaporation process.
Using CFD (Computational Fluid Dynamics), we have recently (Van der Gulik
et al., 2001) investigated temperature and concentration polarization effects in a
ceramic pervaporation membrane reactor for resin production. In this process,
polarization effects are thought to limit the feasibility of pervaporation severely, as
temperature polarization will lead to strong viscosity gradients, thus hampering
mass transfer. We have observed that under laminar conditions, natural convection
induces secondary flow, which reduces the thickness of the hydrodynamic layers.
As a result, the polarization effects are effectively reduced, leading to a flux
increase up to 50% compared to cases without natural convection. Other secondary
flows are Dean vortices, that can be induced by curved membranes, which has
shown to be feasible for polymer (Winzeler et al., 1993; Moulin et al., 1999) and
ceramic mem-branes (Broussous et al., 1997).
In this paper, we report on experimentally determined temperature
distributions occurring during pervaporation. We have implemented the
experimental technique U-CT (Ultrasonic Computer Tomography) in the
pervaporation system, as previously used to study hydrodynamics under laminar
conditions (Van der Gulik et al., 2001). U-CT is based on the dependence of the
propagation velocity of ultrasound on the temperature of the medium. The
measurements provide line average temperatures that can be reconstructed to a 2D-
temperature distribution by using computerized tomography. The technique has
been used to measure temperature distributions in solids (Norton et al., 1984; Peyrin
et al., 1983), gases (Mizutani et al., 1997, 1999; Funakoshi, et al., 2000; Wright et
al., 1998), and liquids (Beckford, 1998; Basarab-Horwath et al., 1994). The U-CT
technique is non-invasive, and can be used for real time temperature measurements
in 2D, with fast computers nowadays. Background on the technique is given in the
experimental section.
Measurement of 2D-Temperature distributions using U-CT 109

This paper will demonstrate the successful implementation of the U-CT


technique for the measurement of temperature profiles during pervaporation.
Additionally, the measurements are used to validate CFD calculations.
110 Chapter 6

6.2 Experimental

6.2.1 The pervaporation system


The pervaporation system, as depicted in Figure 6.1A, was an annulus
configuration. The outer glass tube had a length of 1650 mm, and an inner diameter
of 40 mm. The wall thickness was 5 mm. The inner tube consisted of three parts: a
glass inlet pipe, the membrane, and a glass outlet pipe. The hydrophilic membrane
was supplied by ECN, The Netherlands, and consisted of a support tube, made of α-
Al2O3 on which an amorphous SiO2 top layer was applied, with an average pore
diameter smaller than 0.5 nm, and a thickness δ of 200 nm (Van Veen et al., 1999).
Porosity was estimated to be between 50 and 80%. The length of the membrane was
900 mm, with an outer and inner diameter of 14 and 8 mm, respectively. The inlet
and outlet pipes had lengths of 110 and 680 mm, respectively. The outlet pipe was
hollow, and provided the connection between the vacuum pump and the inside of
the hollow membrane. The complete inner tube could be translated in the axial
direction between two extreme positions as depicted in Figures 6.1A and 6.1B.
In this study, demineralized water was used as the working fluid, which was in
direct contact with the membrane. The water temperature in the storage tank was set
to 313 K. The inlet and outlet temperature of the liquid depended on operating
conditions. The ambient temperature TA was monitored continuously. U-CT
measurements were performed with and without a vacuum applied. In case of
vacuum, the permeate was condensed in a cold trap with liquid N2. The permeate
pressure was around 1 mbar, and the flux was determined gravimetrically.

Tank
(10 l.) Heater T = Temperature sensor

T Block with T Liquid recycle


transducers

A 855 mm
900 mm P

Inlet pipe Membrane Outlet pipe Vacuum pump


Module

B 390 mm
900 mm
Figure 6.1: Set-up of the pervaporation system during a U-CT-measurement at the end of
the membrane (Case A, with the block at x = 855 mm) and during a U-CT-measurement
halfway the membrane (Case B, with the block at x = 390 mm).
Measurement of 2D-Temperature distributions using U-CT 111

Qloss

Fin Fout
membrane
Tin Tout
Qin Qout
Fvap Qvap

Figure 6.2: Schematic representation of the pervaporation system indicating the energy
flows and the mass flows.

In Figure 6.2, the membrane module is represented schematically, in which F


represents volume flows [m3/s], T temperatures [K], and Q energy flows [J/s].
Without vacuum, there is no evaporation, so that Fvap = 0 and Qvap = 0. Then, the
following energy balance over the system can be set up:

Qin + Qout + Qloss = 0 [J/s] (6.1)

As the in- and outlet temperatures are known, the energy losses to the environment,
Qloss, can be calculated:

Qloss = ρc p Fin (Tout − Tin ) [J/s] (6.2)

with ρ the density [kg/m3] and specific heat capacity cp [J/(kg⋅K)] of water. When
applying vacuum, the energy balance becomes:

Qin + Qout + Qloss + Qvap = 0 [J/s] (6.3)

Qloss can be estimated, by assuming that the mass flow at the inlet is equal to the
mass flow at the outlet (i.e., |ρF|in≈|ρF|out>>|ρF|vap) and using the known inlet and
outlet temperatures:

Qloss = ρc p Fin (Tout − Tin ) − ρ∆H vap Fvap [J/s] (6.4)

with ∆Hvap the heat of vaporization for water [J/kg]. The energy flows Qloss and Qvap
have been calculated for each experiment, and are tabulated in Table 6.2. These
data are also required as input for the CFD calculations.
112 Chapter 6

2.2 U-CT Method


U-CT measurements are based on the measurement of time-of-flight (TOF) of
sound between a speaker and a microphone. The TOFL over a distance L is given
by:

1
TOFL = ∫L v( ρ )
dl [s] (6.5)

in which v(ρ) is the sound velocity, that depends on ρ. The velocity depends on
density, being a function of temperature and composition (and pressure when gases
are applied, which is currently not the case). As only water is used, the sound
velocity is only a function of temperature T (Suurmond, 1998):

v(ρ ) = v(T ) = 1402.336 + 5.03358 ⋅ T - 0.0579506 ⋅ T 2 + 3.31636 ⋅ 10 -4 ⋅ T 3


[m/s] (6.6)
- 1.45262 ⋅ 10 -6 ⋅ T 4 + 3.0449 ⋅ 10 -9 ⋅ T 5

with T in °C. From equation 6.6 the average line temperature can be derived. In
Figure 6.3, v(T) and dv/dT are shown as a function of temperature, in which dv/dT
determines the accuracy of the measurements. From the figure, it can be seen that
v(T) has a maximum at 346K (73°C). Obviously no measurements can be
performed when dv/dT = 0.
For speaker and microphone, in-house made transducers were applied in which
piezo-electric elements were mounted. The transducers were placed in an aluminum
block. The block was placed between the two glass tubes that together made up the
membrane module. The block could rotate independently from the glass tubes. The
T [°C]
0 20 40 60 80 100
1560 6

1540 5

1520 4
dv/dT [m/sK]

1500 3
v [m/s]

1480 2

1460 1

1440 0

1420 -1

1400 -2
273 293 313 333 353 373
T [K]

Figure 6.3: Sound velocity and its derivative, plotted against temperature.
Measurement of 2D-Temperature distributions using U-CT 113

Oscilloscope Module

b
e m ra

ne
M
Sound waves

Transducers
Data acquisition
system
Pulse generator

Figure 6.4: Required peripheral equipment for the U-CT-measurements.

fluid under study flowed through the block from one glass tube to the other, while
the transducers were in direct contact with the fluid. The complete construction is
shown in Figure 6.1A.
The peripheral equipment, needed for the U-CT measurements, is depicted in
Figure 6.4. The data acquisition system controlled and synchronized all actions. The
pulse generator (Synthesized Function Generator Yokogawa FG120 / 2 MHz)
generated a sound pulse for the speaker with a frequency of 1.2 MHz. The signal, as
collected by the microphone, was depicted on an oscilloscope (LeCroy 9314M
Quad 300MHz, 100 Ms/s, 50kbps/Channel), and was transferred to the PC that
processed the signal. A typical incoming signal is given in Figure 6.5A. Post-
processing started with a 40 degree low (1.1 MHz) and a high (1.3 MHz)
Cheybyshev bandpass filter, thus focussing on the speaker frequency of 1.2 MHz.
This resulted in the signal as depicted in Figure 6.5B. Then, a Hilbert transform was
applied that enveloped the absolute signal. The TOFL was chosen to be the time that
the signal level was equal to 20% of the maximum of the enveloped signal, as
depicted in Figure 6.5C. The obtained time was not the real TOFL, as a constant
error was introduced, which was removed by calibration. Wakai et al. (1990) have

A B 1.0 - C
10 µs 0.8 -
I /Imax
I [-] 0.6 -
0.4 -
0.2 -
0.0 -
0 time 0 time 0 time
tp
Figure 6.5: Signal processing in three steps. A) The incoming signal, obtained from the
oscilloscope. B) The signal after the Cheybyshev filter. C) The enveloped signal with the
dedicated time at 20% of the maximum in absolute signal.
114 Chapter 6

A B

5
Figure 6.6: A) Location of 8 transducers relative to the membrane with distances in mm. B)
A photograph representing the location of the 16 transducers in three dimensions.

studied different methods for determining the TOFL, which showed that, when
using the enveloped signal, the most accurate results could be obtained. Refraction
effects introduced an error that could not easily be removed. Estimating calculations
showed that the error was at least one order smaller than the errors generated by
noise.
In Figure 6.6A, the location of the transducers in the block, relative to the
membrane, is given in more detail. A 3D representation of the x, y, and z directions
is given in Figure 6.8. The transducers had a diameter of 5mm. In z-direction, 8
transducers were evenly distributed over 7 mm, whereby the top of transducer 8
matched the underside of the membrane. In x-direction, the transducers were evenly
distributed over 28 mm. Although the transducers differed in x-coordinate, for
reconstruction of the 2D temperature distribution it was presumed that the
transducers coincide at x = 14 mm. A 3D representation is given in the photograph
in Figure 6.6B.
Only the rays that are depicted in
Figure 6.4 were used. It was also possible
to use oblique rays up to two transducers
to the left or to the right, but these rays
were not used, as the density of the given
rays was sufficiently high for recon-
struction purposes. By rotating the block
independently from the glass tubes and
the membrane, the straight rays covered
the complete section. It was chosen to
perform measurements after each rotation
over 30°. So, for completion of the total
360°, 12 rotations were needed, providing
a ray distribution as depicted in Figure
6.7.
Calibration of the system was Figure 6.7: Ray distribution after rotation
required, as the exact distance between of the block over 360° in 12 steps.
Measurement of 2D-Temperature distributions using U-CT 115

transducers was unknown. Therefore, TOFL measurements at a constant and


homogeneous temperature were performed. To obtain this situation, the flow and
heating were turned off, and the apparatus was maintained at rest for 24 hours.
During calibration, the temperature of the liquid and the environment were
monitored continuously within ± 0.1 K. Using TOFL and the measured temperature,
the exact distances between the transducers could be calculated using equations 6.5
and 6.6.
The complete inside tube (glass tube – membrane – glass tube) could be
moved in axial position, relative to the U-CT recorder and the outside tube, thus
allowing to perform U-CT measurements at several positions along the membrane.
The U-CT measurements, as presented in Figure 6.1, were performed at the end of
the membrane (x = 855 mm, Figure 6.1A) and approximately halfway the
membrane (x = 390 mm, Figure 6.1B). The experimental set-up was used with and
without vacuum. Applying two flow rates, 300 and 500 mL/min, 6 situations were
studied as summarized in Table 6.1. For each TOFL measurement, the average of 10
consecutive measurements was taken in order to reduce noise contribution.
Standard deviations will be presented as a result. A TOFL measurement then took
about 10 seconds. With 8 rays, the total measurement time of one angle took 80
seconds. After rotation of the block over 30°, the system set-up was given 300
seconds to stabilize before the next TOFL measurement was performed.
Consequently, the measurement of a complete section took 360°/30°∗(300+80) =
4560 seconds.

Table 6.1: Conditions of experiments 1 to 6.


Flow [mL/min] Condition Fig. 1 x-coordinate [mm] Experiment no.
No vacuum A 855 1
300 A 855 2
Vacuum
B 390 3
No vacuum A 855 4
500 A 855 5
Vacuum
B 390 6

For reconstruction of the 2D-temperature distribution it was assumed that all


measurements were performed simultaneously. This implies, that we assumed the
complete system to be at the equilibrium state during the complete U-CT
measurement. Based on the inlet, outlet and ambient temperatures this assumption
could hold as these remained constant within 0.1K.

2.3. Reconstruction
The experiments provided 96 rays, that form the basis for reconstruction of the
temperature distribution. The experimental set-up was symmetrical around the
vertical centerline, which allowed for averaging of symmetrical measurements.
The reconstruction method used, was filtered back-projection. The temperature
distribution that had to be reconstructed, is infinitely-dimensional, and there were
116 Chapter 6

only 96 measurements, indicating that the problem was severely underdetermined.


Therefore, regularization techniques were needed to obtain numerical stability. A
suitable method was a collocation method in which the Moore-Penrose generalized
inverse of a large matrix was computed (Suurmond, 1998; Golub et al., 1983).
A temperature distribution with circular shape had to be constructed. This
shape was built up in a rectangular grid with 7854 cells. The temperatures in these
cells were restricted by using linear independent basis pictures pi. In this case, the
number of basis pictures was equal to the number of measurements, i.e., 96. The
anticipated temperature distribution was built up by these 96 pictures. The basis
pictures were designed to take into account the fact that the speed of sound could
not vary too much in space. One choice of basis pictures that was in agreement with
this, were smooth ‘hills’, as modeled by negative exponential functions:

(
p i ( x ) = exp − λ ⋅ || x − x i ||2 ) (6.7)

where the xi were positioned in a regular grid, that covered the reconstruction region
around position x. i is an index, ranging from 1 to the number of basis pictures (=
96). λ is the damping factor of the basis functions. The use of these basis pictures pi
provided much better results than using the original ART methods (Herman, 1980).
Norton et al. (1984) showed that Bessel-functions are also good candidates for
obtaining smooth distributions.
From the anticipated temperature distribution, as built up by the basis pictures,
a matrix R could be constructed, consisting of the back-projection of discrete ray
transforms of the basis pictures:

( )
R = Ρ p j (x i ,θ i ) ij (6.8)

in which P was a linear operator, representing the back projection pj at position


(xi,θi). j Is an index ranging from 1 to the number of rays (= 96). Now the problem
could be represented as a matrix equation: given measured data y, find a vector ξ
such that Rξ best approximates the measured data y:

Rξ = y (6.9)

In other words, find the solution vector ξ by minimizing:

|| Rξ − y || (6.10)

for which SVD (Singular Value Decomposition) was used, including the Moore-
Penrose generalized inverse matrix (Suurmond, 1998; Golub et al., 1983)..
The solution ξ could be spoiled, because of small singular values, that arose in
the system matrix of the tomography problem. As an efficient approach to
overcome this problem, a filter was used that modified the singular values, such that
the resulting matrix did not have any near-zero singular values. A soft Tichonov
Measurement of 2D-Temperature distributions using U-CT 117

regularization filter was applied with filter parameter η. Using this parameter, the
solution ξ was computed by minimizing the following function:

|| Rξ − y || 2 + η || ξ ||2 (6.11)

In this way, a compromise was obtained between accuracy and smoothness:


large values of η would result in smoother reconstructions, while smaller values of
η would cause Rξ to follow the data y more closely. In this paper a value of 0.001
was used. The choice was based on the maximum values obtained in the legend of
Figure 6.13A, i.e., 311.46 and 304.16K, respectively. These values had to match the
extremes in vector y. This matching introduced an error, as the ‘real’ maximum was
always higher or equal to the maximum ray-averaged temperature. Smaller values
for η could compensate for this, but it was found that smaller values caused
spurious high-frequency oscillations in the obtained temperature distribution.
The temperature distributions were reconstructed for a complete circle, as
surrounded by the glass wall. The circle included the inner area that is covered by
the membrane, although no U-CT measurements could be performed in that area.
Reconstruction was performed here, because it was impossible to implement a
boundary condition in the reconstruction software that describes the temperature on
the membrane. Also, no values were available beforehand to set up a boundary
condition.

2.4. Computational Fluid Dynamics


Flow simulations of the
pervaporation system were Outlet pipe (680 mm)
performed to obtain a refer-
ence for the U-CT experi-
ments. CFX 4.2 was used for Membrane (890 mm)
this purpose, which is a finite
volume fluid dynamics pack- Inlet pipe 14 mm
age (AEA Technology, Har- (110 mm)
well, UK). The software was
installed on a Silicon Graph-
ics Origin 200 workstation z
(IRIX 6.4). To study the
contribution of natural con- x 40 mm
vection in the horizontal set-
up, 3D simulations were per- y
formed on a half-tube. The
Cartesian grid (74×30 ×35),
is depicted in Figure 6.8. A
symmetry plane was applied Figure 6.8: Representation of the grid (74×30×35) of
in the x-z-plane, as the the pervaporation system in position A, as described in
problem was considered Figure 6.1. From the inner tube, only the surface grid
of the membrane is depicted.
118 Chapter 6

symmetrical, thus significantly reducing computation time. To demonstrate the


independence of the solution from the grid, the simulations were repeated with 1.5
times the number of cells, without a noticeable change in solution.
The membrane was simulated as a wall patch, on which a sink term was
defined that represented the water flux at arbitrary numeric cell i (φw,i) as follows:

φ w ,i = C ⋅ Pw∗,i [kg/(m2⋅s)] (6.12)

with C as a constant and P*w,i as the feed-phase vapor pressure of water at the
membrane. Required data for the equation of P*w,i were obtained from Reid et al.
(1988). C was chosen such, that the sum of φw,i was equal to the experimentally
observed flux φw, thus:

∑φ
i =1
w ,i Ai = 3600φ w Am [kg/s] (6.13)

with Ai the membrane area of numeric cell i, and Am the total membrane area. The
amount of energy required for evaporation was accounted for in a sink term as
follows:

φ h ,i = φ w ,i ⋅ ∆H vap [J/(m2⋅s)] (6.14)

with ∆Hvap in J/kg. Viscosity, heat capacity, heat of vaporization, and conductivity
of water were assumed to be temperature-independent.
Heat losses to the environment were also taken into account, because
experiments showed that losses were significant. On the outside wall a combined
Dirichlet-Neumann boundary condition was defined as follows:

Qi = hAi (T A − Ti ) [J/s] (6.15)

in which Qi was the heat loss in a single numeric cell, h the lumped heat transfer
coefficient, Ai the cell area, TA the ambient temperature, and Ti the cell temperature.
In a converged solution, the experimentally determined total heat loss Qloss, was set
equal to the sum of the heat losses in all outer cells:

n n

∑ Q = ∑ hA (T
i =1
i
i =1
i A − Ti ) = Qloss [J/s] (6.16)

h was kept constant for the complete tube. The ambient temperature TA was known
from measurements.
Measurement of 2D-Temperature distributions using U-CT 119

6.3 Results and discussion

6.3.1 Flux measurements


The relevant experimental settings and the measured and calculated quantities
during the flux measurements are given in Table 6.2. The calculated Reynolds
number indicates that conditions are laminar. The measured inlet temperature at 300
mL/min is 0.4 K lower than at 500 mL/min, as losses from the feed tube towards
the environment are larger at lower flow speed. The outlet temperature is lower than
the inlet temperature, also as a result of energy losses towards the environment. The
energy loss over the membrane module, Qloss, can be calculated from the difference
in inlet and outlet temperature, and Fin, using equation 6.2.

Table 6.2: Settings, measured quantities, and calculated energy flows for the 6 experiments
performed.
Set Measured Calculated

Exp Fin vsup Re


Vacuum
Tin Pvac φw Tout Qin-Qout Qvap -Qloss
No. [mL/min] [mm/s] [-] [K] [mbar] [kg/m2⋅h] [K] [W] [W] [W]
1 No - - 309.2 50.4 - 50.4
300 4.53 118 311.6
2&3 Yes 1 2.22 307.0 96.6 53.7 42.9
4 No - - 310.4 57.0 - 57.0
500 7.56 196 312.0
5&6 Yes 2 2.77 308.6 119 63.0 56.0

When the vacuum is applied, water is evaporated at the membrane. The


obtained water flux, φw, appears hardly to be affected by the superficial velocity
vsup, indicating that hydrodynamics are mainly determined by natural convection.
The fluxes are in good agreement with results reported previously (Van der Gulik et
al., 2001). The permeate pressure at 500 mL/min was found to be twice as high as
compared to 300 mL/min. Based on the results presented by Rautenbach and
Albrecht (1980), the permeate pressure should be less than 25% of the vapor
pressure to eliminate the influence of the permeate pressure on the water flux.
Meanwhile, the vapor pressure is determined by the temperature of the membrane,
which is unknown. However, assuming a membrane temperature equal to 293 K,
the vapor pressure is calculated to be 23 mbar. Consequently, no effect of the
permeate pressure is expected.
The losses towards the environment with the vacuum applied, appear to be
lower than the losses without vacuum applied. This is because the average
temperature in the module is lower, resulting in a lower driving force for heat
transfer towards the environment.
120 Chapter 6

1
1

z z
y y
g g
2

A) Experiment 1 B) Experiment 2

Figure 6.9: Secondary flow pattern predicted by CFD for experiment 1 (no vacuum, Fin =
300 mL/min, x = 855mm) and experiment 2 (vacuum, Fin = 300 mL/min, x = 855mm).

6.3.2 CFD simulations


CFD simulations have been performed on the 6 cases, that are summarized in
Table 6.1. Using the measured water and energy fluxes in the boundary conditions
as described in equations 6.12-6.16, secondary flow patterns as given in Figure 6.9
are obtained. In Figure 6.9A, the secondary flow pattern is given for 300 mL/min at
x = 855 mm without vacuum. Due to heat losses towards the environment, the
outside wall temperature is lower than the bulk temperature. Consequently, the
density is higher, inducing a downward velocity at the wall, and an upward velocity
in the bulk near the membrane, creating one large cell at the top of the annulus, as
marked with ‘1’. The temperature distribution at the same x-coordinate (x = 855
mm) is given in Figure 6.10A. The figures show that fluid with low temperature is
expected to be collected at the bottom of the annulus.
When vacuum is applied, the temperature at the membrane also drops.
Although the energy flux as a result of the evaporation is comparable to the losses
to environment (compare Qvap and Qloss in Table 6.2), the temperature drop at the
Measurement of 2D-Temperature distributions using U-CT 121

membrane is larger, because the membrane area is substantially smaller than the
outside wall. As a result, the maximum temperature difference over the annulus is 4
K without vacuum (Figure 6.10A), while a maximum temperature difference of 9 K
is found when the vacuum is applied (Figure 6.10C). Because the temperature
differences are larger, the density differences are also larger. Consequently, the
induced secondary flow around the membrane is stronger than at the outside wall.
As both induced secondary flows are pointed downward, they suppress each other
in the annulus. Therefore, the cell at the outside wall in Figure 6.9A is reduced in
Figure 6.9B (marked with ‘1’), but a second cell is formed at the bottom (marked
with ‘2’). Both cells provide local mixing at the top and the bottom, respectively,
but overall mixing is poor. As a result, a large temperature difference is obtained
between the top and the bottom of the annulus as can be seen in Figures 10C and
10D.

6.3.3 U-CT measurements


In order to determine the calculated temperature profiles experimentally,
Ultrasonic Computer Tomography measurements have been performed. Firstly, the
accuracy of the U-CT measurements will be discussed. Secondly, the performance
of the reconstruction software will be demonstrated, by using average ray
temperatures extracted from a CFD calculation. Finally, the experimental U-CT
measurements are reconstructed into a 2D-temperature distribution, demonstrating
the capabilities of the technique.
To study the noise level of our equipment, standard deviations in temperature
have been determined. These deviations follow from 10 consecutive calibration
measurements, in which the distance between the 8 speaker-microphone
combinations has been determined at constant temperature. The determined
deviations in distance can be converted into deviations in temperature using
equations 6.5 and 6.6. Using equation 6.6, the deviations are extrapolated to the
higher temperatures that are relevant in our experiments. The deviations are given
in Figure 6.11 for all 8 speaker-microphone combinations. At 313 K, the deviations
are between 0.3 and 0.7 K, and at 298 K between 0.2 and 0.4 K, respectively.
Therefore, using water, lower noise levels can be obtained at low temperatures. The
deviations do not correlate with the distance between speaker and microphone and
are thought to be introduced by the transducers themselves. The deviations define
the lower limit in measurable temperature differences.
To test the performance of the reconstruction procedure and the value for the
parameter η of the Tichonov regularization filter, a reconstruction has been made of
the temperature distribution in Figure 6.10C that have been calculated using CFD.
This CFD calculation corresponds to experiment 2, at 300 mL/min, with the
vacuum applied (see Table 6.1 and 2 for details). For this purpose, the average
temperatures over imaginary rays in the temperature distribution in Figure 6.10C
have been determined. These average temperatures are presented in Figure 6.12A as
a function off the rotation angle. At 0°, the rays cover the bottom of the module,
while at 180° the rays cover the top of the module. These ray temperatures are
reconstructed to a temperature distribution again that is given in Figure 6.13A. The
122 Chapter 6

reconstruction performs according to the correspondence with Figure 6.10, although


some finer structures in the temperature distributions near the membrane are missed
due to the absence of rays near the membrane.
Measurement of 2D-Temperature distributions using U-CT 123

Temperature [K] A B C D

Figure 6.10: Temperature distributions predicted by CFD for A) Experiment 1 (no vacuum,
Fin = 300 mL/min, x = 855mm), B) Experiment 4 (no vacuum, Fin = 500 mL/min, x =
855mm), C) Experiment 2 (vacuum, Fin = 300 mL/min, x = 855mm), and D) Experiment 5
(vacuum, Fin = 500 mL/min, x = 855mm).
0.7

1 2
3 4
0.6
5 6
7 8

0.5
[K]
σT

0.4

0.3

0.2
295 300 305 310 315
T [K]

Figure 6.11: Standard deviation in the temperature measurements at 25 °C for the 8


transducer combinations and extrapolated towards higher temperatures. The numbers from
one to eight refer to the transducer numbers, that are given in Figure 6.6.
124 Chapter 6
312 312

310
A 310
B
308 308

Temperature [K]
Temperature [K]

1 1
306 306
2 2
3 3
304 4 304 4
5 5
6 6
302 302
7 7
8 8
300 300
0 90 180 270 360 0 90 180 270 360
Position [°] Position [°]

Figure 6.12: A) Ray temperatures for a CFD calculation, set-up according to experiment 2.
B) Measured ray temperatures for experiment 2 (vacuum, Fin = 300 mL/min, x = 855mm).

A T [K] B T [K]
311.46 311.00
311.15
a 309.15
310.15
307.15
309.15
305.15
308.15

303.15
307.15

301.15
306.15

299.15
305.15

297.15
304.16 296.95

Figure 6.13: A) Reconstruction of the ray temperatures in Figure 6.12A, derived from the
CFD-calculation in Figure 6.10C. B) Reconstruction of measured ray temperatures in Figure
6.12B. Both figures correspond with experiment 2 (vacuum, Fin = 300 mL/min, x = 855mm).

A T [K]
309.13
B T [K]
309.06

308.15
a 307.15 307.15
306.15
305.15
305.15

303.15 304.15
303.15
301.15 302.15
301.15
299.15
300.15
299.15
297.15
296.94 298.48

Figure 6.14: Reconstructed images of the ray temperatures at x = 390 mm in A) experiment 3


(vacuum, Fin = 300 mL/min) and B) experiment 6 (vacuum, Fin = 500 mL/min).
Measurement of 2D-Temperature distributions using U-CT 125

The highest and the lowest values in the legends of Figures 10C and 13A are used
to check the previously determined parameter η. Using η = 0.001 the match appears
to be good, although the highest and the lowest values in the legend of the
reconstructed image in Figure 6.13A are higher than in the original image in Figure
6.10C. This can be attributed to the pronounced extremes at the bottom, and at the
top of the annulus in the CFD calculation.
In Figure 6.12B, the averaged temperatures of U-CT measurements are given
for experiment 2 (300 ml/min, vacuum, x = 855 mm). The experimental conditions
are equal to the CFD calculation in Figure 6.12A. At 0°, the rays measure a lower
temperature than at 180°, which corresponds to the CFD calculations. Overall, the
measured temperatures are lower than in the CFD calculation. The values of the
CFD distribution are thought to represent the experiments better, as the calculations
were fitted on the macroscopic values that are given in Table 6.2.
Figure 6.13B shows the reconstruction of the temperatures in Figure 6.12B.
Figure 6.13B clearly shows a region with a low temperature below the membrane,
and a region with a high temperature above the membrane. This segregation
corresponds to the CFD calculations in Figure 6.11C and 13A. The temperatures at
the top of the annulus (311 K) and at the bottom glass wall (304 K), correspond
very well. The main difference between the reconstructed CFD calculation, and
reconstructed U-CT measurement is the temperature just below the membrane. The
CFD calculation in Figure 6.10C predicts 302 K, while the reconstruction in Figure
6.13B yields 297 K. We assume that reconstruction shows a too low temperature.
The temperature is the outcome of a balance between the ray temperatures in the
reconstruction, and the ray temperature of the measurement according to equation
6.11. Probably the value η = 0.001 for the Tichonov-filter is too strict, leading to a
local unbalance. On the other hand, the value of η = 0.001 seems adequate, as the
too high temperatures of ray 7 in Figure 6.12B do not return in the reconstruction as
they are distributed over the image by the Tichonov regularization filter.
The reconstructions of U-CT experiments 3 and 6 are given in Figures 6.14A
and 6.14B (x = 390 mm, vacuum, 300, and 500 mL/min, respectively). As expected,
the temperature distributions are similar, but overall, the temperatures at 300
mL/min are lower than at 500 mL/min. Compared to the CFD calculation, the
temperatures are about 2 K lower. At 300 mL/min, the temperatures at the bottom
of the module are lower than at 500 mL/min, which is as expected, since at 300
mL/min natural convection determines hydrodynamics to a larger extent than at 500
mL/min (Van der Gulik et al., 2001). Overall, the temperature distributions are
adequate. The distributions, as well as the maximum temperatures, correspond with
the CFD calculations. Also the occurrence of natural convection, as predicted by the
CFD calculation (Van der Gulik et al., 2001), has been shown to exist.
Reconstruction of the ray temperatures of the experiments without vacuum
(experiments 1 and 4), did not result in realistic temperature distributions. The
occurring temperature differences in the module are of the same order of magnitude
as the standard deviations that are given in Figure 6.8. Thus, the noise level has
been too high. Consequently, to be able to perform U-CT measurements without the
vacuum, the technique has to be optimized further, with emphasis on noise
reduction and the determination of TOFL. Wakai et al. (1990) have concluded that
126 Chapter 6

the enveloped signal after Hilbert transformation provides the best results for
determining the TOFL. Analogously, we have chosen the TOFL to be the time that
the signal is equal to 20% of the maximum of the enveloped signal. However, a
detailed study of the signals has shown that declination can occur when the
maximum decreases. Therefore, future developments should include an improved
determination method, probably using a sweep signal that shifts the speaker
frequency from 1.2 to 1.4 MHz. After applying the Cheybyshev-filter on the
microphone signal, the speaker frequency shift will be visible. Then by a unique
overlap of both signals, the TOFL is expected to be determined straightforward.
Measurement of 2D-Temperature distributions using U-CT 127

6.4 Concluding remarks


This study illustrates the implementation of U-CT measurements for the
measurement of temperature distributions in pervaporation modules. The measured
temperature distributions agree well with the CFD calculations. The temperature
distributions prove the occurrence of natural convection in the pervaporation
membrane module, as previously predicted by CFD calculations.
Improving the experimental technique by reducing the noise level can broaden
the application to larger and more complex systems, revealing temperature
distributions inside pervaporation modules with multiple tubes. Also, because in
principle the U-CT technique is based on density differences, it should be possible
to measure composition differences. Moreover, the measurements can be performed
in various types of equipment, provided density differences are present over the
sampling plane. Reducing measurement times can extent the application to dynamic
processes, even with prospects in control systems.

Nomenclature
A area m2
C constant -
cp specific heat J/(kg⋅K)
F volume flow m3/s
h overall heat transfer coefficient W/(m2⋅K)
n number of numeric cells -
P pressure Pa
P* vapor pressure Pa
pi basis picture -
R matrix of back-projected ray transforms
Re Reynolds number -
Q energy flow J/s
T temperature K
TOFL time of flight over length L s
v velocity m/s
x vector position of the basis picture
xi vector position of points around x

Greek

δ thickness of the top layer on the membrane m


∆Hvap heat of vaporization J/kg
φw water flux kg/(m2⋅h)
φh energy flux J/(m2⋅h)
η Tichonov Regularization filter parameter -
λ damping factor of the basis picture
ξ solution vector
ρ density kg/m3
128 Chapter 6

Subscripts

A ambient
h heat
I ith cell
in inlet
loss losses to environment
out outlet
sup superficial velocity
vap vaporization
vac vacuum
w water

References
Basarab-Horwath, I.; Dorozhevets, M.M.; “Measurement of the temperature distribution in
fluids using ultrasonic tomography”, IEEE Ultrasonics Symposium, Cannes, 1994, 1891-
1894.
Beckford, P.; Höfelmann, G.; Luck, H.O.; Franken, D.; “Temperature and velocity flow
fields measurements using ultrasonic computer tomography”, Heat Mass Transfer, 1998,
33, 395-403.
Broussous, L.; Ruiz, J.C.; Larbot, A.; Cot, L.; “Stamped ceramic porous tubes for tangential
filtration”, Proc. Euromembrane '97, Twente, The Netherlands, 1997, 394-396.
Funakoshi, A.; Mizutani, K.; Nagai, K.; Harakawa, K.; Yokoyama, T.; “Temperature
distribution in circular space reconstructed from sampling data at unequal intervals in
small numbers using acoustic computerized tomography”, Jpn. J. Appl. Phys., 2000, 39,
3107-3111.
Golub, G.H.; Van Loan, C.F.; “Matrix computations”, North Oxford Academic, 1983.
Ho, W.S.W.; Sirkar, K.K. (Eds.); “Membrane Handbook”, Chapman & Hall, New York,
1992.
Herman, G.T.; “Image reconstruction from projections. The fundamentals of computerized
tomography”, Academic Press, Inc., 1980.
Karlsson, H.O.E.; Trägårdh, G.; “Heat transfer in pervaporation”, J. Membrane Sci., 1996,
119, 295-306.
Mizutani, K.; Nishizaki, K.; Nagai, K.; Harakawa, K.; “Measurement of temperature
distribution in space using ultrasound computerized tomography”, Jpn. J. Appl. Phys.,
1997, 36, 3176-3177.
Mizutani, K.; Funakoshi, A.; Nagai, K.; Harakawa, K.; “Acoustic measurement of
temperature distribution in a room using a small number of transducers”, Jpn. J. Appl.
Phys., 1999, 38, 3131-3134.
Moulin, P.; Manno, P.; Rouch, J.C.; Serra, C.; Clifton, M.J.; Aptel, P.; ”Flux improvement
by Dean vortices: ultrafiltration of colloidal suspensions and macromolecular solutions”,
J. Membrane Sci., 1999, 156, 109-130.
Norton, S.J.; Testardi, L.R.; Wadley, H.N.G.; “Reconstruction internal temperature
distributions from ultrasonic time-of-flight tomography and dimensional resonance
measurements”, J. Res. Natl. Bur. Stand., 1984, 89(1), 65-74.
Nunes, S.P.; Peinemann, K.-V. (Eds.); “Membrane Technology in the Chemical Industry”,
Wiley-VCH, Weinheim, 2001.
Measurement of 2D-Temperature distributions using U-CT 129

Peyrin, F.; Odet, C.; Fleischmann, P.; Perdrix, M.; “Mapping of internal material
temperature with ultrasonic computed tomography”, Ultrason. Imag., Conference
Proceeding, Halifax, 1983, 31-36.
Rautenbach, R.; Albrecht, R.; “Separation of organic binary mixtures by pervaporation”, J.
Membrane Sci., 1980, 7, 203.
Rautenbach, R.; Albrecht, R.; “On the behavior of asymmetric membranes in pervapora-
tion”, J. Membrane Sci., 1984, 19, 1-22.
Rautenbach, R.; Albrecht, R.; “The separation potential of pervaporation. Part 2. Process
design and economics”, J. Membrane Sci., 1985, 25, 25-54.
Reid, R.C.; Prausnitz, J.M.; Poling, B.E.; “Properties of gases and liquids”, 4th ed. Mc Graw-
Hill Book Co., 1988.
Suurmond, R.T.; “Tomographic reconstruction of temperature distributions from acoustic
measurements”, Postgraduate report, Stan Ackermans Institute, Eindhoven, 1998, ISBN
90-5282-873-3.
Van der Gulik, G.J.S.; Janssen, R.E.G.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodynamics in
a ceramic pervaporation membrane reactor for resin production”, Chem. Eng. Sci., 2001,
56, 371-379.
Van Veen, H.M.; Van Delft, Y.C.; Engelen, C.W.R.; Pex, P.P.A.C.; “Dewatering of organics
by pervaporation with silica membranes”, Proc. Euromembrane '99, Leuven, Belgium,
1999, 209.
Wakai, K.; Shimizu, S.; Ishikawa, S.; “Instantaneous measurement of two-dimensional
distribution of temperature of water by means of ultrasonic-CT”, Proc. of 2nd KSME-
JSME Fluids Engineering Conf., Seoul, 1990, Vol.1, 299-304.
Winzeler, H.B.; Belfort, G.; “Enhanced performance for pressure-driven membrane
processes: the argument for fluid instabilities”, J. Membrane Sci., 1993, 80, 35-47.
Wright, W.M.D.; Schindel, D.W.; Hutchins, D.A.; Carpenter P.W.; Jansen, D.P.; “Ultrasonic
tomographic imaging of temperature and flow fields in gases using air-coupled
capacitance transducers”, J. Acoust. Soc. Am., 1998 104(6), 3446-3455.
FUTURE PERSPECTIVES FOR
7
PROCESS ENGINEERING OF
POLYCONDENSATION REACTIONS

Abstract

In this chapter the most important results described in the


previous chapters will be summarized. Some additional aspects will be
discussed for describing the hydrodynamics in the two types of reactor
investigated here. Additionally, some issues will be discussed for the
design and development of processes for polycondensation reactions
that have not been described in the previous chapters.
132 Chapter 7

7.1 Hydrodynamics in polycondensation reactors

7.1.1 Drais horizontal stirred tank reactor


A large number of flow parameters of the Drais reactor have been established
experimentally. This has been done for turbulent as well as laminar conditions,
which are the occurring flow regimes during the polymerization process. The flow
parameters include the fluid velocities, turbulence parameters, energy dissipation
rates, mixing patterns, mixing times, and power numbers. The results have been
described in Chapters 2 and 3. From these results it can be concluded that mixing in
the Drais reactor is relatively good under laminar conditions due to the occurrence
of chaotic patterns upon mixing, resulting in short macro-mixing times. The
dimensionless macro-mixing Ntm appears to increase upon liquid volume. It was
shown that the hydrodynamics on large and small scale are different. For the
polycondensation reaction investigated, the Drais reactor volume cannot be made
too large, as the macro-mixing time tm will increase to undesirable values.
Increasing the impeller frequency can keep tm constant upon scale-up but this may
lead to unrealistic power requirements.
To increase the understanding of the flow phenomena in the Drais reactor, the
hydrodynamics have been studied numerically using computational fluid dynamics
(CFD) as described in Chapter 4. In these CFD calculations a distinction has been
made between fluid flow and mixing. The performance of two turbulence models
has been tested for describing the turbulence parameters occurring in the Drais
reactor. The simple (and computationally friendly) k-ε model and the complex
Differential Stress Model (DSM) predict fluid flow in the Drais reactor equally
well. However, mixing can only be predicted correctly using the DSM. This implies
that in CFD the choice of an appropriate turbulence model is of key importance for
the correct description of scalar mixing. Moreover, if the choice for a turbulence
model is only based on the comparison of mean flow parameters, the description of
scalar mixing may fail completely.

7.1.2 Membrane reactor


As a result of a combination of temperature polarization and concentration
polarization, the latter leading to an increased reaction rate near the membrane
surface, a strong viscosity polarization has initially been expected to be present in
the pervaporation membrane reactor. With the CFD simulations described in
Chapters 5 and 6, buoyancy forces have shown to be responsible for the occurrence
of a secondary flow, which refreshes the membrane surface and leads to increased
water fluxes through the membrane. Horizontal configurations showed higher
fluxes than vertical configurations. In horizontal configurations, a water flux
increase of 50% was observed due to the secondary flow.
In order to be able to measure temperature distributions, an Ultrasonic
Computer Tomography (U-CT) methodology has been developed successfully. The
Future perspectives for process engineering of polycondensation reactions 133

method has been implemented in the horizontal membrane module as described in


Chapter 6. Using U-CT, it has been shown that the temperature distributions
predicted by means of CFD were indeed present in the module.

4 0,6 1
Ut(r=0.28) ut2(r=0.28)

ut 2 [m2/s 2]
3,9 Ut(r=0.25) 0,5 ut2(r=0.25) Turbulent fluctuations
3,8 Ua(r=0.29) 0,4 ua2(r=0.29) ui2 = u i2,t + u i2, p
3,7 Ua(r=0.28)
0,1
ua2(r=0.28)
0,3
Ua(r=0.25) ua2(r=0.25)
3,6

Ua [m/s]
0,2
Ut [m/s]

3,5
0,1
3,4
0,0 0,01
3,3
3,2 -0,1

3,1 x L = 0.1 m -0,2 x L = 0.1 m

3 -0,3 0,001
0 0,2 0,4 0,6 0,8 1 0 0,2 0,4 0,6 0,8 1
x [-] x [-]

Figure 7.1A-B: A) Tangential and axial velocities as a function of fill ratio.


B) Axial and tangential fluctuations as a function of fill ratio.

1 eq NaOH

1.05 eq ECA 1.05 eq HCl

Figure 7.2: The equivalencies in ECA, HCl and NaOH for chemical
experiments to study the mixing performance in the Drais reactor.
134 Chapter 7

7.2 Design and scale-up of the Drais reactor


In the previous chapters several aspects of the hydrodynamics in the Drais
reactor have been described. Empirical formulas are provided that can be used to
predict macro-mixing times and power consumption on various scales. Also,
detailed data on fluid velocities and turbulence quantities have been determined
using LDA and CFD. In this section, additional LDA data are provided that extend
the applicability of the findings from the LDA measurements to fill ratios below
100%. To allow for a verification of the future implementation of chemical
reactions in CFD, chemical reactions have been performed experimentally.

7.2.1 LDA measurements at intermediate fill ratios


The LDA measurements as described in Chapter 4 have been performed in
completely filled reactors. However, in practice the reactor is used at fill ratios of
40 to 50%. It is interesting to know to what extent the presented data are applicable
to lower fill ratios. Therefore, additional LDA measurements have been performed
at intermediate fill ratios, providing axial and tangential velocities and turbulence
properties. The experimental conditions are similar to the conditions described in
Chapter 4. Measurements have only been performed at the axial coordinate xL =
0.1, and radii r = 0.29, 0.28, and 0.25, respectively.
The results are summarized in Figures 7.1A and 7.1B in which the averaged
velocities and the turbulent fluctuations, respectively, are given. Both figures show
that upon decreasing the fill ratio (x = 1 → 0), both the local velocities and their
fluctuating components remain approximately constant and only deviate below x =
0.5. It is possible that the measurements at fill ratios below x = 0.4 are somewhat
inaccurate because the interface between the liquid ring and gaseous core might
interfere with the measurement volume. Nevertheless, it can reasonably be assumed
that the data as presented in Chapter 4 are applicable down to fill ratios of 40%.
Apparently, for the liquid near the cylindrical wall it does not matter whether a
gaseous core or a liquid that rotates as a solid body is present.

7.2.2 Combining reaction kinetics and mixing


The polymerization reaction that takes place in the Drais reactor has been
classified as a parallel-competitive reaction system (R.1, Chapter 1). A similar
model reaction system can been used to relate reaction selectivity and micro mixing
(Bourne et al., 1994; Schoenmakers, 1998; Verschuren, 2001). The reaction system
is usually referred to as the 3rd Bourne reaction and comprises the following two
reactions:

NaOH + HCl → H2O + NaCl (R.1)


NaOH + ClCH2COOCH2CH3 → C2H5OH + ClCH2COO-Na+ (R.2)
Future perspectives for process engineering of polycondensation reactions 135

12 6

Amount NaOH used for R.2 [%] 10 5

Macro mixing time [s]


8 4

6 3

4 2
Chemical experiments
2 tm PLIF (Chapter 2) L/D = 1.1 1
x = 40%
tm RTD (Chapter 3)
0 0
0 2 4 6 8 10 12 14 16
Impeller speed N [Hz]

Figure 7.3: Percentage NaOH used for Reaction 2 and macro-mixing


time tm against impeller frequency in reactor11 at a fill ratio of 40%.

12 6
Amount NaOH used for R.2 [%]

10 5
Macro mixing time [s]

8 4

6 3

4 2

Chemical experiments L/D = 1.1


2 tm PLIF (Chapter 2)
1
N = 10
tm RTD (Chapter 3)
0 0
0 0.2 0.4 0.6 0.8 1
Fill ratio [-]

Figure 7.4: Percentage NaOH used for Reaction 2 and macro


mixing time tm against fill ratio in reactor11 at N = 10 Hz.

Reaction R.1 is an instantaneous reaction with a reaction coefficient of k1 =


1.3⋅1011 L/(mol⋅s). Reaction R.2 is relatively slow with k2 = 30.4 L/(mol⋅s). In
practice, NaOH is added to premixed HCl and ECA (ClCH2COOCH2CH3).
Depending on the availability on a molecular level, NaOH will react with either
HCl or ECA. When mixing is ideal, there will always be HCl present to react with
NaOH and no conversion of ECA will be obtained. The selectivity for reaction R.1
136 Chapter 7

is then equal to unity. When mixing is moderate, locally all the HCl will be
consumed and the remaining NaOH can react with ECA, resulting in a selectivity
smaller than unity. Thus, the consumed amount of ECA is an indication for the
quality of the mixing process.
The characterization of the macro-mixing performance using this reaction
system has been obtained with the experimental set-up as depicted in Figure 7.2. An
exactly known amount of NaOH is added to a premixed aqueous solution of ECA
and HCl. In Figure 7.3 the percentage of NaOH used for reaction R.2 is given as a
function of impeller speed for reactor11 at a fill ratio of 40%. The percentage of
NaOH used for reaction R.2 increases with decreasing impeller speed. Macro
mixing times as previously presented in Chapters 2 and 3 are also depicted. There
appears to be a strong correlation between both profiles, indicating that macro
mixing determines the selectivity.
Figure 7.4 shows the percentage of NaOH used for reaction R.2 as a function
of fill ratio in reactor11 at an impeller speed of 10 Hz (Re = 3.2⋅105). Also the
macro-mixing times from Chapter 2 and 3 are depicted. The highest ECA
conversion occurs in completely filled reactors, analogous to the macro-mixing
times. The shortest macro-mixing times are obtained at fill ratios of 60-70%. These
do not correspond with the lowest ECA conversion, which has been obtained at a
fill ratio of 40%. Therefore, there seems to be no direct correlation. However, the
injection time of NaOH varied between 0.3 and 1 second over these experiments.
This injection time is relatively large compared to the macro-mixing times.
Therefore, masking effects introduced by the variance in injection time cannot be
ruled out.

7.2.3 Scale-up
In a patent, several reactor designs have been described that provide Twaron®
polymer with the required product quality (Bannenberg-Wiggers et al., 1998). Two
of these designs are selected to show how the production process can be scaled-up
to 2.5 m3 while keeping the macro-mixing times constant. Data are provided in
Table 7.1.
The reaction in example 4A of this patent takes 7 minutes in a liquid volume
of 1.75 m3. With the provided dimensions of the reactor, the macro-mixing times
tm,t and tm,l for turbulent and laminar conditions can be calculated using equations
4.7 and 4.9, respectively. When this process is scaled up to a liquid volume of 2.5
m3 according to route 1 in Table 7.1, while keeping x, N and L/D constant, this
leads to a larger diameter (1.55 m) but to comparable mixing times. However, the
required power inputs would double to 24 and 124 kW for turbulent and laminar
conditions, respectively. The reactor can also be scaled-up to a liquid volume of 2.5
m3 by keeping the diameter D constant, aiming at a reduced power input because
Pt~D5. Then the reactor length has to increase to 4 m, and consequently the impeller
speed to 3.2 Hz in order to keep the mixing times constant. The resulting power
inputs of 28 and 136 kW for turbulent and laminar conditions, respectively, are
even larger than for the first route.
Future perspectives for process engineering of polycondensation reactions 137

The reaction in example 2C of the patent takes 30 minutes in a liquid volume


of 1.3 m3. Chemical conditions differ from example 4A for which the reader is
referred to Bannenberg-Wiggers et al. (1998). Scale-up of this process to 2.5 m3
while maintaining x and L/D = 2 requires a diameter of 1.72 m and a slight increase
in impeller speed to 2 Hz. The required power has more than quadrupled to 15 and
124 kW, respectively. Based on these findings it can be concluded that the
bottleneck for scale-up will probably be the enormous power input under laminar
conditions, required to keep mixing times constant.

Table 7.1: Typical numbers for three different strategies for scale-up of the Twaron® Process to
2.5m3.
Vl D L L/D N tm,t tm,l(1) εt εl Pt Pl(2)
Conditions
[m3] [m] [m] [-] [Hz] [s] [s] [m2/s3] [kW]

Original 1.75 1.37 2.73 2 2.5 27 240 6.4 37 11 65


[PPT] = 11.2 wt%
[H2O] = 675 ppm

Scale
2.5 1.55 3 2 2.5 27 270 8.5 50 21 124
Example 4A

-up 1
tr = 7 min.

x = 0.437

Scale
2.5 1.37 4 3 3.2 28 280 11 53 28 136
-up 2

Original 1.3 1.38 1.66 1.1 1.7 30 250 2.3 18 4 24


[PPT] = 12.4 wt%
[H2O] = 400 ppm
Example 2C
tr = 30 min.

x = 0.52

Scale
2.5 1.72 2.07 1.1 2 28 270 6.2 50 15 124
-up 3

1
The equation for the mixing time is only valid for x = 1.0;
2
The power input at laminar conditions is based on equation 4.8 at Re=100 with k4 = x⋅200.
138 Chapter 7

7.3 Design of a pervaporation membrane reactor for


polycondensation processes
The secondary flow pattern, which has been calculated using CFD and has
been demonstrated using U-CT, provides an efficient route to reduce concentration
polarization in the model system studied. Although it might be somewhat premature
to design a membrane reactor based on the work described in this thesis, there are
some thoughts on reactor designs in which advantage is taken from the secondary
flows as observed in the membrane module. The next paragraph will elaborate on
these reactor designs. It is expected that it will be difficult to control the residence
time distribution in membrane reactors because of the possible presence of
shortcuts. An experimental technique that can be used to study these shortcuts is
MRI (Magnetic Resonance Imaging). The shortcuts can be determined by
measuring velocities using Pulse Field Gradients techniques in MRI (Neling et al.,
1997; Gibss et al., 1997). They can also be studied by measuring proton densities
(Hornak, 2002). This has been done as will be discussed in paragraph 7.3.2. Some
preliminary results will be presented.

7.3.1 Module design


First, a process has to be defined for which a membrane module is required.
The object process scheme has been given in Figure 1.3c: a once-through
continuous process in which a low molecular weight alkyd resin (Mw = 1,000) is
produced with an average polymerization degree of 10. From a production capacity
of 10,000 ton/year, it can be calculated that around 1,600 ton/year of water has to be

Products

External
loops

Reactants Heat exchangers


Figure 7.5: Proposed geometry (1) for a membrane reactor
for low molecular resins operated at laminar conditions.
Future perspectives for process engineering of polycondensation reactions 139

removed. Assuming an average water flux of 4


kg/(m2⋅h), a membrane surface area of 50 m2 is
required. Using a packing density equivalent to
tube-and-shell heat exchangers of 100 m2/m3, this
leads to a reactor volume of approximately 0.5 m3.
i=1
An example of a possible reactor is given in
Figure 7.5, consisting of a horizontal bundle of
membrane tubes in a larger manifold. Along these
membranes the reactants are fed in parallel. Due to
evaporation of the water at the membrane surface, Pn = 1
temperature segregation occurs. The cold and thus
heavy liquid flows to the bottom of the module
where it is removed from the module. After
∆H
heating, it is fed back to the top of the reactor. The
liquid is removed and heated at several axial
positions along the tube to compensate the energy
loss caused by evaporation.
The most difficult part in the design will be to
maintain plug flow behavior, which is mandatory
i=2
to obtain a narrow MWD. Using the relatively
small buoyancy forces to force the liquid in the
right direction might result in unsteady flow
behavior, and shortcuts for reactants directly to the
outlet are likely to occur. Particle tracking in CFD Pn = 2
calculations or MRI (see below) can give insight
into these unwanted events, thus providing means
to prevent shortcuts.
A possibility to control the average degree of
polymerization is by removing cold fluid at the
bottom where the polycondensation is in equili-
brium. Then, just enough energy has to be added in
the recycle for the evaporation of water that is
released by a next propagation step. Theoretically,
i=n
a reactor with 10 recycles can then provide an
average degree of polymerization of 10. For such a
process a set-up of several reactors in series is
presented in Figure 7.6.
Pn = n

Figure 7.6: Proposed geometry (2)


for a set-up of several membrane re-
actors in series for the production of
low molecular resins operated at
laminar conditions.
140 Chapter 7

7.3.2 Magnetic Resonance Imaging (MRI)


A possible experimental technique for studying shortcuts is Magnetic
Resonance Imaging (MRI). In a nuclear magnetic resonance experiment the
magnetic moments of the hydrogen nuclei are manipulated by suitably chosen
alternating radio-frequency fields, resulting in a so called spin-echo signal. The
amplitude of the spin-echo signal is proportional to the amount of nuclei excited by
the radio-frequency field. The resonance condition for the nuclei is given by:
γ
f= Bo (7.1)

Here f is the frequency of the alternating radio-frequency field, γ is the
gyromagnetic ratio (γ /2π = 42.58 MHz/T for 1H), and Bo is the externally applied
static magnetic field. Because of the resonance condition, the method can be made
sensitive to hydrogen only. When also a known magnetic field gradient is applied,
the resonance condition and hence the NMR signal will depend on the position of
the nuclei. By applying the magnetic field gradients along a large number of
different directions and using back-projecting techniques, analogous to the methods
described in Chapter 6, the proton density can be made visible in a 2-dimensional
plot (Kak, 1984).
The spin-echo signal also gives information about the rate at which the
excitation of the magnetic moments decays. The system will return to equilibrium
by two mechanisms: interactions between the nuclei themselves, causing the so-
called spin-spin relaxation, and interactions between the nuclei and their environ-

Figure 7.7A-B: A) Relative NMR signal strength as a function of repetition time. The
percentage in the legend refers to the amount of 1,4-butanediol. T1 is the spin lattice
relaxation time determined from fits of an exponential function to the data. B) T1 as a
function of the 1,4-butanediol mass fraction (Van der Sande, 2000).
Future perspectives for process engineering of polycondensation reactions 141

ment, causing the so-called spin-lattice


relaxation. w% H2O
Assuming that both mechanisms
give rise to a single exponential
relaxation and that the spin-lattice relax-
ation is much slower than the spin-spin
relaxation, the magnitude of the NMR
spin-echo signal is given by
(Vlaardingerbroek et al., 1996):

S = ρ e −TE / T2 (1 − e −TR / T1 ) (7.2)

In this expression, ρ is the density of the


hydrogen nuclei, T1 the spin-lattice or
longitudinal relaxation time, TR the
repetition time of the spin-echo experi-
ments, T2 the spin-spin or transverse
relaxation time, and TE the so-called
spin-echo time. T1 and T2 depend on the
nuclei surrounding the hydrogen nuclei Figure 7.8: Mass fraction distribution of
at resonance, and hence the T1 and T2 of water in a 67/33 w% mixture of 1,4-
water and 1,4-butanediol will be butanediol and water along a pervapo-
different. In this study, it was chosen to ration membrane at x = 0.85m.
relate T1 of the mixture to the 1,4-
butanediol mass fraction. If TE << T2, the T1 can be determined by changing the TR
of the spin-echo experiment. In Figure 7.7A the signal is given as a function of TR
for various 1,4-butanediol mass fractions. As can be seen the T1 of the mixture
clearly depends on the 1,4-butanediol mass fraction. In Figure 7.7B the calibration
of the T1 of the mixture as a function of the 1,4-butanediol mass fraction is given.
To illustrate the power of MRI, a lot of effort has been made to experimentally
show the presence of concentration polarization in the membrane module. For such
a MRI experiment, a 67/33 wt% mixture of 1,4-butanediol and water has been
chosen. This mixture combines a moderate viscosity (0.015 Pa⋅s) and moderate
water fluxes: at an inlet temperature of 70-80°C and 300 mL/min the flux is
between 1 and 1.5 kg/m2⋅h (Chapter 5 and 6; Van der Gulik et al., 2001). A
dedicated CFD calculation has been performed for these conditions to show the
polarization effects that could be expected. The calculated concentration
distribution in mass fractions at the end of the membrane (at x = 0.85 m of a
membrane 0.9 m long) is plotted in Figure 7.8. Water and 1,4-butanediol have
approximately the same density as a function of temperature (Sun et al., 1992,
Hawrylak et al., 1998). Therefore, the buoyancy effects are not profound, resulting
in a strong concentration polarization, which is helpful in MRI as a large
concentration difference is easier to measure.
A special MRI set-up has been built for measuring concentration polarization
in the membrane module, which is depicted in Figure 7.9 (Blok, 1999; Van der
Sande, 2000). This MRI set-up uses an electromagnet generating a magnetic field of
142 Chapter 7

Main magnet

Membrane
module
Heating unit
for liquid
Transport
unit

MRI Electronics

Figure 7.9: Setup of the MRI equipment.

0.4 Tesla. Unfortunately, with this equipment no concentration polarization could


be demonstrated, since the noise levels were too high and the spatial resolution of
the MRI equipment was too low. The mediocre spatial resolution was demonstrated
by analyzing the signal strength over a line from the liquid bulk to the center of the
membrane. A very sharp transition in proton density exists across the membrane
surface going from unity in the bulk liquid (high proton density) to zero in the
membrane tube (low proton density). In a 2-dimensional representation of the
measured proton density, 4 mm is required to span the decrease from unity to zero.
This length is too much to visualize the 0.75-mm thick boundary layer in Figure 7.8
obtained with CFD calculations. The lack of spatial resolution was mainly a result
of inhomogeneity of the main magnetic field. The best result obtained is depicted in
Figure 7.10 in which a measured concentration distribution is given at conditions
that apply for the distribution given in Figure 7.8. The off-circle shape is a result of
inhomogeneity of the main magnetic field. A slightly higher 1,4-butanediol
concentration (+3 w%) can be observed in red below the membrane.
Measuring concentration polarization requires MRI equipment with more
restrictive characteristics. Currently, a new 1.5 Tesla MRI machine has been
installed, providing a much more homogeneous magnetic field, in which
concentration polarization might be measurable. However, with the current set-up
shortcuts in membrane modules might still be measurable because in that case the
bulk of the liquid will be monitored, which requires less spatial resolution.
Future perspectives for process engineering of polycondensation reactions 143

Bulk liquid

Membrane tube

Increased
concentration

Figure 7.10: Measured concentration distribution around membrane tube in the horizontal
membrane module. Green is the bulk liquid, orange indicates an increased concentration.
144 Chapter 7

7.4 Outlook to the future


From the work presented in this thesis it can be concluded that hydrodynamics
have a major impact on the design and development of reactors for
polycondensation reactions. A major challenge for the future will be to combine the
insights in hydrodynamics with kinetic reaction schemes. This should allow for the
incorporation of all hydrodynamic characteristics and evaluation of their effect on
final product properties like the molecular weight distribution. Obviously, CFD
packages should eventually be able to handle this type of complex question. This
should allow for a proper comparison between different reactor geometries and
modes of operation, leading to the development of reliable and sustainable
processes for the future.

References
Blok, R.; “Design and validation of a 3d MRI setup to study chemical separation processes”,
M.Sc. Thesis, FIK/FTI99-05, Eindhoven University of Technology, 1999.
Bourne, J.R.; Yu, S.; “Investigation of micromixing in stirred tank reactors using parallel
reactions”, Ind. Eng. Chem. Res., 1994, 33(1), 41-55.
Bannenberg-Wiggers, A.E.M.; Van Omme, J.A.; Surquin, J.M.; “Process for the batchwise
preparation of poly-p-terephtalamide”, U.S. Pat., 5,726,275, 1998.
Gibbs, S.J.; Haycock, D.E.; Frith, W.J.; Ablett, S.; Hall, L.D.; “Strategies for rapid NMR
Rheometry, by magnetic resonance imaging velocimetry”, J. Magn. Reson., 1997, 125,
43-51.
Hawrylak, B.; Gracie, K.; Palepu, R.; “Thermodynamic properties of binary mixtures of
butanediols with water”, J. Solution Chem., 1998, 27(1), 17-31.
Hornak, J.P.; “The basics of MRI”, http://www.cis.rit.edu/class/schp730/bmri/bmri.htm.
Kak, A.C.; “Image reconstruction from projections”, Digital image processing techniques,
Academic Press Inc., ISBN 0-12-236760-X, 1984.
Neling, B.; Gibbs, S.J.; Derbyshire, J.A.; Xing, D.; Hall, L.D.; Haycock, D.E.; Firth, W.J.;
Ablett, S.; “Comparisons of magnetic imaging velocimetry with computational fluid
dynamics”, J. Fluids Eng., 1997, 119, 103-109.
Schoenmakers, J.H.A.; “Turbulent feed stream mixing in agitated vessels”, Ph.D. Thesis,
Eindhoven University of Technology, 1998.
Sun, T.; DIGuillo, M.; Teja, A.S.; “Densities and viscosities of four butanediols between 293
and 463 K”, J. Chem. Eng. Data, 1992, 37, 246-248.
Van der Gulik, G.J.S.; Janssen, R.E.G.; Wijers, J.G.; Keurentjes, J.T.F.; “Hydrodyna-mics in
a ceramic pervaporation membrane reactor for resin production”, Chem. Eng. Sci., 2001,
56, 371-379.
Van der Sande, J.J.; “Design and validation of a NMR-measurement system”, M.Sc. Thesis,
NF/FIK 2000-07, Eindhoven University of Technology, 2000.
Verschuren, I.L.M.; “Feed stream mixing in stirred tank reactors”, Ph.D. Thesis, Eindhoven
University of Technology, 2001.
Vlaardingerbroek, M.T.; Den Boer, J.A.; “Magnetic Resonance Imaging”, Springer-Verlag,
Germany, ISBN 3-540-60080-0, 1996.
DANKWOORD

Hoewel dit dankwoord voor velen waarschijnlijk het meest begrijpelijke deel
van dit proefschrift is, was het misschien wel het moeilijkste deel om te schrijven.
Is het bij het schrijven van een wetenschappelijk getinte tekst vaak al moeilijk de
juiste afbakening te vinden, bij een dankwoord is dit welhaast onmogelijk. Vandaar
dat ik allereerst wil beginnen met iedereen, maar dan ook íedereen, te bedanken, die
op welke wijze dan ook geholpen heeft bij de totstandkoming van dit boekwerk.
Toch wil ik een aantal mensen bij naam bedanken.
Te beginnen met mijn promotor Jos Keurentjes en mijn coach Johan Wijers.
Jos, ik bewonder je onuitputtelijke enthousiasme dat ook nog eens aanstekelijk
werkt. Daarmee heb je mij over de streep getrokken om te gaan promoveren. Ik wil
je bedanken voor het vertrouwen, de begeleiding en de bijzonder vlotte afhandeling
van de stapels papier. Johan wil ik graag bedanken voor zijn grote betrokkenheid,
de open deur en zijn humoristische doch kritische blik op de alledaagse dingen.
Tenslotte is het mij opgevallen dat jullie opmerkingen t.a.v. een wetenschappelijke
tekst volledig complementair zijn. Soms leek het afgesproken werk.
De sponsoren van het verrichtte werk wil ik bedanken voor de financiële steun.
Hoewel vaak op grote afstand, wil ik Jan Surquin speciaal bedanken, ook gezien
zijn voortdurende interesse in het onderzoek. Ook is het prettig te constateren dat
sponsoren bereid zijn het vervolgonderzoek van 8 manjaar mee te willen
financieren.
Dit proefschrift staat bol van het experimentele werk. Voor het vervaardigen
van de opstellingen heb ik waarschijnlijk iedereen van de Technische Dienst wel
eens ingeschakeld. Zonder de illusie te hebben volledig te zijn, wil ik in elk geval
de volgende mensen bedanken: Peter Brinkgreve, Anton Bombeeck, Erwin
Dekkers, Piet van Eeten, Henk Hermans, Rinus Janssen, Frans Kuijpers, Theo
Maas, Jovita Moerel en Hans Wijtvliet.
Baukje Osinga wil ik bedanken voor het uitvoeren van een groot aantal
pervaporatie-experimenten en het veelvuldig verslepen van de opstelling van de
Faculteit Scheikundige Technologie naar Technische Natuurkunde v.v. Daar
werden we altijd hartelijk ontvangen door Leo Pel, Roland Blok en/of Joris van der
Sande van de groep Fysische informatica en klinische fysica (FIK). Een speciaal
woord van dank gaat uit naar hen. Met grote inzet en doorzettingsvermogen hebben
zij een complete MRI-opstelling gebouwd om aan concentratiepolarisatie in de
membraanmodule te kunnen meten. Het is nauwelijks voor te stellen dat uit zoveel
noeste arbeid, zo weinig bevredigend resultaat kan voortkomen. Gelukkig kunnen
jullie de opstelling ook nog voor andere doeleinden gebruiken.
Het bedrijf Innovation Handling verdient zeker een bedankje voor de
assistentie bij het berekenen van temperatuurverdelingen zoals die ontstaan tijdens
de pervaporatie-experimenten. Ook al is onze samenwerking niet gelopen zoals
gepland, jullie reconstructiesoftware is uit de kunst!
Ik heb het genoegen gehad een vijftal afstudeerders te mogen begeleiden. Zij
hebben veel werk verricht waarvan het grootste gedeelte is opgenomen in dit
proefschrift. Achtereenvolgens waren dit: René van der Zande (de man met de
bekertjes glycerol), Eric Rossou (voor de mooiste PLIF-plaatjes), Roger Janssen
(“Het personeel”), Hanny van Enschot (chaotisch mengen) en Dirk van Asseldonk
(de duizendpoot). Allen bedankt voor jullie inzet, de goede samenwerking en de
gezelligheid, ook na vijven.
Alle leden van de capaciteitsgroep Procesontwikkeling bedank ik voor de
plezierige werksfeer, in het bijzonder alle kamergenoten die ik in de loop der tijd als
promovendus heb versleten, te weten Aico de Volder, Geurt Swanenberg, Maartje
Kemmere, Jaco Boelhouwer en Marc Jacobs.
Twee heren zullen 10 december nog een beetje op (me) moeten letten, Marc en
Peter, de beide paranimfen. Bedankt dat jullie deze belangrijke taak willen
uitvoeren. Kunnen jullie meteen een beetje wennen aan het podium…
Tenslotte wil ik alle vrienden en familie bedanken voor de steun en de
getoonde interesse aangaande de voortgang van het boekje, in het bijzonder
natuurlijk Pa, Ma, Pieter en Floor. Uiteraard wil ik Arina en de kinderen bedanken.
Arina, het zal je nog tot vervelens toe gezegd worden dat dit niet mogelijk was
zonder jouw liefde, steun en ruimte. Maar dat klopt dan ook als een bus. Joanne,
Huub en Mieke, jullie wil ik bedanken voor het opeisen van tijd voor ontspanning.
Jullie geven mij het inzicht om alles op de juiste waarde in te schatten.

Ommen, Oktober 2002


CURRICULUM VITAE

Gert-Jan van der Gulik werd op 3 juni 1969 geboren te Uithuizen. In 1986
behaalde hij het HAVO diploma aan het Ludgercollege te Doetinchem. Aansluitend
studeerde hij Chemische Technologie aan de Hogeschool Enschede. Na een jaar
onder de wapenen te hebben gestaan, werd de studie Chemische Technologie
vervolgd aan de Universiteit Twente. In 1995 studeerde hij af onder leiding van
prof.dr.ir. J.A. Lercher in de Vakgroep Katalytische Processen en Materialen. In
1996 begon hij aan de 2de-fase opleiding Proces- en productontwerp aan de
Technische Universiteit Eindhoven. De eindopdracht omvatte het opschalen van de
Drais reactor waarin het polymeer voor de Twaron-vezel wordt geproduceerd. Deze
opdracht is vervolgens 2,5 jaar voortgezet en uitgebreid met onderzoek naar de
hydrodynamica in pervaporatie membraanmodules. De resultaten van beide zijn
beschreven in dit proefschrift. Het onderzoek is uitgevoerd binnen de
capaciteitsgroep Procesontwikkeling onder supervisie van prof.dr.ir. J.T.F.
Keurentjes.
Sinds 1 november 2000 is Gert-Jan van der Gulik werkzaam als CFD-
consultant bij BuNova Development BV te Zwolle.

You might also like